mBioMBiombiombiomBiomBio2150-7511American Society of Microbiology1752 N St., N.W., Washington, DC260602754471563mBio00595-1510.1128/mBio.00595-15Research ArticlecsrR, a Paralog and Direct Target of CsrA, Promotes Legionella pneumophila Resilience in WaterCsrR Is Paralog and Target of CsrAAbbott et al.AbbottZachary D.aYakhninHelenbBabitzkePaulbSwansonMichele S.aDepartment of Microbiology and Immunology, University of Michigan Medical School, Ann Arbor, Michigan, USADepartment of Biochemistry and Molecular Biology, Center for RNA Molecular Biology, Pennsylvania State University, State College, Pennsylvania, USATaylorRonald K.EditorDartmouth Medical SchoolAddress correspondence to Michele S. Swanson, mswanson@umich.edu.

This article is a direct contribution from a Fellow of the American Academy of Microbiology.

962015May-Jun201563e00595-151042015852015Copyright © 2015 Abbott et al.2015Abbott et al.This is an open-access article distributed under the terms of the Creative Commons Attribution-Noncommercial-ShareAlike 3.0 Unported license, which permits unrestricted noncommercial use, distribution, and reproduction in any medium, provided the original author and source are credited.ABSTRACT

Critical to microbial versatility is the capacity to express the cohort of genes that increase fitness in different environments. Legionella pneumophila occupies extensive ecological space that includes diverse protists, pond water, engineered water systems, and mammalian lung macrophages. One mechanism that equips this opportunistic pathogen to adapt to fluctuating conditions is a switch between replicative and transmissive cell types that is controlled by the broadly conserved regulatory protein CsrA. A striking feature of the legionellae surveyed is that each of 14 strains encodes 4 to 7 csrA-like genes, candidate regulators of distinct fitness traits. Here we focus on the one csrA paralog (lpg1593) that, like the canonical csrA, is conserved in all 14 strains surveyed. Phenotypic analysis revealed that long-term survival in tap water is promoted by the lpg1593 locus, which we name csrR (for “CsrA-similar protein for resilience”). As predicted by its GGA motif, csrR mRNA was bound directly by the canonical CsrA protein, as judged by electromobility shift and RNA-footprinting assays. Furthermore, CsrA repressed translation of csrR mRNA in vivo, as determined by analysis of csrR-gfp reporters, csrR mRNA stability in the presence and absence of csrA expression, and mutation of the CsrA binding site identified on the csrR mRNA. Thus, CsrA not only governs the transition from replication to transmission but also represses translation of its paralog csrR when nutrients are available. We propose that, during prolonged starvation, relief of CsrA repression permits CsrR protein to coordinate L. pneumophila’s switch to a cell type that is resilient in water supplies.

IMPORTANCE

Persistence of L. pneumophila in water systems is a public health risk, and yet there is little understanding of the genetic determinants that equip this opportunistic pathogen to adapt to and survive in natural or engineered water systems. A potent regulator of this pathogen’s intracellular life cycle is CsrA, a protein widely distributed among bacterial species that is understood quite well. Our finding that every sequenced L. pneumophila strain carries several csrA paralogs—including two common to all isolates—indicates that the legionellae exploit CsrA regulatory switches for multiple purposes. Our discovery that one paralog, CsrR, is a target of CsrA that enhances survival in water is an important step toward understanding colonization of the engineered environment by pathogenic L. pneumophila.

cover-dateMay/June 2015
INTRODUCTION

Many Gram-negative and Gram-positive bacteria rely on a CsrA protein to regulate gene expression (13). Some of these highly conserved and well-studied RNA binding proteins repress translation by binding to the 5′ untranslated region (5′-UTR) of target mRNAs and preventing ribosome access to the Shine-Dalgarno (SD) sequence. Several species of bacteria, including Escherichia coli (47), Pseudomonas aeruginosa (8), Bacillus subtilis (9), Salmonella enterica serovar Typhimurium (10), and Legionella pneumophila (11), practice this mode of posttranscriptional regulation. CsrA binding can also regulate mRNA translation or stability by other mechanisms (3). For example, the E. coli CsrA protein binds in the 5′ coding region of the sdiA transcript to repress its translation (12).

In the Gram-negative, aquatic bacterium L. pneumophila, CsrA regulates a key intracellular life cycle switch within host amoeba or mammalian macrophages (1315). L. pneumophila differentiates from a replicative form—capable of multiplying to high numbers within a host cell when nutrients are abundant—to a transmissive form—equipped to lyse out of and infect new host cells (16). CsrA is essential for transmissive-phase cells to convert to the replicative phase, and it actively represses transmissive phenotypes when nutrients are available to support growth (11, 13, 15, 17). When nutrients become limiting, the LetA/LetS two-component system activates transcription of two noncoding RNAs, rsmY and rsmZ, that directly bind and sequester CsrA, thus relieving its repression of transmission traits, including expression of dot and icm substrates (11, 1720).

L. pneumophila can also persist in diverse freshwater reservoirs, ranging from ponds to engineered water systems and water-cooling towers (2125). When machinery vaporizes water contaminated with L. pneumophila, the bacteria can gain access to the human lung. Remarkably, strategies that evolved to promote persistence within phagocytic amoebae also equip the pathogen to survive and replicate in alveolar macrophages (22, 26). Therefore, understanding environmental survival and persistence is critical to designing strategies to reduce the risks L. pneumophila poses to human health.

Although distinct extracellular forms of L. pneumophila, including biofilm communities, mature intracellular forms (MIFs), and viable but nonculturable cells (VBNCs), have been described previously (2731), the mechanisms that dictate the switch into these environmental forms are not yet known. By applying phenotypic, biochemical, and molecular assays, we identified in the L. pneumophila core genome a dual-CsrA system whose integrated design is predicted to promote reciprocal expression of distinct cell types.

RESULTSIdentification of 71 <italic>csrA</italic>-like genes in <italic>L. pneumophila</italic> species.

Initial genome sequencing projects revealed three csrA-like genes in the carbon storage regulator family (32). By performing a bioinformatic search of the L. pneumophila Philadelphia-1 genome, we identified five csrA-like genes, including the well-characterized canonical csrA (Fig. 1A) (13). The amino acid sequences of the four paralogs exhibit 28% to 58% identity and 59% to 80% similarity.

CsrA paralogs in 12 strains of legionellae identified in silico. (A) Alignment of the five CsrA-like proteins in L. pneumophila Philadelphia-1, with the essential arginine residue highlighted and the two RNA binding pockets outlined (1). (B) Sequence identity of the 63 CsrA-like proteins found in 12 strains of legionellae. Darker shading indicates greater percent identity. Each of the five CsrA proteins encoded by L. pneumophila Philadelphia-1 is marked, and clusters are named for two core and the remaining ICE-associated genes.

The amino acid residues determined to be critical for CsrA function in E. coli are retained in CsrA proteins across myriad genera surveyed, including the canonical L. pneumophila CsrA protein (1). By aligning the amino acid sequences of the csrA paralogs of L. pneumophila Philadelphia-1, we confirmed that several key residues, including the arginine residue that is essential for the RNA binding activity, are retained in all five (1) (Fig. 1A). Based on the conserved amino acid sequence, the five L. pneumophila Philadelphia-1 loci are likely valid csrA paralogs.

To determine if the Philadelphia-1 strain is unique in encoding multiple csrA paralogs, the in silico search was expanded to include 13 more fully sequenced strains of legionellae: 8 additional independent isolates of L. pneumophila, 2 laboratory strains derived from Philadelphia-1, 2 strains of L. longbeachiae, and 1 strain of L. drancourtii (see Table S1 in the supplemental material). Altogether, 71 csrA-like genes were identified, and each retained several residues predicted to be key for RNA binding (1). In particular, the arginine residue critical for RNA binding was conserved in 70 of the 71 proteins; in the one exception, another cationic residue, lysine, was in place of the arginine (see Table S2). Each of the 14 surveyed strains contained between 4 and 7 csrA-like genes. Without exception, all 14 encoded not only the canonical csrA gene but also a copy of a previously uncharacterized paralog, lpg1593, which we name here csrR. Accordingly, both the csrA and csrR genes appear to belong to the core genome of Legionella species. In stark contrast, the remainder of the csrA-like genes are located within previously characterized or putative integrative conjugative elements (ICEs), large genomic islands known or predicted to transfer horizontally between strains of legionellae (see Table S3) (3335).

To compare the sequences of the 63 Legionella csrA genes found in the 12 strains, excluding the two laboratory-derived strains Lp02 and JR32, we generated a heat map that displays the amino acid similarity of each protein to each of the other paralogs. By this approach, we recognized three clusters: two that correspond to the presumptive core CsrA and CsrR proteins and a third that encompasses a more diverse group of ICE-related CsrA-like proteins (Fig. 1B). Because csrR is encoded by all surveyed strains, is highly conserved, and retains key RNA binding residues, we hypothesized that this gene would regulate traits fundamental to L. pneumophila’s life style, as CsrA does.

<italic>L. pneumophila</italic> requires the conserved <italic>csrR</italic> locus to persist under nutrient-poor conditions.

Since csrR is a paralog of the essential regulator csrA, we first tested whether the two genes regulate the same developmental stage. To do so, we either deleted or induced expression of the csrR gene and then assayed phenotypes known to be regulated by the canonical csrA gene (13). Analyzed under standard laboratory culture conditions (37°C in rich ACES-buffered yeast extract with thymidine [AYET] medium), neither loss nor gain of csrR function altered expression of any L. pneumophila replication-phase or transmission-phase traits tested. These included, for broth cultures, growth, pigmentation, motility, and sensitivity to sodium, heat, and osmotic stress, as well as infection and intracellular growth in primary mouse macrophages and amoebae (data not shown) (13). Since csrR appeared unlikely to regulate traits that are hallmarks of the commonly studied replicative and transmissive cell types, we next considered conditions that L. pneumophila likely encounters in the environment. Indeed, a dual-CsrA system (RsmA/RsmF) equips P. aeruginosa to modulate biofilm formation (36).

The legionellae naturally thrive in freshwater and soil habitats, and L. pneumophila can survive for long periods in water in a VBNC state (29, 30). Therefore, we tested whether csrR coordinates differentiation to a resilient environmental cell type. Employing the method of Garduño and colleagues (30), we inoculated autoclaved tap water at 45°C with wild-type (WT) L. pneumophila, a csrR mutant, or a csrR mutant strain in which a wild-type csrR allele had been integrated into the chromosome. After 24 days, we quantified bacterial viability by fluorescence microscopy using a LIVE/DEAD BacLight stain and culturability on rich nutrient agar (37). The wild-type culture retained 79% viability, which was significantly higher than the mutant culture’s 50% viability (P value = 0.049; Fig. 2A). Likewise, the level of colony forming units (CFU) recovered at 24 days relative to day 0 was significantly higher for the wild-type culture than for the mutant culture (5.6% of WT, 0.3% of mutant; P value = 0.0027) (Fig. 2B). Although the merodiploid mutant strain exhibited an intermediate level of viability as measured by microscopy assay (Fig. 2A) (P > 0.05 compared either to the wild type or to the mutant), genetic complementation was incomplete. Two control experiments indicated that the lack of full complementation was likely due to poor expression of csrR. First, quantitative reverse transcriptase PCR (qRT-PCR) revealed that the merodiploid strain contained approximately 5% of the wild-type levels of csrR RNA (data not shown). Second, we quantified reduced viability relative to wild-type L. pneumophila by both microscopy and CFU assays after isolating and analyzing nine independent csrR mutant strains constructed by three different strategies.

csrR enhances survival of L. pneumophila in water. After incubation for 24 days in 45°C tap water, survival of the wild-type strain (MB1368; black), ΔcsrR deletion mutants (MB1370 and MB1371; light gray), and merodiploid strains (MB1373 and MB1374; dark gray) was assessed microscopically by LIVE/DEAD fluorescence staining (A) and by enumerating CFU (B). Error bars represent standard deviations calculated from the results determined for duplicate samples from two independent isolates of each genotype analyzed in parallel. Significance was calculated using a Student’s t test. Differences reported for the wild-type and ΔcsrR strains are representative of the results of multiple independent experiments.

<italic>csrR</italic> mRNA contains multiple CsrA binding sites.

Because knowledge of a protein’s regulation can provide insight into its function, we next analyzed the upstream sequences of csrR. Bioinformatic analysis identified motifs typical of CsrA binding sites in the 5′ noncoding region of csrR (11). As a first step to analyze csrR regulation, we mapped by 5′ rapid amplification of cDNA ends (5′-RACE) its transcriptional start site to −52 relative to the translational start (data not shown). Three independent sequencing events revealed that the transcript starting with the −52 start site was predominant, although rare transcripts started at −53 or −54. On the basis of the data indicating the predominant species, we made several predictions about the csrR mRNA. Putative −10 and −35 sites as well as the likely SD sequence approximately 10 bases 5′ of the translational start codon were identified (Fig. 3A). RNA-fold software predicted a secondary structure for the csrR transcript in which the SD sequence is positioned in a loop extending from a long stem (Fig. 3B), a common binding motif for CsrA protein (38). Accordingly, we postulated that the csrR transcript is bound and its stability regulated by CsrA.

A putative CsrA binding site on the csrR RNA overlaps with a ribosome binding site motif. (A) The csrR transcriptional start site (TSS, +1; black arrow) was mapped by 5′ RACE. Likely −10 and −35 promoter elements are underlined and labeled; a putative Shine-Dalgarno (SD) sequence and the associated flanking bases predicted to form a stem-loop structure recognized by CsrA and located 10 bases 5′ of the translational start are underlined; and the csrR open reading frame is marked with a gray arrow and capitalized letters. (B) The csrR RNA encodes a putative SD (+37) sequence located on a stem-loop and an ATG start codon (+51) positioned at the base of two stem-loops, as predicted by the IDT UNAFold algorithm. CsrA protein is predicted to bind the csrR RNA SD sequence (boxed).

To test directly whether CsrA protein binds to csrR mRNA, we performed an in vitro assay. Purified CsrA protein was incubated with 5′-end-labeled csrR RNA, and then binding was analyzed by gel electromobility shift assay. As predicted by the in silico analysis, CsrA protein bound tightly to csrR RNA, with a calculated dissociation constant of 32 ± 9 nM (Fig. 4A). This physical interaction was specific, since incubation with an excess of unlabeled csrR RNA effectively competed with CsrA protein binding to labeled csrR RNA (Fig. 4B). In contrast, incubation with even higher concentrations of a heterologous competitor RNA, E. coli phoB, did not inhibit CsrA binding to csrR. Therefore, CsrA protein binds csrR RNA tightly and specifically in vitro.

CsrA binds tightly and specifically to csrR RNA. (A) 5′-end-labeled csrR RNA (0.1 nM) was incubated with CsrA protein at the indicated concentration (nM), and the complexes were separated by gel electrophoresis. Positions of bound (marked B) and free (marked F) RNA species are shown. The binding constant (Kd) was calculated from the binding curve shown at right. (B) Labeled csrR RNA (0.1 nM) was incubated with CsrA protein in the absence or presence of unlabeled specific (csrR) or nonspecific (E. coli phoB) competitor RNA at the concentrations indicated. (C) 5′-End-labeled csrR RNA was treated with RNase T1 in the presence of CsrA protein at the concentration indicated. Also shown are partial alkaline hydrolysis (column OH) and RNase T1 digestion (column T1) ladders, as well as a control sample that lacked RNase T1 treatment (column C). Labels indicate residues for which CsrA protein reduced RNase T1 cleavage (BS1 to BS3), the csrR Shine-Dalgarno (SD) sequence, and the translation initiation codon (Met). Numbering is with respect to the start site of csrR transcription (+1).

To verify our observation that CsrA protein binds to the csrR transcript at the predicted stem-loop structure surrounding the SD sequence, we performed RNA footprinting experiments. Interestingly, although CsrA did bind at the predicted SD site (“BS2” in Fig. 4C), the protein also protected a GGA motif within the first two codons of the csrR coding region (“BS3” in Fig. 4C). Although binding in the 5′ untranslated region of its target gene is typical, in some cases CsrA does bind in coding regions (12). The RNA footprinting assay also revealed a potential third binding site (BS1); however, because this third site was protected only weakly, we focused next on the stronger BS2 and BS3 binding sites.

Transcriptional and translational fusions reveal a contribution of BS3 to <italic>csrR</italic> regulation.

To analyze in live cells whether either or both BS2 and BS3 binding sites contribute to csrR posttranscriptional regulation, we constructed a series of green fluorescent protein (GFP) reporters (Fig. 5A). For the BS2-only transcriptional reporter, gfp expression is driven by a sequence corresponding to approximately 300 bp immediately 5′ of the csrR coding region that includes BS2. Although this reporter contains the csrR SD sequence, its gfp SD sequence likely controls translation; thus, the BS2-only construct is referred to as a transcriptional reporter. The BS2/BS3 translational reporter includes both the BS2 and BS3 directly 5′ and in frame with gfp coding sequences, eliminating the gfp SD sequence and including the first 5 codons of csrR. For the BS2 stem-loop-disrupted (BS2-sld)/BS3 reporter, the BS2 stem-loop structure was disrupted on the translational reporter by changing the six consecutive adenines immediately 5′ of the SD sequence to uracils, thereby preventing stem formation with the eight uracils 3′ of the SD sequence.

BS3 mediates repression of csrR-GFP reporters. (A) Schematics of csrR-GFP reporters to analyze csrR regulation. The BS2-only transcriptional reporter (MB1375; circles) encodes ~300 bp 5′ to the translational start sequence, including the csrR SD sequence, which overlaps the BS2 CsrA binding site, and the gfp ribosome binding site (RBS). In the BS2/BS3 translational reporter (MB1376; squares) the gfp RBS is replaced by the first five codons of csrR, which encode the BS3 CsrA binding site. The BS2-sld/BS3 translational reporter (MB1377; triangles) was constructed by eliminating the stem-loop base pairing to disrupt the BS2 CsrA binding site on the dual reporter. (B) E-phase cultures of L. pneumophila carrying the reporter indicated were incubated for 24 to 28 h, and GFP fluorescence was quantified at the times indicated and expressed in relative fluorescence units (RFU). Reference strains were L. pneumophila carrying the empty gfp vector (diamonds) or flaA-GFP reporter (gray line) as an indicator of the transition into PE phase. Shown are means ± standard deviations calculated from the results determined from triplicate samples; error bars for the stem-loop mutant are too small to be depicted. Similar patterns were observed in at least two other experiments.

L. pneumophila strains carrying each reporter were cultured in broth, and then aliquots were assessed for GFP fluorescence. As a marker for the growth phase, we also analyzed a reporter of the flaA flagellin gene, which is strongly induced in post-exponential (PE) phase (39). The BS2-only reporter generated a high level of fluorescence in exponential (E) phase (~15,000 relative fluorescence units [RFU]) that then decreased (to ~5,000 RFU) in PE phase (Fig. 5B), the period when transmission traits are induced (16). In comparison to the BS2-driven expression, in E phase, the BS2/BS3 and the BS2-sld/BS3 translational reporters demonstrated 3-fold and 6-fold less fluorescence, respectively (Fig. 5B). Therefore, the BS3 site appears to destabilize the mRNA and/or inhibit translation of the transcript. In addition, on the basis of the reduced fluorescence seen when the stem-loop was disrupted, the BS2 site may somehow stabilize the mRNA.

Mutation of the CsrA binding site relieves CsrA repression <italic>in vivo</italic>.

To assess whether CsrA protein destabilizes the csrR transcript in vivo, we utilized a previously characterized conditional csrA mutant (13). In this strain, the chromosomal csrA locus is deleted, and a plasmid carries a functional copy of the gene under IPTG (isopropyl-β-d-thiogalactopyranoside)-inducible control. Thus, we measured the impact of CsrA protein on csrR transcript levels by culturing cells in the presence and absence of IPTG.

To determine whether CsrA alters the stability of csrR transcripts, the conditional csrA mutant was cultured with or without IPTG for 9 h, which corresponded to E phase, a period when csrR promoter activity is high (Fig. 5B) (40). Next, transcription was inhibited by treatment with rifampin. RNA was isolated 0 and 15 min later, and then the relative levels of csrR mRNA were quantified by qRT-PCR. There was an inverse correlation between the presence of CsrA protein and csrR mRNA levels (Fig. 6A), indicating that the presence of CsrA decreased csrR transcript stability. A similar pattern has been reported for other CsrA-mRNA interactions (41, 42).

CsrA regulates csrR expression in vivo. (A) To analyze the impact of CsrA protein on csrR RNA stability, a conditional csrA mutant (MB464) was cultured without (black) or with (gray) IPTG to the E phase, rifampin was added to inhibit new transcription, and RNA was isolated 0 and 15 min later. Error bars indicate standard deviations of the results from duplicate samples, and significance was calculated using a Student’s t test. Results are representative of two independent experiments. (B) csrR transcript stability was assessed via qRT-PCR of E-phase L. pneumophila cultured without or with IPTG to induce expression of either the WT strain (MB1378, black) or BS3 mutant csrR (MB1379, gray). Data shown represent the fold increase in the transcript level in the induced strains relative to their uninduced controls. Error bars indicate standard deviations of the results from duplicate samples, and significance was calculated using a Student’s t test. Results are representative of three independent experiments. (C) The strains whose results are shown in panel B were cultured into E phase with or without IPTG, and then protein levels were analyzed by Western assay using anti-His antibody. L. pneumophila encoding His-tagged CsrA on a plasmid served as the positive control. Data are representative of the results of four independent experiments done with two independent isolates of each strain.

We then tested in vivo whether CsrA destabilization of csrR transcript depends on its BS3 motif (Fig. 4 and 5). To do so, we performed qRT-PCR on wild-type L. pneumophila strains that carried plasmids in which an IPTG-inducible promoter drives either a wild-type or BS3 mutant csrR allele that also encodes a 6-histidine epitope. To ensure the presence of active CsrA, RNA was isolated from strains cultured with or without IPTG for 9 h in E phase. Induction of the BS3 mutant generated >8-fold more csrR transcript than the corresponding uninduced control, whereas the wild-type allele showed a <2-fold increase (P value = 0.0327; Fig. 6B). Thus, in cultures of wild-type L. pneumophila, csrR RNA is destabilized by a mechanism that requires its BS3 binding site for the CsrA repressor protein.

Finally, we tested the prediction that, by decreasing csrR transcript stability, CsrA repressor binding reduces CsrR protein levels. For this purpose, we performed Western analysis on the strains that encode His-tagged CsrR on mRNA that either contained or lacked the BS3 RNA binding site for CsrA protein. CsrR protein was detected only in L. pneumophila that carried the csrR allele that lacked the BS3 binding site for CsrA (Fig. 6C). Therefore, CsrR protein expression is inhibited by the BS3 sequence positioned at the start of the csrR opening reading frame, most likely by direct binding of csrR mRNA by the CsrA repressor protein (Fig. 3 to 5).

DISCUSSION

In nature, legionellae must adapt to the stresses of the intracellular environment of professional phagocytes and nutrient-poor aqueous habitats (22, 2628, 30, 43). L. pneumophila relies on the CsrA RNA binding protein to regulate directly and indirectly the transition between the replicative and transmissive phases of its pathogenic life cycle (11, 13, 15, 17). Here we demonstrate that L. pneumophila survival during aquatic stress is enhanced by the locus encoding the CsrA-like protein CsrR, a broadly conserved factor whose translation is directly repressed by CsrA.

Of the numerous csrA paralogs encoded by Legionella species, only two are completely conserved in all surveyed strains: canonical csrA (13) and the gene we name csrR here (Fig. 1; see also Table S3 in the supplemental material). Four genetic observations establish that CsrR is not redundant with CsrA. First, a csrA deletion mutant cannot replicate despite the presence of chromosomal csrR (13). Second, inducing expression of csrR fails to complement the csrA growth defect (data not shown). Third, neither loss nor gain of csrR function alters expression of the panel of traits regulated by CsrA (data not shown). Instead, the csrR locus enhances L. pneumophila persistence in 45°C tap water (Fig. 2), a condition that induces resilient VBNC cells (30). Together, our initial phenotypic analyses predict that CsrR regulates environmental resilience of L. pneumophila whereas CsrA controls the intracellular replication-transmission cycle (13).

In addition, CsrA represses CsrR protein expression by a posttranscriptional mechanism. CsrA directly binds csrR mRNA in vitro (Fig. 4); in vivo, CsrA destabilizes csrR mRNA (Fig. 5 and 6A) and prevents CsrR protein expression (Fig. 6B). CsrA binding is mediated by a GGA motif in the second codon of csrR (Fig. 4C), and mutation of this motif relieves posttranscriptional repression (Fig. 6B). CsrA’s direct repression of csrR ensures reciprocal expression of this dual-CsrA system, a design that creates two mutually exclusive regulons for L. pneumophila.

CsrA represses CsrR protein expression by binding directly to its mRNA at BS3, a site comprised of a GGA motif within the first two codons of the csrR-coding region (Fig. 3). It is noteworthy that BS3 overlaps the codon for an aspartic acid residue in the second amino acid position, a striking deviation from the ubiquitous leucine residue found not only in 55 of the 57 non-CsrR CsrA paralogs encoded in the Legionella pan-genome but also in every CsrA homologue of over 30 surveyed species of Gram-negative and -positive bacteria (see Table S2 in the supplemental material) (1). This leucine-to-aspartic acid replacement is perfectly conserved in all 14 surveyed csrR genes, indicating that the BS3 motif is selected for and retained specifically, likely to conserve the CsrA-mediated repression identified here.

CsrA binds at BS3 in the coding region of the csrR transcript (Fig. 4), an uncommon, though not unprecedented, location (11, 12). A second binding site, BS2, was predicted bioinformatically and verified by RNA footprinting (Fig. 4). Nevertheless, in our reporter assays, BS2 was not necessary for inhibition. Instead, we observed that mutation of the bases surrounding BS2, which likely disrupt the stem-loop formation that is favorable for CsrA binding, actually decreased expression of the reporter (Fig. 5), indicating that BS2 (or at least the stem-loop) may have a stabilizing effect. Alternatively, disruption of the stem-loop may inhibit GFP expression by a CsrA-independent mechanism or it may enhance rather than disrupt binding by CsrA. More-detailed biochemical studies can investigate whether CsrA alternately stabilizes or destabilizes the csrR transcript depending on whether it binds at BS2 or BS3, respectively. By analogy to the E. coli molybdenum cofactor system (44), perhaps a conditional riboswitch governs CsrA binding site selection to ensure csrR expression during prolonged starvation.

At the amino acid level, CsrR is only 28% identical to canonical CsrA. Despite significant genetic drift, CsrR retains many of the residues necessary for RNA binding by E. coli CsrA (1) (Fig. 1; see also Table S2 in the supplemental material). Likewise, among the homologues of CsrA encoded by many different species of bacteria, the RNA binding pockets are largely conserved, whereas the rest of the coding region varies considerably (1), as do the targets of regulation. For example, CsrA proteins repress motility of B. subtilis by binding directly to the flagellar subunit mRNA (hag) but induce motility of E. coli by binding to and stabilizing mRNA of the flagellar regulator flhDC (4547). Evidently, L. pneumophila CsrR is under selective pressure to retain RNA binding function. The significant drift in the residues outside the two binding pockets defined by Mercante et al. (1) likely confers CsrR specificity for a distinct regulon.

Our discovery that csrR is posttranscriptionally repressed by CsrA (Fig. 3 to 5) may explain why its deletion conferred no discernible phenotype under standard laboratory conditions, which are repressive. Likewise, inducing transcription of a plasmid-borne csrR allele in cells cultured in rich media may be inconsequential because csrR mRNAs are bound and degraded by CsrA before they can be translated (Fig. 3 to 5). However, inducing CsrR protein expression from a csrR allele that lacks the CsrA BS3 binding site (Fig. 5) was also not sufficient to trigger L. pneumophila differentiation into a resilient, nonreplicative cell type. In rich AYET broth, bacterial growth, pigmentation, and motility were not affected by ectopic CsrR protein expression (data not shown). It is possible that the mRNA targets of CsrR are not expressed under the luxuriant culture conditions used, negating the effect of accumulated CsrR protein. We speculate that L. pneumophila differentiation into an environmental, resilient cell type requires additional factors to be expressed and/or CsrA repression of other determinants to be relieved.

Although the pathway that triggers CsrA derepression of CsrR expression remains to be discovered, the results reported here suggest an expanded model of the Legionella life cycle (Fig. 7). When a phagocytic cell is encountered, the transmissive form of the bacterium infects it and establishes a protective vacuole (16, 48). The CsrA repressor then equips the intracellular bacterium to differentiate into a replicative form and multiply to high numbers (13). In replication-phase cells, CsrA actively represses not only transmissive traits but also accumulation of CsrR, a regulator of environmental persistence (11, 13, 15, 17) (Fig. 2 to 5). When nutrients are exhausted within the host cell, LetA and LetS induce expression of the noncoding RNAs RsmY and RsmZ to relieve CsrA repression of transmissive traits (11, 19, 40, 49, 50). At the same time, csrR transcription declines (Fig. 5). As a consequence, the cells transition back to the motile, infectious transmissive form that can spread efficiently from one phagocytic cell to another. However, if L. pneumophila does not readily encounter another host cell and remains in a nutrient-poor extracellular environment for a prolonged period, we propose that, by mechanisms that remain to be discovered, CsrR is activated and promotes the pathogen’s long-term survival (Fig. 2).

Model of CsrA and CsrR regulation of the L. pneumophila life cycle. Within host cells, CsrA induces transition into the replicative phase (blue bacteria) and represses both transmissive traits and csrR translation. When nutrients become limiting, CsrA repression is relieved, transcription of csrR declines, and L. pneumophila differentiates to the transmissive phase (yellow bacteria). After prolonged exposure to nutrient-poor water, CsrR is activated and promotes conversion to a resilient cell type (green bacteria).

We favor a model in which the mechanism to relieve CsrA repression of transmissive traits is distinct from its derepression of CsrR protein expression. As nutrients are consumed by replicating bacteria, CsrA repression of virulence traits is relieved at least in part by two noncoding RNAs, RsmY and RsmZ (17, 19, 49). In contrast, csrR transcript—but not protein—levels are elevated in the replicative phase, as shown by our reporter assay (Fig. 5), qRT-PCR (data not shown), Western analysis (Fig. 6), and deep sequencing of RNA pools (40). The high activity of the csrR promoter in the replicative phase indicates that the L. pneumophila is spring-loaded to respond to abrupt stress, an advantage should the bacteria lack the time or resources needed to complete their natural transition into the transmissive phase. We therefore postulate that some dedicated and yet unknown mechanism relieves CsrA repression of CsrR protein expression. For example, when conditions require rapid deployment of CsrR, replicating cells could induce expression of a direct inhibitor of CsrA binding, or the conditions could facilitate a change in the secondary structure of the csrR mRNA that promotes its stabilization by binding of CsrA to BS2 over BS3.

Having two functional csrA-like genes in their core genome is an unusual feature of the legionellae. We propose that duplication and genetic drift of the gene endowed these environmental intracellular pathogens with a new regulatory protein that retained the mRNA binding function of CsrA but acquired a new cohort of targets. This evolutionary sequence may have bestowed on the legionellae the capacity to alternate between a facultative intracellular life cycle and a resilient state that can withstand a broad range of environmental conditions. Packaged within this new regulatory csrR gene is an on/off switch: a CsrA binding site in its coding region. We propose that this parsimonious design equips L. pneumophila’s two CsrA paralogs to mediate reciprocal expression of distinct regulons under appropriate conditions.

MATERIALS AND METHODSBacterial strains, culture conditions, and reagents.

Strains, plasmids, and primers used in this study are listed in Table S4 in the supplemental material. L. pneumophila Philadelphia-1 laboratory-derived strain Lp02, a thymidine auxotroph, was cultured at 37°C in AYE broth and on ACES-buffered charcoal (Fisher)-yeast extract (Becton, Dickinson) (CYE) agar supplemented with 100 µg/ml thymidine (Sigma), 400 µg/ml cysteine (Fisher), and 135 µg/ml ferric nitrate (J. T. Baker) (51). Thymidine was omitted when culturing thymidine prototroph strains. When necessary for antibiotic selection of mutants or plasmids, media were supplemented with kanamycin (Sigma) (10 µg/ml); chloramphenicol (Fisher) (5 µg/ml); or gentamicin (Gibco) (10 µg/ml). Where indicated, gene expression was induced by adding IPTG (Gold Biotechnology) to reach a final concentration of 200 µM. For all experiments, colonies were first inoculated into broth, incubated overnight, and diluted to an optical density at 600 nm (OD600) of 0.05 to 0.2 and then cultured to the E phase (OD600 of 1.0 to 2.0) or PE phase (OD600 of 3.7 to 4.0), as indicated.

<italic>csrA</italic> gene identification and heat map generation.

To identify the csrA-like genes in the Legionella pan-genome (see Table S1 in the supplemental material), the amino acid sequence of CsrA from L. pneumophila Philadelphia-1 was submitted to BLAST (http://blast.ncbi.nlm.nih.gov/Blast.cgi). The heat map was generated by first entering the percent amino acid identity of each protein pair into an Excel spreadsheet and then assigning a grayscale to the values. The proteins were then manually clustered by similarity. The order of the proteins from top to bottom (and left to right) is listed in Table S2.

<italic>csrR</italic> deletion and merodiploid.

A chloramphenicol resistance cassette was amplified from pKD3 (52) using primers ZA44 and ZA45. Flanking homology to ~500 bp 5′ and 3′ of csrR was generated from Lp02 genomic DNA using primers ZA37 and ZA43 and primers ZA46 and ZA38, respectively. The three PCR products were then joined by splicing by overhang extension (SOE) PCR using outermost primers ZA37 and ZA38 to generate the csrR::camR allele used to replace the csrR gene by natural transformation (53), producing strain MB1369.

To insert a wild-type csrR locus into a neutral site on the chromosome of the csrR mutant, the csrR gene and its native promoter were integrated into the 154-bp intergenic region between lpg2528 and lpg2529 (54). To do so, we first constructed plasmid pSU-pZLKm. The multicloning site of pSU2719 (55) was amplified using primers SC1 and SC2 and cloned into the SalI restriction site of pZL790 (54) to create vector pZL790-MCS. Next, a kanamycin resistance cassette flanked by FLP recombination target (FRT) sites was amplified from pKD4 (52) with primers Sc3 and Sc4 and cloned into the SacI site of pZL790-MCS, creating vector pZL790-MCS-Km. Then, the lpg2528-MCS-Km-lpg2529 fragment from pZL790-MCS-Km was amplified with primers SC5 and SC6 and ligated into pSU2719 at the Nco1 and HindIII sites, after the HindIII site was made blunt with Klenow fragment, producing vector pSU-pZLKm. csrR with its native promoter (~300 bp 5′ of the translational start) was amplified from Lp02 genomic DNA with primers ZA49 and ZA60 and ligated into the EcoRI site of pSU-pZLKm. This plasmid was then integrated by natural transformation into the 154-bp intergenic region between lpg2528 and lpg2529 in csrR mutant MB1369 to generate strain MB1372. Merodiploid mutants were verified by growth on CYETcam+kan and by PCR.

Restoration of thymidine prototrophy.

L. pneumophila is naturally a thymidine prototroph, but the Lp02 laboratory strain was made a thymidine auxotroph in 1993 to exploit “thymineless death” as a strategy to identify intracellular growth mutants (56). To avoid the potentially confounding phenotype of thymineless death under conditions of prolonged starvation, thymidine prototrophy was restored to strains MB110, MB1369, and MB1372 by replacing by natural transformation their mutant thyA allele with the wild-type thyA gene from L. pneumophila Philadelphia-1 encoded on pJB3395 as described previously (53), generating strains MB1368, MB1370, MB1371, MB1373, and MB1374.

Water incubation and survival.

Thymidine prototroph strains MB1368, MB1370, MB1371, MB1373, and MB1374 were analyzed to avoid thymidineless death during the prolonged incubation. The wild type and two independent isolates of each mutant and merodiploid strain were cultured in AYET medium to the E phase, diluted to an OD600 of 0.2, and then cultured to the PE phase. The cultures were then washed twice by suspension in 50 ml of autoclaved tap water and centrifugation at 4,500 × g for 15 min. Cells were then resuspended in 10 ml of autoclaved tap water in 15-ml conical tubes with the caps loosely affixed to allow air exchange and incubated statically in a 45°C incubator. The viability of duplicate samples was assessed by LIVE/DEAD BacLight (Life Technologies) staining. A 165-μl volume of culture was mixed with 165 µl of autoclaved tap water, 1 µl of each stain was added, and the samples were incubated at room temperature in the dark for 15 min. Aliquots (10 µl) were mounted on slides pretreated for 5 min with 5 µl poly-l-lysine, and 100 bacteria were scored for each replicate of each strain. Culturability of duplicate samples was quantified by plating 10-fold serial dilutions in autoclaved tap water on CYE agar.

RNA isolation.

For all experiments analyzing RNA, 2 ml of liquid culture was pelleted by centrifugation at 10,000 × g for 5 min at 4°C, the cells were resuspended in 1 ml of Trizol reagent (Life Technologies), and RNA was isolated following the manufacturer’s protocol. Residual DNA was degraded using a Turbo DNA-free kit (Life Technologies) following the manufacturer’s protocol, and RNA was stored at −20°C.

5′-RACE and RNA fold prediction.

5′-RACE was performed using a ligation-anchored PCR strategy described previously (57). Briefly, RNA was isolated as previously described, and first-strand cDNA synthesis was performed using SuperScript II reverse transcriptase (Life Technologies), the manufacturer’s protocol, and ZA54 as the csrR gene-specific primer. The ZA80 adaptor oligonucleotide, with 5′ phosphorylation for the ligation and a 3′ amino modifier to prevent its ligation, was ligated to cDNA using T4 RNA ligase (NEB) following the manufacturer’s protocol. PCR amplification was performed using ZA81 and ZA54 with platinum Taq DNA polymerase (Invitrogen), and ligation into pGEM T-easy (Promega) was performed following the manufacturer’s protocols. Sequencing from the pGEM plasmid was performed on three independent clones using primer SL3. mRNA fold prediction was done using the IDT DNA UNAFold tool, inputting the first 70 bp of the csrR mRNA sequence by using the experimentally determined transcriptional start.

CsrA protein purification.

csrA with a His-tagged epitope was amplified from Lp02 genomic DNA using primers ZA33 and ZA34, ligated into plasmid p206gent, and transformed into E. coli Rosetta(DE3) pLysS. The strain was cultured in Terrific Broth media containing 20 µg/ml chloramphenicol and 10 µg/ml gentamicin at 37°C to an OD600 of ~0.5, supplemented with 500 µM IPTG, and cultured for 16 h at 22°C. Cells were then harvested by centrifugation and flash frozen with liquid N2. Protein was purified using a 5-ml Hi-Trap metal affinity column (GE Healthcare) following the manufacturer’s instructions. Protein was loaded in a His buffer (25 mM NaH2PO4, 500 mM NaCl, 20 mM imidazole) and eluted using an imidazole gradient of 20 to 300 mM. Protein-containing fractions were then diluted 1:6 in the elution buffer (due to low solubility) and dialyzed overnight at 4°C in freezing buffer (10 mM Tris-HCl [pH 7.4], 10 mM MgCl2, 100 mM KCl, 10% glycerol).

pYH215 construction.

A csrR fragment (−168 to +103) was amplified from Lp02 genomic DNA using primers HY2 and HY4 and cloned into the EcoRI and HindIII sites of the pTZ18U polylinker (Stratagene).

Gel mobility shift assay.

RNA was synthesized using a MEGAscript kit (Life Technologies) and a PCR fragment containing a T7 promoter derived from primer HY1. A DNA fragment containing csrR sequence from +3 to +84 relative to the start of transcription was used as the template. Gel-purified RNA was 5′ end labeled with [γ-32P]ATP. RNA suspended in Tris-EDTA (TE) buffer was heated to 90°C for 1 min followed by slow cooling to room temperature. Binding reaction mixtures (10 µl) contained 10 mM Tris-HCl (pH 7.5), 10 mM MgCl2, 100 mM KCl, 200 ng/µl yeast RNA, 0.2 mg/ml bovine serum albumin (BSA), 7.5% glycerol, 20 mM dithiothreitol (DTT), 0.1 nM csrR RNA, CsrA-H6 (various concentrations), and 0.1 mg/ml xylene cyanol. Competition assay mixtures also contained unlabeled competitor RNA. To allow CsrA-RNA complex formation, reaction mixtures were incubated for 30 min at 37°C. Samples were then fractionated through native 15% polyacrylamide gels using 0.5× Tris-borate-EDTA (TBE) running buffer. Radioactive bands were visualized with a phosphorimager and quantified using ImageQuant 5.2 software. The apparent equilibrium binding constant (Kd) of CsrA-csrR RNA interaction was calculated as described previously (58).

RNA footprint assay.

RNA was synthesized using a MEGAscript kit and a PCR fragment containing a T7 promoter derived from primer HY3. The template was a DNA fragment containing csrR sequence extending from 32 to +84 relative to the start of transcription. Binding reactions (10 µl) containing 5 nM csrR RNA and the indicated concentrations of CsrA-H6 were otherwise identical to those in the gel shift assay. After the initial binding reaction, 0.075 U RNase T1 (Roche) was added to the reaction mixtures, and incubation continued for 15 min at 37°C. Reactions were terminated by addition of 10 µl of gel loading buffer, and reaction mixtures were placed on ice. Partial alkaline hydrolysis and RNase T1 digestion ladders of each transcript were prepared as described previously (59). After samples were fractionated through the use of standard 6% polyacrylamide sequencing gels, radioactive bands were visualized with a phosphorimager.

Construction and assays of GFP reporters.

To construct the csrR-GFP BS2-only transcriptional reporter, ~300 bp immediately 5′ of the csrR coding region was amplified from Lp02 genomic DNA using primers ZA49 and ZA50, ligated directly 5′ of GFPmut3 in vector pBH6119, a derivative of pJB98 that encodes a promoterless GFPmut3 locus (39), and then electroporated into Lp02 to generate strain MB1375. For the csrR-GFP BS2/BS3 translational reporter, the sequence incorporated into the BS2 reporter was extended to include the first 5 codons of csrR. After PCR amplification using primers ZA49 and ZA73, the fragment was fused directly 5′ and in frame with the second codon of GFPmut3 amplified from pBH6119 using primers ZA72 and ZA84 via the SOE PCR strategy using primers ZA49 and ZA84. This product was then ligated into pBH6119, replacing the promoterless GFPmut3, and electroporated into Lp02 to generate strain MB1376. To construct the csrR-GFP BS2-sld/BS3 translational reporter, the stem-loop structure of BS2 predicted to favor CsrA binding was disrupted by changing the six consecutive adenines immediately 5′ of the SD sequence to thymidines in the DNA sequence (resulting in uracils in the RNA). To do so, the mutation was inserted into two overlapping PCR products using pcsrR-GFP_BS2/BS3 as the template, primer pair ZA49 and ZA75, and primer pair ZA74 and ZA84. The two products were then joined via SOE PCR using primers ZA49 and ZA84 to generate the BS2 mutant allele, which was then ligated into pBH6119, replacing the promoterless GFPmut3, and electroporated into Lp02 to produce strain MB1377. For fluorescence experiments, colonies of MB1375 to MB1377 were cultured in broth to the E phase and then diluted to an OD600 of 0.1. Aliquots were collected at the times shown and normalized to an OD600 of 1, and their fluorescence was quantified. Strain MB355, containing pflaA-GFP, served as a marker of the entry into the PE phase, and strain BH006 carrying pBH6119, encoding GFP but no promoter, was the negative control (39). Growth levels of all strains were similar, as assessed by OD600 readings throughout the experiment (data not shown).

Inducible CsrR constructs.

Wild-type csrR was amplified from Lp02 genomic DNA using primers ZA59 and ZA60 and the PCR product ligated into p206cam (60). This plasmid was electroporated into Lp02 to generate strain MB1378.

The BS3 mutant allele of csrR was constructed by SOE PCR. Two PCR products were generated. Product 1 started ~500 bp 5′ of the transcriptional start generated by forward primer ZA37 and included a GAT-to-TTG mutation in the second codon of csrR that was inserted using reverse primer ZA83. Product 2 encoded the csrR coding region, included the GAT-to-TTG mutation in the second codon that was inserted by the forward primer ZA82, and extended ~500 bp 3′ of the coding region by using reverse primer ZA38. One microliter of each of these products then served as the template in a SOE PCR using outermost primers ZA37 and ZA38. Next, this PCR product containing the BS3 mutation was used as the template to amplify a His-tagged version of only the coding region, using primers ZA59 and ZA60. Finally, the product was ligated into p206cam. This plasmid was electroporated into Lp02 to generate strain MB1379.

Rifampin and RNA stability experiment.

The conditional csrA mutant strain MB464 was cultured on CYE agar supplemented with 200 µM IPTG and then inoculated into AYET medium containing 200 µM IPTG. After culture to the E phase, cells were collected by centrifugation, resuspended in AYET medium without IPTG, and then divided between two tubes that contained or lacked 200 µM IPTG. These samples were then cultured to an OD600 of ~1 and aliquots collected for RNA isolation. Next, 100 µg/ml rifampin was added and cultures were incubated for 15 min before a second aliquot was collected for RNA isolation.

qRT-PCR.

cDNA was generated using 1 µl of isolated RNA in a 20-µl reaction mixture with an iScript cDNA synthesis kit from Bio-Rad following the manufacturer’s protocol. A 1:50 dilution of cDNA served as the template for qRT-PCR in 20-µl reaction mixtures using iQ SYBR green Supermix (Bio-Rad). Primers ZA53 and ZA54 were used to assess csrR mRNA levels, which were normalized to the internal control 16S rRNA amplified using primers SL1 and SL2 and threshold cycle (ΔΔCT) analysis.

Western analysis.

Strains MB1378 and MB1379 were inoculated from colonies and cultured to the E phase, diluted to an OD600 of 0.05, and divided into two tubes containing or lacking 200 µM IPTG. After incubation for ~15 h to the E phase, cultures were normalized to an OD600 of 10 by centrifugation of aliquots at 13,000 rpm for 2 min, and then cells were resuspended in 100 µl of Laemmli buffer (2% SDS, 10% glycerol, 5% 2-mercaptoethanol, 0.005% bromophenol blue, 62.5 mM Tris-HCl, pH 6.8). Samples were then lysed by boiling for 5 min and debris pelleted at 13,000 rpm for 2 min. Proteins were separated on a 12% mini-Protean TGX precast gel (Bio-Rad); a Precision Plus Kaleidoscope protein standard ladder (Bio-Rad) was used as a size marker and to verify transfer. To determine relative loading levels and transfer of each sample, Ponceau S (Fisher) staining was performed on a duplicate membrane. CsrR was detected via the 6×His epitope that was inserted in frame at its carboxy terminus using a 1:5,000 dilution of anti-His (C-terminal)-horseradish peroxidase (HRP) antibody (Novex; Life Technologies) and SuperSignal West Pico chemiluminescent substrate (Thermo Scientific). A His-tagged derivative of CsrA paralog Lpg2094 was the positive control.

SUPPLEMENTAL MATERIAL

Legionella strains surveyed for csrA-like genes, with NCBI taxonomy identification numbers indicated. TaxID, taxonomy identification number. The superscript a indicates previously published data (C. Rao, H. Benhabib, and A. W. Ensminger, PLoS One 8:e64129, 2013).

Table S1, DOCX file, 0.1 MB

Each gene is listed with its locus tag identifier in the order in which it is found in the heat map (Fig. 1B). Canonical CsrA (Lpg0780) and CsrR (Lpg1593) are marked in bold.

Table S2, DOCX file, 0.1 MB

The superscript a indicates a gene name in one selected strain. Most have unique gene annotations in each strain. The superscript b indicates the size (in nucleotides) of the putative translated region. The superscript c indicates the prevalence in each of the 14 fully sequenced strains of legionellae surveyed, as available via NCBI BLAST. The superscript d indicates putative integrative conjugative elements (ICEs) determined by the presence of type IVa secretion system genes immediately adjacent to the csrA-like gene. The superscript e indicates canonical csrA, as characterized previously (13).

Table S3, DOCX file, 0.1 MB

Strains, plasmids, and primers used in this study.

Table S4, DOCX file, 0.1 MB

Citation Abbott ZD, Yakhnin H, Babitzke P, Swanson MS. 2015. csrR, a paralog and direct target of CsrA, promotes Legionella pneumophila resilience in water. mBio 6(3):e00595-15. doi:10.1128/mBio.00595-15.

ACKNOWLEDGMENTS

We thank Brenda Byrne and Kaitlin Flynn for many conversations and suggestions and for copious negative phenotypic data not shown here; Nicole Koropatkin for the Rosetta(DE3) pLysS strain, high-performance liquid chromatography (HPLC) equipment, and protein purification guidance; Paul Carlson for qRT-PCR guidance; and Mary O’Riordan for sharing equipment. We are grateful to Nicholas Cianciotto (Northwestern University) and Zhao-Qing Luo (Purdue University) for the kind gifts of plasmid pSU2719 and pZL790, respectively. And we especially thank the French Canadian, Sébastien Crépin, for technical guidance, construction of the pSU-pZLKm plasmid, and writing the relevant Materials and Methods section. Finally, we thank Zach Dalebroux for helping launch this project.

This research was supported by the Endowment for the Basic Sciences at the University of Michigan Medical School, a University of Michigan Rackham Merit Fellowship (Z.D.A.), the Cellular Biotechnology Training Program (T32 GM008353; Z.D.A.), and NIH grant GM059969 (P.B.).

REFERENCESMercanteJ, SuzukiK, ChengX, BabitzkeP, RomeoT 2006 Comprehensive alanine-scanning mutagenesis of Escherichia coli CsrA defines two subdomains of critical functional importance. J Biol Chem 281:3183231842. doi:10.1074/jbc.M606057200.16923806TimmermansJ, Van MelderenL 2010 Post-transcriptional global regulation by CsrA in bacteria. Cell Mol Life Sci 67:28972908. doi:10.1007/s00018-010-0381-z.20446015VakulskasCA, PottsAH, BabitzkeP, AhmerBMM, RomeoT 2015 Regulation of bacterial virulence by Csr (Rsm) systems. Microbiol Mol Biol Rev 79:193224. doi:10.1128/MMBR.00052-14.25833324BakerCS, MorozovI, SuzukiK, RomeoT, BabitzkeP 2002 CsrA regulates glycogen biosynthesis by preventing translation of glgC in Escherichia coli. Mol Microbiol 44:15991610. doi:10.1046/j.1365-2958.2002.02982.x.12067347BhattS, EdwardsAN, NguyenHT, MerlinD, RomeoT, KalmanD 2009 The RNA binding protein CsrA is a pleiotropic regulator of the locus of enterocyte effacement pathogenicity island of enteropathogenic Escherichia coli. Infect Immun 77:35523568. doi:10.1128/IAI.00418-09.19581394PannuriA, YakhninH, VakulskasCA, EdwardsAN, BabitzkeP, RomeoT 2012 Translational repression of NhaR, a novel pathway for multi-tier regulation of biofilm circuitry by CsrA. J Bacteriol 194:7989. doi:10.1128/JB.06209-11.22037401DubeyAK, BakerCS, SuzukiK, JonesAD, PanditP, RomeoT, BabitzkeP 2003 CsrA regulates translation of the Escherichia coli carbon starvation gene, cstA, by blocking ribosome access to the cstA transcript. J Bacteriol 185:44504460. doi:10.1128/JB.185.15.4450-4460.2003.12867454BrencicA, LoryS 2009 Determination of the regulon and identification of novel mRNA targets of Pseudomonas aeruginosa RsmA. Mol Microbiol 72:612632. doi:10.1111/j.1365-2958.2009.06670.x.19426209YakhninH, PanditP, PettyTJ, BakerCS, RomeoT, BabitzkeP 2007 CsrA of Bacillus subtilis regulates translation initiation of the gene encoding the flagellin protein (hag) by blocking ribosome binding. Mol Microbiol 64:16051620. doi:10.1111/j.1365-2958.2007.05765.x.17555441JonasK, EdwardsAN, AhmadI, RomeoT, RömlingU, MeleforsO 2010 Complex regulatory network encompassing the Csr, c-di-GMP and motility systems of Salmonella Typhimurium. Environ Microbiol 12:524540. doi:10.1111/j.1462-2920.2009.02097.x.19919539NevoO, ZusmanT, RasisM, LifshitzZ, SegalG 2014 Identification of Legionella pneumophila effectors regulated by the LetAS-RsmYZ-CsrA regulatory cascade, many of which modulate vesicular trafficking. J Bacteriol 196:681692. doi:10.1128/JB.01175-13.24272784YakhninH, BakerCS, BerezinI, EvangelistaMA, RassinA, RomeoT, BabitzkeP 2011 CsrA represses translation of sdiA, which encodes the N-acylhomoserine-l-lactone receptor of Escherichia coli, by binding exclusively within the coding region of sdiA mRNA. J Bacteriol 193:61626170. doi:10.1128/JB.05975-11.21908661MolofskyAB, SwansonMS 2003 Legionella pneumophila CsrA is a pivotal repressor of transmission traits and activator of replication. Mol Microbiol 50:445461. doi:10.1046/j.1365-2958.2003.03706.x.14617170OhnoA, KatoN, SakamotoR, KimuraS, YamaguchiK 2008 Temperature-dependent parasitic relationship between Legionella pneumophila and a free-living amoeba (Acanthamoeba castellanii). Appl Environ Microbiol 74:45854588. doi:10.1128/AEM.00083-08.18502936Forsbach-BirkV, McNealyT, ShiC, LynchD, MarreR 2004 Reduced expression of the global regulator protein CsrA in Legionella pneumophila affects virulence-associated regulators and growth in Acanthamoeba castellanii. Int J Med Microbiol 294:1525. doi:10.1016/j.ijmm.2003.12.003.15293450ByrneB, SwansonMS 1998 Expression of Legionella pneumophila virulence traits in response to growth conditions. Infect Immun 66:30293034.9632562RasisM, SegalG 2009 The LetA-RsmYZ-CsrA regulatory cascade, together with RpoS and PmrA, post-transcriptionally regulates stationary phase activation of Legionella pneumophila Icm/Dot effectors. Mol Microbiol 72:9951010. doi:10.1111/j.1365-2958.2009.06705.x.19400807EdwardsRL, DalebrouxZD, SwansonMS 2009 Legionella pneumophila couples fatty acid flux to microbial differentiation and virulence. Mol Microbiol 71:11901204. doi:10.1111/j.1365-2958.2009.06593.x.19170883SahrT, BrüggemannH, JulesM, LommaM, Albert-WeissenbergerC, CazaletC, BuchrieserC 2009 Two small ncRNAs jointly govern virulence and transmission in Legionella pneumophila. Mol Microbiol 72:741762. doi:10.1111/j.1365-2958.2009.06677.x.19400772HammerBK, TatedaES, SwansonMS 2002 A two-component regulator induces the transmission phenotype of stationary-phase Legionella pneumophila. Mol Microbiol 44:107118. doi:10.1046/j.1365-2958.2002.02884.x.11967072RowbothamTJ 1980 Preliminary report on the pathogenicity of Legionella pneumophila for freshwater and soil amoebae. J Clin Pathol 33:11791183. doi:10.1136/jcp.33.12.1179.7451664FieldsBS 1996 The molecular ecology of legionellae. Trends Microbiol 4:286290. doi:10.1016/0966-842X(96)10041-X.8829338CohnPD, GleasonJA, RudowskiE, TsaiSM, GeneseCA, FaglianoJA 2015 Community outbreak of legionellosis and an environmental investigation into a community water system. Epidemiol Infect 143:13221331. doi:10.1017/S0950268814001964.25083716QinT, ZhouH, RenH, GuanH, LiM, ZhuB, ShaoZ 2014 Distribution of sequence-based types of Legionella pneumophila serogroup 1 strains isolated from cooling towers, hot springs, and potable water systems in China. Appl Environ Microbiol 80:21502157. doi:10.1128/AEM.03844-13.24463975OsawaK, ShigemuraK, AbeY, JikimotoT, YoshidaH, FujisawaM, ArakawaS 2014 A case of nosocomial Legionella pneumoniae associated with a contaminated hospital cooling tower. J Infect Chemother 20:6870. doi:10.1016/j.jiac.2013.07.007.24462430HorwitzMA, SilversteinSC 1980 Legionnaires’ disease bacterium (Legionella pneumophila) multiples intracellularly in human monocytes. J Clin Invest 66:441450. doi:10.1172/JCI109874.7190579MurgaR, ForsterTS, BrownE, PrucklerJM, FieldsBS, DonlanRM 2001 Role of biofilms in the survival of Legionella pneumophila in a model potable-water system. Microbiology 147:31213126.11700362DeclerckP, BehetsJ, MargineanuA, van HoefV, De KeersmaeckerB, OllevierF 2009 Replication of Legionella pneumophila in biofilms of water distribution pipes. Microbiol Res 164:593603. doi:10.1016/j.micres.2007.06.001.17644359SteinertM, EmödyL, AmannR, HackerJ 1997 Resuscitation of viable but nonculturable Legionella pneumophila Philadelphia JR32 by Acanthamoeba castellanii. Appl Environ Microbiol 63:20472053.9143134Al-BanaBH, HaddadMT, GarduñoRA 2014 Stationary phase and mature infectious forms of Legionella pneumophila produce distinct viable but non-culturable cells. Environ Microbiol 16:382395. doi:10.1111/1462-2920.12219.23968544GarduñoRA, GarduñoE, HiltzM, HoffmanPS 2002 Intracellular growth of Legionella pneumophila gives rise to a differentiated form dissimilar to stationary-phase forms. Infect Immun 70:62736283. doi:10.1128/IAI.70.11.6273-6283.2002.12379706ChienM, MorozovaI, ShiS, ShengH, ChenJ, GomezSM, AsamaniG, HillK, NuaraJ, FederM, RineerJ, GreenbergJJ, SteshenkoV, ParkSH, ZhaoB, TeplitskayaE, EdwardsJR, PampouS, GeorghiouA, ChouIC, IannuccilliW, UlzME, KimDH, Geringer-SamethA, GoldsberryC, MorozovP, FischerSG, SegalG, QuX, RzhetskyA, ZhangP, CayanisE, De JongPJ, JuJ, KalachikovS, ShumanHA, RussoJJ 2004 The genomic sequence of the accidental pathogen Legionella pneumophila. Science 305:19661968. doi:10.1126/science.1099776.15448271Doléans-JordheimA, AkermiM, GinevraC, CazaletC, KayE, SchneiderD, BuchrieserC, AtlanD, VandeneschF, EtienneJ, JarraudS 2006 Growth-phase-dependent mobility of the lvh-encoding region in Legionella pneumophila strain Paris. Microbiology 152:35613568. doi:10.1099/mic.0.29227-0.17159208FlynnKJ, SwansonMS 2014 Integrative conjugative element ICE-βox confers oxidative stress resistance to Legionella pneumophila in vitro and in macrophages. mBio 5:e01091-14. doi:10.1128/mBio.01091-14.24781744WeeBA, WoolfitM, BeatsonSA, PettyNK 2013 A distinct and divergent lineage of genomic island-associated type IV secretion systems in Legionella. PLoS One 8:e82221. doi:10.1371/journal.pone.0082221.24358157MardenJN, DiazMR, WaltonWG, GodeCJ, BettsL, UrbanowskiML, RedinboMR, YahrTL, WolfgangMC 2013 An unusual CsrA family member operates in series with RsmA to amplify posttranscriptional responses in Pseudomonas aeruginosa. Proc Natl Acad Sci U S A 110:1505515060. doi:10.1073/pnas.1307217110.23980177GiãoMS, WilksSA, AzevedoNF, VieiraMJ, KeevilCW 2009 Validation of SYTO 9/propidium iodide uptake for rapid detection of viable but noncultivable Legionella pneumophila. Microb Ecol 58:5662. doi:10.1007/s00248-008-9472-x.19043657DubeyAK, BakerCS, RomeoT, BabitzkeP 2005 RNA sequence and secondary structure participate in high-affinity CsrA-RNA interaction. RNA 11:15791587. doi:10.1261/rna.2990205.16131593HammerBK, SwansonMS 1999 Co-ordination of Legionella pneumophila virulence with entry into stationary phase by ppGpp. Mol Microbiol 33:721731. doi:10.1046/j.1365-2958.1999.01519.x.10447882SahrT, RusniokC, Dervins-RavaultD, SismeiroO, CoppeeJY, BuchrieserC 2012 Deep sequencing defines the transcriptional map of L. pneumophila and identifies growth phase-dependent regulated ncRNAs implicated in virulence. RNA Biol 9:503519. doi:10.4161/rna.20270.22546937LiuMY, YangH, RomeoT 1995 The product of the pleiotropic Escherichia coli gene csrA modulates glycogen biosynthesis via effects on mRNA stability. J Bacteriol 177:26632672.7751274WangX, DubeyAK, SuzukiK, BakerCS, BabitzkeP, RomeoT 2005 CsrA post-transcriptionally represses pgaABCD, responsible for synthesis of a biofilm polysaccharide adhesin of Escherichia coli. Mol Microbiol 56:16481663. doi:10.1111/j.1365-2958.2005.04648.x.15916613KimC, JeonS, JungJ, OhY, KimY, LeeJ, ChoiS, ChaeY, LeeYK 2015 Isolation of Legionella pneumophila from cooling towers, public baths, hospitals, and fountains in Seoul, Korea, from 2010 to 2012. J Environ Health 77:5862.25619037Patterson-FortinLM, VakulskasCA, YakhninH, BabitzkeP, RomeoT 2013 Dual posttranscriptional regulation via a cofactor-responsive mRNA leader. J Mol Biol 425:36623677. doi:10.1016/j.jmb.2012.12.010.23274138WeiBL, Brun-ZinkernagelAM, SimeckaJW, PrüssBM, BabitzkeP, RomeoT 2001 Positive regulation of motility and flhDC expression by the RNA-binding protein CsrA of Escherichia coli. Mol Microbiol 40:245256. doi:10.1046/j.1365-2958.2001.02380.x.11298291MukherjeeS, BabitzkeP, KearnsDB 2013 FliW and FliS function independently to control cytoplasmic flagellin levels in Bacillus subtilis. J Bacteriol 195:297306. doi:10.1128/JB.01654-12.23144244YakhninAV, BakerCS, VakulskasCA, YakhninH, BerezinI, RomeoT, BabitzkeP 2013 CsrA activates flhDC expression by protecting flhDC mRNA from RNase E-mediated cleavage. Mol Microbiol 87:851866. doi:10.1111/mmi.12136.23305111RoyCR 2002 The Dot/lcm transporter of Legionella pneumophila: a bacterial conductor of vesicle trafficking that orchestrates the establishment of a replicative organelle in eukaryotic hosts. Int J Med Microbiol 291:463467.11890545EdwardsRL, JulesM, SahrT, BuchrieserC, SwansonMS 2010 The Legionella pneumophila LetA/LetS two-component system exhibits rheostat-like behavior. Infect Immun 78:25712583. doi:10.1128/IAI.01107-09.20351136DalebrouxZD, YagiBF, SahrT, BuchrieserC, SwansonMS 2010 Distinct roles of ppGpp and DksA in Legionella pneumophila differentiation. Mol Microbiol 76:200219. doi:10.1111/j.1365-2958.2010.07094.x.20199605FeeleyJC, GibsonRJ, GormanGW, LangfordNC, RasheedJK, MackelDC, BaineWB 1979 Charcoal-yeast extract agar: primary isolation medium for Legionella pneumophila. J Clin Microbiol 10:437441.393713DatsenkoKA, WannerBL 2000 One-step inactivation of chromosomal genes in Escherichia coli K-12 using PCR products. Proc Natl Acad Sci U S A 97:66406645. doi:10.1073/pnas.120163297.10829079SextonJA, VogelJP 2004 Regulation of hypercompetence in Legionella pneumophila. J Bacteriol 186:38143825. doi:10.1128/JB.186.12.3814-3825.2004.15175295LiuY, GaoP, BangaS, LuoZQ 2008 An in vivo gene deletion system for determining temporal requirement of bacterial virulence factors. Proc Natl Acad Sci U S A 105:93859390. doi:10.1073/pnas.0801055105.18599442MartinezE, BartoloméB, de la CruzF 1988 pACYC184-derived cloning vectors containing the multiple cloning site and lacZ alpha reporter gene of pUC8/9 and pUC18/19 plasmids. Gene 68:159162. doi:10.1016/0378-1119(88)90608-7.2851489BergerKH, IsbergRR 1993 Two distinct defects in intracellular growth complemented by a single genetic locus in Legionella pneumophila. Mol Microbiol 7:719. doi:10.1111/j.1365-2958.1993.tb01092.x.8382332TrouttAB, McHeyzer-WilliamsMG, PulendranB, NossalGJ 1992 Ligation-anchored PCR: a simple amplification technique with single-sided specificity. Proc Natl Acad Sci U S A 89:98239825. doi:10.1073/pnas.89.20.9823.1409706YakhninAV, TrimbleJJ, ChiaroCR, BabitzkeP 2000 Effects of mutations in the l-tryptophan binding pocket of the Trp RNA-binding attenuation protein of Bacillus subtilis. J Biol Chem 275:45194524. doi:10.1074/jbc.275.6.4519.10660627BevilacquaJM, BevilacquaPC 1998 Thermodynamic analysis of an RNA combinatorial library contained in a short hairpin. Biochemistry 37:1587715884. doi:10.1021/bi981732v.9843393MoralesVM, BäckmanA, BagdasarianM 1991 A series of wide-host-range low-copy-number vectors that allow direct screening for recombinants. Gene 97:3947. doi:10.1016/0378-1119(91)90007-X.1847347