Skip to main content
Advertisement
Browse Subject Areas
?

Click through the PLOS taxonomy to find articles in your field.

For more information about PLOS Subject Areas, click here.

  • Loading metrics

Microarray Analysis of the Ler Regulon in Enteropathogenic and Enterohaemorrhagic Escherichia coli Strains

  • Lewis E. H. Bingle ,

    Contributed equally to this work with: Lewis E. H. Bingle, Chrystala Constantinidou

    Current address: Faculty of Applied Sciences, University of Sunderland, Sunderland, United Kingdom,

    Affiliation School of Biosciences, University of Birmingham, Edgbaston, Birmingham, United Kingdom

  • Chrystala Constantinidou ,

    Contributed equally to this work with: Lewis E. H. Bingle, Chrystala Constantinidou

    Current address: Division of Microbiology & Infection, University of Warwick, Coventry, United Kingdom,

    Affiliation School of Biosciences, University of Birmingham, Edgbaston, Birmingham, United Kingdom

  • Robert K. Shaw,

    Affiliation School of Biosciences, University of Birmingham, Edgbaston, Birmingham, United Kingdom

  • Md. Shahidul Islam,

    Current address: Department of Biotechnology, Bangladesh Agricultural University, Mymensingh-2202, Bangladesh,

    Affiliation School of Biosciences, University of Birmingham, Edgbaston, Birmingham, United Kingdom

  • Mala Patel,

    Affiliation School of Biosciences, University of Birmingham, Edgbaston, Birmingham, United Kingdom

  • Lori A. S. Snyder,

    Current address: School of Life Sciences, Kingston University, Kingston upon Thames, United Kingdom

    Affiliation School of Biosciences, University of Birmingham, Edgbaston, Birmingham, United Kingdom

  • David J. Lee,

    Affiliation School of Biosciences, University of Birmingham, Edgbaston, Birmingham, United Kingdom

  • Charles W. Penn,

    Affiliation School of Biosciences, University of Birmingham, Edgbaston, Birmingham, United Kingdom

  • Stephen J. W. Busby,

    Affiliation School of Biosciences, University of Birmingham, Edgbaston, Birmingham, United Kingdom

  • Mark J. Pallen

    m.pallen@warwick.ac.uk

    Current address: Division of Microbiology & Infection, University of Warwick, Coventry, United Kingdom,

    Affiliation School of Biosciences, University of Birmingham, Edgbaston, Birmingham, United Kingdom

Abstract

The type III protein secretion system is an important pathogenicity factor of enteropathogenic and enterohaemorrhagic Escherichia coli pathotypes. The genes encoding this apparatus are located on a pathogenicity island (the locus of enterocyte effacement) and are transcriptionally activated by the master regulator Ler. In each pathotype Ler is also known to regulate genes located elsewhere on the chromosome, but the full extent of the Ler regulon is unclear, especially for enteropathogenic E. coli. The Ler regulon was defined for two strains of E. coli: E2348/69 (enteropathogenic) and EDL933 (enterohaemorrhagic) in mid and late log phases of growth by DNA microarray analysis of the transcriptomes of wild-type and ler mutant versions of each strain. In both strains the Ler regulon is focused on the locus of enterocyte effacement – all major transcriptional units of which are activated by Ler, with the sole exception of the LEE1 operon during mid-log phase growth in E2348/69. However, the Ler regulon does extend more widely and also includes unlinked pathogenicity genes: in E2348/69 more than 50 genes outside of this locus were regulated, including a number of known or potential pathogenicity determinants; in EDL933 only 4 extra-LEE genes, again including known pathogenicity factors, were activated. In E2348/69, where the Ler regulon is clearly growth phase dependent, a number of genes including the plasmid-encoded regulator operon perABC, were found to be negatively regulated by Ler. Negative regulation by Ler of PerC, itself a positive regulator of the ler promoter, suggests a negative feedback loop involving these proteins.

Introduction

Enteropathogenic (EPEC) and enterohaemorrhagic (EHEC) Escherichia coli are two pathotypes of this important gastrointestinal bacterium that can cause serious diarrhoeal disease in humans [1]. Many EHEC and EPEC strains possess a type III secretion system (T3SS) encoded by a pathogenicity island called the locus of enterocyte effacement (LEE) that is also found in the related bacterium Citrobacter rodentium, a mouse pathogen that is widely used as a model for the EHEC and EPEC strains [2]. Pathogenicity factors encoded within the LEE, specifically the type III secretion system and secreted effector proteins, are responsible for formation of the attaching and effacing (AE) lesion on the gut epithelium that is characteristic of these strains and required for intimate attachment of the bacteria [3]. The 41 genes of the LEE are arranged in 5 major polycistronic operons called LEE1-5 along with a number of smaller transcriptional units [4]. Attaching and effacing pathogens, including EPEC strains such as E2348/69, O157:H7 EHEC strains and non-0157 EHEC strains, have distinct evolutionary histories but carry an overlapping core repertoire of pathogenicity genes, including the LEE and many effector genes outside the LEE, that have been acquired via horizontal gene transfer [5], [6], [7]. However, there are significant differences in overall pathogenicity between EHEC and EPEC strains, for example EHEC strains cause a more severe bloody diarrheal disease (haemorrhagic colitis) that is often accompanied by the life threatening complication, haemolytic uraemic syndrome (HUS) [8]. Such differences are presumably mainly determined by the differing contributions of the extra-LEE factors. Examples include differing arrays of T3SS effector proteins and the fact that the EHEC genome encodes a Shiga-like toxin responsible for serious pathology in the human host, while EPEC does not [8].

In addition to variation in the genomic arsenal of determinants, appropriate control of gene transcription may be critical in optimising pathogenicity [9], [10]. Type III secretion systems (T3SS) are generally acquired through horizontal gene transfer and therefore should employ a means of regulation that is easily integrated into the existing regulatory networks of the cell [11]. One way to achieve this integration is to have T3SS gene expression under the control of a master regulator, which multiple environmental signaling pathways can feed into. The master regulator for the LEE is the Ler protein, encoded by the first gene in the LEE1 operon [12]. Ler is a transcriptional activator of the LEE: a homologue and also an antagonist of the genome organizer and silencer H-NS [13]. In addition to its H-NS-dependent role in activating most promoters of the LEE, Ler can activate the LEE5 promoter in an H-NS-independent manner (reviewed in [14]). Ler has also previously been shown to act as a specific autorepressor of the LEE1 promoter [15] while the LEE encoded regulator GrlA and the plasmid encoded regulator PerC (EPEC), or its EHEC homologues PchABC, have been shown to specifically activate ler transcription [13], [16], [17], [18], [19]. LEE gene expression is responsive to population status, via the AI-3 quorum sensing system activating the LysR type regulator QseA which in turn activates LEE1 (ler) transcription [20], [21], [22]. Expression of the LEE is also known to be responsive to many environmental factors (reviewed in [23], [24]). One example is temperature: transcription of the LEE is up-regulated at 37°C and repressed (by H-NS) at 27°C [25]. Expression of the LEE is also dependent on the physiological state of the cell, for example growth phase. In glucose MOPS minimal medium, gene expression as assessed by microarray transcriptomics is maximal in late exponential phase and down-regulated during the transition to stationary phase [26]. Under some other growth conditions (LB broth) expression from LEE promoters, measured via transcription of a lacZ reporter gene, seemed to increase during the transition to stationary phase [20], [27].

Extra-LEE genes that are known to be members of the Ler regulon in EPEC include espC, encoding an autotransporter (Type V) extracellular serine protease, that is thought to play various roles in pathogenicity [28], [29], [30], [31]. The espC gene has previously been shown to be strongly activated by Ler, however in contrast the EHEC homologue of this gene espP was not found to be Ler-regulated [12]. Extra-LEE members of the Ler regulon in EHEC include stcE, a pO157–borne gene encoding a metalloprotease that is involved in intimate adherence of bacterium to gut epithelium [32] and nleA, encoding a T3SS-secreted effector protein [33]. However some of the many extra-LEE T3SS effectors of EHEC were previously thought not to be regulated by Ler e.g. EspJ and TccP [6], [34]. In addition, expression of long polar fimbriae of EHEC has been found to be reciprocally regulated by H-NS repression and Ler antagonism [35].

Here we will characterise and compare the Ler regulons for EPEC strain E2348/69 and EHEC strain EDL933. The regulon for Ler has previously been loosely defined at the transcriptional level for the closely-related Sakai strain of EHEC [36] where Ler regulation was mostly found to be confined to horizontally-transferred DNA. The LEE is inserted at the same selC locus in both EDL933 and E2348/69 strains, the most parsimonious interpretation being a single insertion event in a common ancestral strain [37]. Any differences in the Ler regulon between these two strains, within or outside the LEE, will reflect divergent adaptation to subsequent changes in the genome, for example plasmid acquisition, and are of interest from a regulon evolution point of view.

Results

We constructed two validated mutant strains of E. coli: LBEC1 (EDL933 Δler) and LBEC2 (E2348/69 Δler), grew cultures of WT and parental mutant strains under conditions known to be inducing for the LEE to two different growth phases (mid and late log phase), harvested RNA and used this to perform microarray analysis of the transcriptomes. Microarray data has been deposited with the GEO database (http://www.ncbi.nlm.nih.gov/geo) with accession code GSE38876.

Enteropathogenic E. coli

In mid-log phase cells a total of 85 genes are transcriptionally regulated: 62 genes, at 14 different loci, are activated by Ler (table 1) while 23 genes, at 6 loci, are repressed by Ler (table 2). Of the activated genes 49 (79%) are carried on or directly adjacent to mobile genetic elements (MGEs: prophage, integrative element or plasmid), while 11 of the repressed genes (48%) are carried on MGEs. If one compares the genes that are activated and repressed, two repressed genes (E2348C_0084 and E2348C_2114) are potentially expressed from promoters that are immediately divergent from an activated promoter.

thumbnail
Table 1. EPEC genes activated 2-fold or more by Ler at mid-log phase (OD600 = 0.4).

https://doi.org/10.1371/journal.pone.0080160.t001

thumbnail
Table 2. EPEC genes repressed 2-fold or more by Ler at mid-log phase (OD600 = 0.4).

https://doi.org/10.1371/journal.pone.0080160.t002

In late log phase cells 97 genes in total are regulated by Ler. Of these, 85 genes at 23 genetic locations are activated, of which 62 genes (73%) are carried on or directly adjacent to MGEs (table 3). Twelve genes are repressed by Ler, of which only 1 is adjacent to a MGE (table 4).

thumbnail
Table 3. EPEC genes activated 2-fold or more by Ler at late-log phase (OD600 = 0.9).

https://doi.org/10.1371/journal.pone.0080160.t003

thumbnail
Table 4. EPEC genes repressed 2-fold or more by Ler at late-log phase (OD600 = 0.9).

https://doi.org/10.1371/journal.pone.0080160.t004

The strongest activation was generally observed for LEE genes, with the mostly high activated genes being eae at mid-log phase (58-fold) and orf29 in late log phase (100-fold). Extra LEE genes with comparable levels of activation included espC (mid-log only), pagP and the gene encoding the T3SS secreted effector NleA. The maximum fold repression observed outside of the LEE was approximately 9-fold in mid-log phase (fimD) and approximately 6-fold in late log phase cells (chuT-hmuV heme utilization operon).

Enterohaemorrhagic E. coli

In mid-log phase cells, only one gene passed the Benjamini and Hochberg MTC filter as being repressed (2-fold) by Ler. This was Z2974 on prophage CP-933T, encoding an unknown protein.

In late-log phase cells, 39 genes were found to be transcriptionally activated by Ler (2-fold or more; table 5). Thirty five of these genes are within the LEE (representing all major transcriptional units; activation between 4 and 32-fold). The remaining 4 extra-LEE activated genes encode: StcE (4-fold), EtpC (3-fold), SfpA (5-fold) and the putative cytochrome YhaI (36-fold). The stcE and etpC genes are located on plasmid pO157; SfpA is prophage-encoded and yhaI is not associated with a mobile genetic element.

thumbnail
Table 5. EHEC genes activated 2-fold or more by Ler at late-log phase (OD600 = 1.1).

https://doi.org/10.1371/journal.pone.0080160.t005

Discussion

It is clear that in both the EPEC and EHEC strains of E. coli examined here, the LEE is the primary target for Ler activation: all major transcriptional units of the LEE are regulated by Ler, although the regulation of LEE1 is growth phase dependent in EPEC, as noted below. Otherwise, in EPEC the Ler regulon is quite small, covering about 2% of the genome; in EHEC the regulon is even smaller and contains very few genes outside of the LEE. As the positive regulatory activity of Ler is known to be due to antagonism of H-NS repression (where studied) we would predict that all activated members of the Ler regulon are repressed by H-NS. However the H-NS regulon is very large and clearly not all H-NS repressed genes are activated by Ler [38]. An important question that therefore remains to be answered is: what provides specificity to Ler regulation? The specificity of action that we have observed (i.e. most of the strongly regulated genes are located within the LEE) is in agreement with the observations of Abe et al. relating to EHEC [36]. This specificity is consistent with Ler binding to a specific DNA structural motif, via an indirect readout mechanism, as suggested by an NMR analysis of Ler C-terminal domain-DNA complexes [39]. While some studies have shown evidence for specific binding of LEE promoters by Ler, the Chip-CHIP analysis of Abe et al. suggested that Ler was binding extensively (although not evenly) across the Sakai genome and the authors concluded that Ler has a low binding specificity [36].

Many but not all of the extra LEE members of the EPEC Ler regulon are located on mobile genetic elements (MGEs) and it is particularly striking that Ler negatively regulates a disproportionately high number of plasmid-borne genes, at least in mid-log phase EPEC: 9 genes from 3 different operons (10% of the total of 90 genes) on plasmid pMAR2 are shown to be regulated, while only 0.3% of the chromosomal genes (14 genes) are repressed. However by late log phase, no plasmid-borne genes are repressed by Ler. Similarly it is striking that 4 of the 5 extra-LEE genes found to be Ler-regulated in EHEC (likely members of the same operon) are located on a MGE (plasmid or prophage). In both bacteria the GC contents of chromosome is slightly higher than that of the plasmid: EDL933 and E2348/69 chromosomes are 50.4 and 50.6% respectively, while pO157 and pMAR2 are 47.6% and 48%. As genes located on MGEs with lower GC content are selectively silenced by H-NS [40], in evolutionary terms Ler activation could have been a useful means to “liberate” the expression of newly acquired pathogenicity factors.

Across the genome, the Ler regulon is notably growth phase dependent in EPEC: in mid log phase (OD600 = 0.4) 27 extra-LEE genes are activated by Ler, while in late log phase (OD600 = 0.9) the number of activated extra-LEE genes is 43. In EPEC the regulation of the LEE1 operon, but not the other operons of the LEE, differs between mid-log and late-log growth phases: at late log phase, all 41 genes within the LEE are strongly activated by Ler, along with the flanking predicted sugar transporter gene yicJ, while at mid-log phase the 7 genes in the LEE1 operon before escU are not strongly (>2-fold) regulated (Figure 1; note that we do not comment on the regulation of the ler gene itself as the coding sequence is partly deleted in the mutant). While it is possible that we have introduced some artefactual corruption of LEE1 regulation during mutation of ler, the observed activation of late-log phase cells suggests that there is no gross defect in the Ler regulatory circuit. This result indicates that the regulation of the LEE1 promoter is somewhat different to that of other LEE promoters, possibly due to a complex balance between Ler autoregulation and activation. It is noteworthy that, while previous reporter gene analysis of the LEE1 promoter has indicated that it is autorepressed by Ler, our results indicate that it may be activated, a difference that may reflect the growth phase dependence of the effects observed here [15]. No corresponding differential regulation of the LEE1 operon was observed in late log phase EHEC; in the mid-log phase cultures none of the LEE genes passed the MTC filter, but if the filter is not applied then LEE1 seems to be similarly regulated in mid-log and late-log phase cultures. While Sperandio et al. found that the LEE4 operon (sepL-espF) was constitutively expressed at a high level in EHEC and insensitive to Ler regulation [20], we have found it to be clearly Ler-dependent in both EHEC and EPEC strains. The observed difference could have resulted from selection of a promoter fragment for reporter gene assays that lacks the full complement of H-NS binding sites.

thumbnail
Figure 1. Growth phase dependent Ler regulation of the EPEC LEE1 operon.

While the other major operons of the EPEC LEE are regulated by Ler in a similar manner in both mid and late-log phase cultures, the LEE1 operon (lerescU) was strongly activated by Ler only in late log phase cultures. The LEE1 operon and flanking genes are shown as block arrows and are coloured according to the fold activation seen in ler+ cells. Fold activation values are not shown for the ler gene as this is partly deleted in ler cells. The intergenic region between ler and espG has been contracted for clarity.

https://doi.org/10.1371/journal.pone.0080160.g001

There are a number of Ler-activated genes in the EPEC regulon that are outside of the LEE but may be involved in pathogenicity. As noted above, espC is already known to be ler-regulated and is one of the mostly highly (22-fold) activated genes in mid-log phase cells. PagP, the palmitoyl transferase for lipid A is strongly regulated at both mid and late log phases (9-fold and 15-fold respectively). Palmitoylated lipid A supposedly protects bacteria from host immune defences (e.g. CAMPs) and attenuates their activation through the TLR4 signal transduction pathway [41]. E2348C_0684, strongly regulated along with its downstream neighbour, encodes a SfpA (systemic factor protein A)-like protein: SfpA is a porin involved in systemic disease in Yersinia enterocolitica [42]. A homologue of sfpA (ECs0814) in the Sakai strain of EHEC was previously observed to be Ler regulated [36]. The rcsA gene, which encodes a positive regulator of the serotype-specific group I K (capsular) antigen is activated by Ler in late log phase, although not at the earlier growth point [43]. This may reflect an impact of capsule production on the intimate attachment of EPEC bacteria to the gut epithelium, however, no regulation of the wza promoter (target for RcsA in E. coli K-12) was apparent. It is worth noting at this point that a ler mutant of EPEC was previously found to be defective for colonisation of Caenorhabditis elegans [44]. This requirement for Ler was found to be independent of T3SS encoded by the LEE. This effect is presumably due to one or more of these extra-LEE members of the Ler regulon which are essential pathogenicity factors in a C. elegans infection but are not involved in T3S (and are not effectors delivered by the T3SS). Several non-LEE encoded effector genes, whose products are secreted via the T3SS, are Ler-regulated in EPEC, including the operon of five genes from nleI/G to nleF (Ler regulation of a homologue of nleA is already known to occur in EHEC [33]) and a homolog of the espG gene located next to the espC gene which is also Ler regulated (see above). There is also clear evidence for the transcriptional regulation of nleH and espJ homologues at late log phase. While it may be unsurprising that effectors secreted via the T3SS are coregulated with the LEE, previous studies in EHEC and C. rodentium have not found these two genes to be regulated by Ler [34], [45]. Only 4 extra-LEE genes were identified as part of the EHEC Ler regulon: sfpA, as discussed above; stcE, encoding a protease that is known to be involved in intimate adherence and inhibition of complement-mediated lysis [32], [46]; etpC, located immediately downsteam of stcE and encoding a component of the pO157-encoded type II secretion system for StcE is also known to be involved in adherence and intestinal colonization [47] and the putative cytochrome gene yhaI. Assuming that etpC is in the same operon as stcE, only the last of these is a novel observation.

We have also identified a number of EPEC genes that are repressed by Ler, including the “plasmid-encoded regulator” operon perABC, located on the EPEC adherence factor (EAF) plasmid pMAR2 [48]. PerA protein activates transcription of the bfp operon, encoding bundle-forming pili [49]. These pili are involved in formation of an initial attachment between EPEC cells and the gut epithelium that occurs prior to AE lesion formation, therefore down-regulation of bfp expression with LEE expression is consistent with the known program of infection [50]. PerC protein is known to activate ler [17], [18], [51] and therefore this result suggests the existence of a negative feedback loop, previously undescribed, that ultimately autoregulates expression of Ler (and therefore the LEE) and may be involved in a down-regulation of ler transcription after the initial stages of infection [52]. Regulation of the per operon by Ler, the gene for which is known to be regulated by quorum sensing (QS), would account for the previously observed “indirect” QS regulation of perA [20]. The repressive effect of Ler on perA presumably also explains the up-regulation of the bundle-forming pili (bfp) operon in the ler knockout mutant. Neither of these phenomena (which were only observed in mid-log phase cells) have so far been reported in the literature, although Elliot et al. reported Ler regulation of non-BFP fimbriae, while Leverton and Kaper described an inverse relationship between expression of ler and bfpA in the presence of HEp-2 cells [12], [52]. Ler repression of acid resistance genes – previously noted by Abe et al. in the Sakai strain [36] - may reflect an accessory mechanism to assist in tight regulation of these genes, preventing inappropriate expression in the lower regions of the GI tract where acid resistance is not required.

Overall the data reported here suggests that the Ler regulon for enteropathogenic and enterohaemorrhagic strains of E. coli is mainly focused on the type III secretion system genes in the LEE, but also includes unlinked pathogenicity genes. The regulon is growth phase dependent and, at least in strain E2348/69, is composed of both positively and negatively regulated genes. Additionally, in enteropathogenic E. coli, the observed negative regulation by Ler of PerC, itself a positive regulator of the ler promoter, suggests the existence of a negative feedback loop involving these two proteins.

Materials and Methods

Bacterial strains and plasmids

Bacterial strains used or constructed during this study are detailed in table 6 and plasmids used or constructed are detailed in table 7. Standard techniques for recombinant DNA manipulations were used throughout this work. All cloned sequences were checked using the University of Birmingham Functional Genomics Facility (http://www.genomics.bham.ac.uk/sequencing.htm).

thumbnail
Table 6. Bacterial strains used or constructed in this study.

https://doi.org/10.1371/journal.pone.0080160.t006

Construction and validation of E. coli mutant strains and ler expression plasmid

The ler expression plasmid pSI04 was derived from pJW15Δ1-100 [53] by cloning EHEC ler CDS as a NsiI-HindIII fragment under the control of the melR promoter and SD site [54].

Non-polar ler knockout mutants of E. coli strains EDL933 and E2348/69 were constructed using a λRed based method (GeneDoctoring) [55] to replace the majority of the ler gene with a kanamycin resistance cassette. Recombination cassettes designed to replace the central portion of ler with a kanamycin resistance gene (aphA) were amplified from a pDOC-K template using a conserved forwards primer (LER-KO-F: taatagcttaaaatattaaagcATGCGGAGATTATTTATTATGAATATGG-TGGCTGGAGCTGCTTCGAA) in combination with strain specific reverse primers (LER-KO-EHEC-R: catttaattatttcatgTTAAATATTTTTCAGCGGTATTATTTCTTCT-CTCGAGATATGAATATCCTCCTTAG and LER-KO-EPEC-R: catttaattattttatgTTAAATATTTTTCAGCGGTATTATTTCTTCT-CTCGAGATATGAATATCCTCCTTAG) and a proofreading DNA polymerase (Velocity, Bioline). The blunt-ended cassettes were ligated into EcoRV-digested donor plasmid pDOC-C to generate donor plasmids carrying EHEC and EPEC specific leraphA+ knockout cassettes. These plasmids were used together with pACBSCE to replace the Ler coding sequence with the kanamycin resistance gene cassette. The antibiotic resistance cassette was subsequently removed via flanking Flp recombination target (FRT) sites using the temperature sensitive FLP expression plasmid pCP20 [56]. The Δler locus in the resulting unmarked mutants encoded the first and last 9 aa of Ler (first 3 aa of the shorter Ler protein as described by Mellies et al. [57]) sandwiching a central “scar region” derived from the FLP recombinase sites encoding 29 (non-Ler) amino acids. Loss of all three plasmids (pDOC-derived donor plasmids and pACBSCE, pCP20) involved in mutagenesis was confirmed by antibiotic resistance profiling. The DNA sequence surrounding the recombination site was checked by sequencing across the knockout locus from primers designed to bind flanking sites. Recombinant strains were designated LBEC1 (EDL933 Δler) and LBEC2 (E2348/69 Δler).

The absence of gross unwanted deletions in the mutant strains was confirmed by comparative genomic hybridization (CGH) of labeled genomic DNA extracted from wild-type and mutant (Δler) strains. No missing loci, other than the desired deletion of ler, were apparent. Growth curves were assessed for LBEC1 and LBEC2 strains in comparison to parental wild-types and no gross defects in growth were observed (a small growth advantage consistent with predicted increased fitness due to reduced expression of T3SS was sometimes observed for the mutant strains on growth in inducing Dulbecco's Modified Eagle Medium (DMEM) medium, but this was neither statistically significant nor reproducible).

The ler mutation in strains LBEC1 and LBEC2 was successfully complemented using the ler expression plasmid pSI04 resulting in the restoration of a functional T3SS, as confirmed by the fluorescent actin staining (FAS) test (i.e. via microscopic assessment of AE lesion formation (table 8) [58]. Subconfluent HeLa cell monolayers on glass coverslips were infected for 4 hours at 37°C with a 1∶100 dilution of an overnight LB broth culture of E. coli diluted in DMEM buffered with 25 mM HEPES. Following fixation in 4% formalin for 20 minutes and permeabilization in 0.1% Triton in PBS for 4 minutes, cells were stained with 12 µg/ml FITC conjugated phalloidin (Sigma) for 20 minutes at room temperature [54]. Bacterial cells were simultaneously stained with 10 µg/ml propidium iodide (Invitrogen).

thumbnail
Table 8. Summary of FAS1 testing of complementation of E. coli ler mutant strains.

https://doi.org/10.1371/journal.pone.0080160.t008

RNA Purification

Quadruplicate overnight cultures of WT and Δler strains, grown in LB broth (Miller formulation) were diluted 1/100 into DMEM buffered with 25 mM HEPES and incubated at 37°C, with aeration by shaking at 200 rpm (i.e. inducing conditions for expression of the LEE). Samples were harvested at mid and late log phases of growth (OD600 of 0.4 and 0.9 for EPEC; 0.5 and 1.1 for EHEC). Messenger RNA was stabilized immediately by pipetting the samples directly into RNAprotect Bacteria reagent (Qiagen) before purification of total RNA using the RNeasy Mini Kit with on-column DNase digestion (Qiagen).

Microarray labelling and hybridization

The concentration of RNA was determined using a spectrophotometer (ND-1000; NanoDrop). Five hundred nanograms of total RNA was used for labelling, and aRNA was synthesized with the Ambion MessageAmp™ II-Bacteria RNA Amplification Kit according to the recommendations of the manufacturer and labeled with the Cy3 or Cy5 monoreactive dye pack (GE Healthcare). Labeled aRNA was purified with Qiagen RNeasy MinElute clean up kit according to the manufacturer's instructions and quantified using a spectrophotometer (ND-1000; NanoDrop). The 8×15,000 (15K) DNA high-density microarrays of E2348/69 and EDL933 were designed by Oxford Gene Technology (Oxford OX5 1PF, United Kingdom) and validated by the University of Birmingham E. coli Centre (UBEC) (United Kingdom). During validation, three 60-mer probes per predicted gene were designed for all the open reading frames (ORFs) in the chromosome and plasmids of each one of the two E. coli strains used in this study. For each of the designed probes, a mismatch probe (containing 3 mismatches per 60-mer probe at positions 10, 25, and 40) was also generated. These mismatch probes and the perfect-match probes designed against each strain were placed on an array (4×44k) in triplicate. This array was hybridized with genomic DNA and a pool of mRNA representing conditions in which as many genes as practicable would be induced (derived from an equimolar pool of total RNA from E. coli grown in morpholinepropanesulfonic acid (MOPS) minimal medium at 30°C mid-log phase, 37°C for mid-log phase, and 37°C for stationary phase). The results were processed to select the best-performing probe for each gene. This derived and optimized probe set was printed in a random pattern in triplicate by Agilent Technologies on an 8×15K array for each strain and used in this study. For each of the four biological replicates equal quantities (300 ng) of Cy5- and Cy3-labeled aRNA were added to hybridization solution, and hybridization was performed using the Gene Expression hybridization kit (Agilent Technologies).

Analysis of Microarray Data

The microarray images were analyzed using GenePix software v6 (Axon Instruments). The data were imported into GeneSpring, version 7 (Agilent). A Lowess curve (locally weighted linear regression curve) was fitted to the plot of log intensity versus log ratio, and 40% of the data were used to calculate the Lowess fit at each point. The curve was used to adjust the control value for each measurement. If the control channel signal was below a threshold value of 10, then 10 was used instead.

For each strain data set a list of genes was prepared showing at least 2-fold differential expression levels between the ler and wild type samples for each one of the two growth conditions by using Student's t-test and applying the Benjamini and Hochberg false discovery rate (multiple testing correction, MTC) test with a p value cut off of 0.05.

Acknowledgments

The authors would like to thank Dr Y. Sevastsyanovich for commenting on a draft of this manuscript.

Author Contributions

Conceived and designed the experiments: LEHB CC LASS CWP SJWB MJP. Performed the experiments: LEHB CC MP RKS MSI DJL. Analyzed the data: LEHB CC LASS. Contributed reagents/materials/analysis tools: MSI DJL. Wrote the paper: LEHB CC MSI RKS MJP.

References

  1. 1. Kaper JB, Nataro JP, Mobley HLT (2004) Pathogenic Escherichia coli. Nat Rev Microbiol 2: 123–140.
  2. 2. Elliott SJ, Wainwright LA, McDaniel TK, Jarvis KG, Deng YK, et al. (1998) The complete sequence of the locus of enterocyte effacement (LEE) from enteropathogenic Escherichia coli E2348/69. Mol Microbiol 28: 1–4.
  3. 3. McDaniel TK, Jarvis KG, Donnenberg MS, Kaper JB (1995) A genetic locus of enterocyte effacement conserved among diverse enterobacterial pathogens. Proc Natl Acad Sci USA 92: 1664–1668.
  4. 4. Frankel G, Phillips AD, Rosenshine I, Dougan G, Kaper JB, et al. (1998) Enteropathogenic and enterohaemorrhagic Escherichia coli: more subversive elements. Mol Microbiol 30: 911–921.
  5. 5. Ogura Y, Ooka T, Iguchi A, Toh H, Asadulghani M, et al. (2009) Comparative genomics reveal the mechanism of the parallel evolution of O157 and non-O157 enterohemorrhagic Escherichia coli. Proc Natl Acad Sci USA 106: 17939–17944.
  6. 6. Tobe T, Beatson SA, Taniguchi H, Abe H, Bailey CM, et al. (2006) An extensive repertoire of type III secretion effectors in Escherichia coli O157 and the role of lambdoid phages in their dissemination. Proc Natl Acad Sci USA 103: 14941–14946.
  7. 7. Wong ARC, Pearson JS, Bright MD, Munera D, Robinson KS, et al. (2011) Enteropathogenic and enterohaemorrhagic Escherichia coli: even more subversive elements. Mol Microbiol 80: 1420–1438.
  8. 8. Spears KJ, Roe AJ, Gally DL (2006) A comparison of enteropathogenic and enterohaemorrhagic Escherichia coli pathogenesis. FEMS Microbiol Lett 255: 187–202.
  9. 9. Vanaja SK, Springman AC, Besser TE, Whittam TS, Manning SD (2010) Differential expression of virulence and stress fitness genes between Escherichia coli O157:H7 strains with clinical or bovine-biased genotypes. Appl Environ Microbiol 76: 60–68.
  10. 10. Abu-Ali GS, Ouellette LM, Henderson ST, Whittam TS, Manning SD (2010) Differences in adherence and virulence gene expression between two outbreak strains of enterohaemorrhagic Escherichia coli O157 : H7. Microbiology 156: 408–419.
  11. 11. Hueck CJ (1998) Type III protein secretion systems in bacterial pathogens of animals and plants. Microbiol Mol Biol Rev 62: 379–433.
  12. 12. Elliott SJ, Sperandio V, Giron JA, Shin S, Mellies JL, et al. (2000) The locus of enterocyte effacement (LEE)-encoded regulator controls expression of both LEE- and non-LEE-encoded virulence factors in enteropathogenic and enterohemorrhagic Escherichia coli. Infect Immun 68: 6115–6126.
  13. 13. Bustamante VH, Santana FJ, Calva E, Puente JL (2001) Transcriptional regulation of type III secretion genes in enteropathogenic Escherichia coli: Ler antagonizes H-NS-dependent repression. Mol Microbiol 39: 664–678.
  14. 14. Mellies JL, Benison G, McNitt W, Mavor D, Boniface C, et al. (2011) Ler of pathogenic Escherichia coli forms toroidal protein-DNA complexes. Microbiology 157: 1123–1133.
  15. 15. Berdichevsky T, Friedberg D, Nadler C, Rokney A, Oppenheim A, et al. (2005) Ler is a negative autoregulator of the LEE1 operon in enteropathogenic Escherichia coli. J Bacteriol 187: 349–357.
  16. 16. Barba J, Bustamante VH, Flores-Valdez MA, Deng WY, Finlay BB, et al. (2005) A positive regulatory loop controls expression of the locus of enterocyte effacement-encoded regulators Ler and GrlA. J Bacteriol 187: 7918–7930.
  17. 17. Mellies JL, Elliott SJ, Sperandio V, Donnenberg MS, Kaper JB (1999) The Per regulon of enteropathogenic Escherichia coli: identification of a regulatory cascade and a novel transcriptional activator, the locus of enterocyte effacement (LEE)-encoded regulator (Ler). Mol Microbiol 33: 296–306.
  18. 18. Porter ME, Mitchell P, Free A, Smith DGE, Gally DL (2005) The LEE1 promoters from both enteropathogenic and enterohemorrhagic Escherichia coli can be activated by PerC-like proteins from either organism. J Bacteriol 187: 458–472.
  19. 19. Islam MS, Bingle LEH, Pallen MJ, Busby SJW (2011) Organization of the LEE1 operon regulatory region of enterohaemorrhagic Escherichia coli O157:H7 and activation by GrlA. Mol Microbiol 79: 468–483.
  20. 20. Sperandio V, Mellies JL, Nguyen W, Shin S, Kaper JB (1999) Quorum sensing controls expression of the type III secretion gene transcription and protein secretion in enterohemorrhagic and enteropathogenic Escherichia coli. Proc Natl Acad Sci USA 96: 15196–15201.
  21. 21. Sperandio V, Li CC, Kaper JB (2002) Quorum-sensing Escherichia coli regulator A: a regulator of the LysR family involved in the regulation of the locus of enterocyte effacement pathogenicity island in enterohemorrhagic E. coli. Infect Immun 70: 3085–3093.
  22. 22. Russell RM, Sharp FC, Rasko DA, Sperandio V (2007) QseA and GrlR/GrlA regulation of the locus of enterocyte effacement genes in enterohemorrhagic Escherichia coli. J Bacteriol 189: 5387–5392.
  23. 23. Tree JJ, Wolfson EB, Wang D, Roe AJ, Gally DL (2009) Controlling injection: regulation of type III secretion in enterohaemorrhagic Escherichia coli. Trends in Microbiology 17: 361–370.
  24. 24. Mellies JL, Barron AMS, Carmona AM (2007) Enteropathogenic and enterohemorrhagic Escherichia coli virulence gene regulation. Infect Immun 75: 4199–4210.
  25. 25. Umanski T, Rosenshine I, Friedberg D (2002) Thermoregulated expression of virulence genes in enteropathogenic Escherichia coli. Microbiology 148: 2735–2744.
  26. 26. Bergholz TM, Wick LM, Qi W, Riordan JT, Ouellette LM, et al. (2007) Global transcriptional response of Escherichia coli O157:H7 to growth transitions in glucose minimal medium. BMC Microbiol 7: 97.
  27. 27. Sharma VK, Zuerner RL (2004) Role of hha and ler in transcriptional regulation of the esp operon of enterohemorrhagic Escherichia coli O157 : H7. J Bacteriol 186: 7290–7301.
  28. 28. Xicohtencatl-Cortes J, Saldaña Z, Deng W, Castañeda E, Freer E, et al. (2010) Bacterial macroscopic rope-like fibers with cytopathic and adhesive properties. J Biol Chem 285: 32336–32342.
  29. 29. Vidal JE, Navarro-Garcia F (2006) Efficient translocation of EspC into epithelial cells depends on enteropathogenic Escherichia coli and host cell contact. Infect Immun 74: 2293–2303.
  30. 30. Salinger N, Kokona B, Fairman R, Okeke IN (2009) The plasmid-encoded regulator activates factors conferring lysozyme resistance on enteropathogenic Escherichia coli strains. Appl Environ Microbiol 75: 275–280.
  31. 31. Mellies JL, Navarro-Garcia F, Okeke I, Frederickson J, Nataro JP, et al. (2001) espC pathogenicity island of enteropathogenic Escherichia coli encodes an enterotoxin. Infect Immun 69: 315–324.
  32. 32. Grys TE, Siegel MB, Lathem WW, Welch RA (2005) The StcE protease contributes to intimate adherence of enterohemorrhagic Escherichia coli O157:H7 to host cells. Infect Immun 73: 1295–1303.
  33. 33. Roe AJ, Tysall L, Dransfield T, Wang D, Fraser-Pitt D, et al. (2007) Analysis of the expression, regulation and export of NleA-E in Escherichia coli O157:H7. Microbiology 153: 1350–1360.
  34. 34. Garmendia J, Frankel G (2005) Operon structure and gene expression of the espJ-tccP locus of enterohaemorrhagic Escherichia coli 0157 : H7. FEMS Microbiol Lett 247: 137–145.
  35. 35. Torres AG, Lopez-Sanchez GN, Milflores-Flores L, Patel SD, Rojas-Lopez M, et al. (2007) Ler and H-NS, regulators controlling expression of the long polar fimbriae of Escherichia coli O157 : H7. J Bacteriol 189: 5916–5928.
  36. 36. Abe H, Miyahara A, Oshima T, Tashiro K, Ogura Y, et al. (2008) Global regulation by horizontally transferred regulators establishes the pathogenicity of Escherichia coli. DNA Res 15: 25–38.
  37. 37. Wieler LH, McDaniel TK, Whittam TS, Kaper JB (1997) Insertion site of the locus of enterocyte effacement in enteropathogenic and enterohemorrhagic Escherichia coli differs in relation to the clonal phylogeny of the strains. FEMS Microbiol Lett 156: 49–53.
  38. 38. Hommais F, Krin E, Laurent-Winter C, Soutourina O, Malpertuy A, et al. (2001) Large-scale monitoring of pleiotropic regulation of gene expression by the prokaryotic nucleoid-associated protein, H-NS. Mol Microbiol 40: 20–36.
  39. 39. Cordeiro TN, Schmidt H, Madrid C, Juárez A, Bernadó P, et al. (2011) Indirect DNA readout by an H-NS related protein: structure of the DNA complex of the C-terminal domain of Ler. PLoS Pathog 7: e1002380.
  40. 40. Navarre WW, Porwollik S, Wang Y, McClelland M, Rosen H, et al. (2006) Selective silencing of foreign DNA with low GC content by the H-NS protein in Salmonella. Science 313: 236–238.
  41. 41. Bishop RE, Kim S-H, El Zoeiby A (2005) Role of lipid A palmitoylation in bacterial pathogenesis. J Endotox Res 11: 174–180.
  42. 42. Mildiner-Earley S, Miller VL (2006) Characterization of a novel porin involved in systemic Yersinia enterocolitica infection. Infect Immun 74: 4361–4365.
  43. 43. Keenleyside WJ, Jayaratne P, MacLachlan PR, Whitfield C (1992) The rcsA gene of Escherichia coli O9:K30:H12 is involved in the expression of the serotype-specific group I K (capsular) antigen. J Bacteriol 174: 8–16.
  44. 44. Mellies JL, Barron AMS, Haack KR, Korson AS, Oldridge DA (2006) The global regulator Ler is necessary for enteropathogenic Escherichia coli colonization of Caenorhabditis elegans. Infect Immun 74: 64–72.
  45. 45. Garcia-Angulo VA, Deng W, Thomas NA, Finlay BB, Puente JL (2008) Regulation of expression and secretion of NleH, a new non-locus of enterocyte effacement-encoded effector in Citrobacter rodentium. J Bacteriol 190: 2388–2399.
  46. 46. Lathem WW, Bergsbaken T, Welch RA (2004) Potentiation of C1 esterase inhibitor by StcE, a metalloprotease secreted by Escherichia coli O157:H7. J Exp Med 199: 1077–1087.
  47. 47. Ho TD, Davis BM, Ritchie JM, Waldor MK (2008) Type 2 secretion promotes enterohemorrhagic Escherichia coli adherence and intestinal colonization. Infect Immun 76: 1858–1865.
  48. 48. Gomez-Duarte OG, Kaper JB (1995) A plasmid-encoded regulatory region activates chromosomal eaeA expression in enteropathogenic Escherichia coli. Infect Immun 63: 1767–1776.
  49. 49. Tobe T, Schoolnik GK, Sohel I, Bustamante VH, Puente JL (1996) Cloning and characterization of bfpTVW, genes required for the transcriptional activation of bfpA in enteropathogenic Escherichia coli. Mol Microbiol 21: 963–975.
  50. 50. Cleary J, Lai L-C, Shaw RK, Straatman-Iwanowska A, Donnenberg MS, et al. (2004) Enteropathogenic Escherichia coli (EPEC) adhesion to intestinal epithelial cells: role of bundle-forming pili (BFP), EspA filaments and intimin. Microbiology 150: 527–538.
  51. 51. Bustamante VH, Villalba MI, Garcia-Angulo VA, Vazquez A, Martinez LC, et al. (2011) PerC and GrlA independently regulate Ler expression in enteropathogenic Escherichia coli. Mol Microbiol 82: 398–415.
  52. 52. Leverton LQ, Kaper JB (2005) Temporal expression of enteropathogenic Escherichia coli virulence genes in an in vitro model of infection. Infect Immun 73: 1034–1043.
  53. 53. Wade JT, Belyaeva TA, Hyde EI, Busby SJW (2000) Repression of the Escherichia coli melR promoter by MelR: evidence that efficient repression requires the formation of a repression loop. Mol Microbiol 36: 223–229.
  54. 54. Islam MS (2011) Studies on gene expression in a pathogenic bacterium. PhD Thesis. University of Birmingham, UK.
  55. 55. Lee DJ, Bingle LEH, Heurlier K, Pallen MJ, Penn CW, et al. (2009) Gene doctoring: a method for recombineering in laboratory and pathogenic Escherichia coli strains. BMC Microbiol 9: 252.
  56. 56. Cherepanov PP, Wackernagel W (1995) Gene disruption in Escherichia coli: TcR and KmR cassettes with the option of Flp-catalyzed excision of the antibiotic-resistance determinant. Gene 158: 9–14.
  57. 57. Mellies JL, Larabee FJ, Zarr MA, Horback KL, Lorenzen E, et al. (2008) Ler interdomain linker is essential for anti-silencing activity in enteropathogenic Escherichia coli. Microbiology 154: 3624–3638.
  58. 58. Knutton S, Baldwin T, Williams PH, McNeish AS (1989) Actin accumulation at sites of bacterial adhesion to tissue culture cells: basis of a new diagnostic test for enteropathogenic and enterohemorrhagic Escherichia coli. Infect Immun 57: 1290–1298.
  59. 59. Perna NT, Plunkett G, Burland V, Mau B, Glasner JD, et al. (2001) Genome sequence of enterohaemorrhagic Escherichia coli O157:H7. Nature 409: 529–533.
  60. 60. Iguchi A, Thomson NR, Ogura Y, Saunders D, Ooka T, et al. (2009) Complete genome sequence and comparative genome analysis of enteropathogenic Escherichia coli O127:H6 strain E2348/69. J Bacteriol 191: 347–354.