Skip to main content
Advertisement
  • Loading metrics

Succinate utilisation by Salmonella is inhibited by multiple regulatory systems

  • Nicolas Wenner ,

    Roles Conceptualization, Data curation, Formal analysis, Funding acquisition, Investigation, Methodology, Project administration, Supervision, Validation, Visualization, Writing – original draft, Writing – review & editing

    nicolas.wenner@gmail.com

    Current address: Biozentrum, University of Basel, Basel, Switzerland

    Affiliation Clinical Infection, Microbiology & Immunology, Institute of Infection, Veterinary & Ecological Sciences, University of Liverpool, Liverpool, United Kingdom

  • Xiaojun Zhu,

    Roles Conceptualization, Formal analysis, Investigation, Methodology, Validation, Visualization, Writing – review & editing

    Affiliation Clinical Infection, Microbiology & Immunology, Institute of Infection, Veterinary & Ecological Sciences, University of Liverpool, Liverpool, United Kingdom

  • Will P. M. Rowe,

    Roles Formal analysis, Investigation

    Affiliation Clinical Infection, Microbiology & Immunology, Institute of Infection, Veterinary & Ecological Sciences, University of Liverpool, Liverpool, United Kingdom

  • Kristian Händler,

    Roles Investigation

    Current address: Institute of Human Genetics, University Hospital of Schleswig-Holstein, University of Lübeck, Lübeck, Germany

    Affiliation Department of Microbiology, School of Genetics and Microbiology, Moyne Institute of Preventive Medicine, Trinity College Dublin, Dublin, Ireland

  • Jay C. D. Hinton

    Roles Conceptualization, Funding acquisition, Methodology, Project administration, Resources, Supervision, Writing – review & editing

    Affiliation Clinical Infection, Microbiology & Immunology, Institute of Infection, Veterinary & Ecological Sciences, University of Liverpool, Liverpool, United Kingdom

Abstract

Succinate is a potent immune signalling molecule that is present in the mammalian gut and within macrophages. Both of these infection niches are colonised by the pathogenic bacterium Salmonella enterica serovar Typhimurium during infection. Succinate is a C4-dicarboyxlate that can serve as a source of carbon for bacteria. When succinate is provided as the sole carbon source for in vitro cultivation, Salmonella and other enteric bacteria exhibit a slow growth rate and a long lag phase. This growth inhibition phenomenon was known to involve the sigma factor RpoS, but the genetic basis of the repression of bacterial succinate utilisation was poorly understood. Here, we use an experimental evolution approach to isolate fast-growing mutants during growth of S. Typhimurium on succinate containing minimal medium. Our approach reveals novel RpoS-independent systems that inhibit succinate utilisation. The CspC RNA binding protein restricts succinate utilisation, an inhibition that is antagonised by high levels of the small regulatory RNA (sRNA) OxyS. We discovered that the Fe-S cluster regulatory protein IscR inhibits succinate utilisation by repressing the C4-dicarboyxlate transporter DctA. Furthermore, the ribose operon repressor RbsR is required for the complete RpoS-driven repression of succinate utilisation, suggesting a novel mechanism of RpoS regulation. Our discoveries shed light on the redundant regulatory systems that tightly regulate the utilisation of succinate. We speculate that the control of central carbon metabolism by multiple regulatory systems in Salmonella governs the infection niche-specific utilisation of succinate.

Author summary

During infection, Salmonella enterica serovar Typhimurium must efficiently utilise the nutrients provided by the host and the microbiota. Within host niches colonised by Salmonella, key carbon source is succinate, which is abundant in both the gut and within macrophages. Although S. Typhimurium possesses all the molecular machinery to import and catabolise succinate, the utilisation of this carbon source is strongly inhibited under laboratory conditions. Here, we have used diverse genetic approaches to discover that several distinct genetic regulatory systems act in concert to repress succinate utilisation by Salmonella. Our study sheds new light on how carbon source utilisation is managed by this important model enteropathogenic bacterium.

Introduction

Metabolic versatility is a key property that allows pathogenic enteric bacteria to thrive both during infection of mammals and in the wider environment [1]. C4-dicarboyxlates are an important part of the bacterial catabolic repertoire, which can be utilised as a sole carbon and energy source (C-source) [2]. In the mammalian gut, the C4-dicarboyxlate succinate is an abundant C-source that is provided by the microbiota in response to the presence of dietary fibre [3]. Salmonella enterica serovar Typhimurium (S. Typhimurium) is one of the best understood enteropathogenic bacterium [4] which efficiently catabolises succinate during intestinal colonisation to enhance the growth [5].

As well as colonising the mammalian gut, Salmonella can also cross the intestinal epithelial barrier and invade several types of tissues [6]. An important element of the pathogenic lifestyle of S. Typhimurium involves the hijacking of macrophages, and the intracellular proliferation of the bacteria within Salmonella-containing vacuoles (SCVs) [7]. Recently, it has been discovered that macrophages undergo metabolic reprogramming during bacterial infection, leading to the build-up of tricarboxylic acid (TCA) cycle intermediates, including succinate [8,9]. This C4-dicarboxylate acts as an important proinflammatory molecule that is also involved in hypoxic and metabolic signalling [10,11]. Following infection by S. Typhimurium, high levels of succinate accumulate within macrophages [12]. However, Salmonella does not use this succinate to fuel growth as glucose and the glycolytic intermediate 3-phosphoglycerate are the key intra-macrophage C-sources [1214]

The inactivation of key succinate catabolic genes does not reduce the ability of S. Typhimurium to replicate in murine macrophages, but stimulates intracellular proliferation [15]. Although succinate is not utilised as a C-source by Salmonella in the SCV, the metabolite does act as a crucial signal molecule for the induction of Salmonella genes associated with survival and virulence within macrophages [14].

During in vitro cultivation, S. Typhimurium exhibits a particularly extended lag phase in minimal axenic media containing succinate as sole C-source; in contrast, succinate supports the rapid growth of other enteric bacteria such as Escherichia coli [16], via the succinate dehydrogenase (SDH) multi-enzyme complex that oxidises succinate into fumarate [17]. Subsequent, bacterial replication with succinate involves the generation of all cellular components via gluconeogenesis [2,18].

The stress response sigma factor σ38 (RpoS) is a global transcriptional regulator that modulates diverse facets of Salmonella biology including stress-resistance, immobilised growth, virulence and nutrient assimilation [1922]. RpoS inhibits in vitro growth upon succinate by repressing transcription of the sdhCDAB operon (sdh) and other TCA cycle genes [2325].

Because Salmonella utilises succinate for colonisation of the inflamed gut [5] but not for intra-macrophage proliferation [14,15], we hypothesised that Salmonella had evolved multiple genetic regulatory mechanisms to tightly control the niche-dependent utilisation of this infection-relevant molecule.

Here, we devised an in vitro experimental strategy to search for novel regulatory mechanisms involved in the modulation of succinate utilisation under aerobic conditions. This type of experimental evolution has a successful track record [26] and has previously been used to generate Salmonella mutants that grow on novel carbon sources [27]. Our subsequent genetic dissection identified two novel RpoS-independent regulatory mechanisms that repress succinate utilisation via the CspC and IscR regulatory proteins. In addition, the modulation of RpoS activity by RbsR also impacted upon succinate utilisation. We propose that this multi-factorial system ensures that succinate is only catabolised at the right place and at the right time during infection to permit effective niche adaptation.

Results and discussion

Growth inhibition and experimental evolution in succinate minimal medium

Enteric bacteria possess the catabolic enzymes and efficient uptake systems required to grow with C4-dicarboxylates as sole C-source [2,28]. However, some environmental and clinical isolates of Escherichia coli and Salmonella have surprisingly slow growth rates in succinate-containing minimal media [2931]. To investigate this phenomenon in pathogenic and non-pathogenic enteric bacteria, we assessed the growth of four bacterial species on agar plates containing succinate as a sole C-source (M9+Succ). We studied growth for up to 96 hours, and used a variety of laboratory strains with an emphasis on Salmonella (Fig 1).

thumbnail
Fig 1. Inhibition of growth of most Salmonella serovars and K. pneumoniae on succinate minimal medium.

The indicated strains of E. coli, Salmonella, Citrobacter and Klebsiella were spread on M9+Succ agar plates and incubated at 37°C. Photographs of bacterial growth were taken every 24 hours for 2 days or 4 days. For Salmonella enterica isolates, the serovar and the clade A/B status [35,36] are indicated by the colour of the picture frame. Experiments were carried out as biological triplicates, and a representative picture is shown for each strain.

https://doi.org/10.1371/journal.pgen.1011142.g001

For S. enterica serovars Typhimurium and Enteritidis, we tested the growth of the well-characterised S. Typhimurium strains LT2, 4/74 and 14028 and of the S. Enteritidis strain P125109. In addition, we assessed the growth of multidrug resistant S. Typhimurium ST313 strain D23580 and S. Enteritidis strain D7795, two representative strains that cause invasive non-typhoidal Salmonella disease in Africa [3234]. We included the reptile-associated Salmonella serovars Soahanina, Hadar, Newport and Infantis which belong to the two metabolically-distinct clades A and B of S. enterica [35,36]. We and others have previously generated genome sequences of most of the Salmonella strains that were used [34,35,37]. We also tested a multidrug resistant strain of Klebsiella pneumoniae (strain KP52.145) [38], a Citrobacter rodentium strain (ATCC51459) [39] and the classic laboratory strain E. coli K-12 strain MG1655.

After 24 hours of incubation at 37°C, the only strains that displayed substantial growth with succinate as a C-source were S. Typhimurium LT2, E. coli MG1655 and C. rodentium ATCC51459. The growth of the S. Typhimurium laboratory strain LT2 upon succinate at 24h is the result of the mutated variant of rpoS carried by this strain [40], which explains its faster growth [16]. We found that growth on succinate is a serovar-dependent phenotype. Of the six Salmonella serovars that were tested, only S. Newport displayed growth after 2 days. Subsequently, following 2–3 days of incubation, large isolated colonies were observed within the bacterial lawns of the K. pneumoniae strain and the other Salmonella isolates.

Previous experiments in liquid minimal medium containing succinate as sole C-source showed that S. Typhimurium exhibited a particularly long lag phase [16,23]. This extended lag time could reflect a particularly slow metabolic remodelling, preparing Salmonella for the exponential phase in the presence of succinate. Alternatively, robust inhibition of succinate assimilation might be occurring under these conditions, preventing growth until spontaneous fast-growing mutants (hereafter referred as Succ+ mutants) have emerged.

To test these hypotheses, we used an experimental evolution approach to generate Succ+ mutants. We assessed the growth of the well-characterised S. Typhimurium strain 4/74 (henceforth referred to as 4/74 or Salmonella) in liquid M9+Succ media inoculated with a stationary phase culture made in rich medium (LB). The four independent 4/74 cultures (cultures I-IV) exhibited the reported 30–35 hour lag time at 37°C [16,23] (Fig 2A). We collected the Salmonella that eventually reached stationary phase and cultured the bacteria in LB for two passages before re-inoculation in M9+Succ. For all the succinate-evolved cultures, the lag time in M9+Succ was shortened to 4–5 hours, and stationary phase was reached after approximately 14 hours (Fig 2B).

thumbnail
Fig 2. Experimental evolution of S. Typhimurium in succinate minimal medium.

(A) Growth of S. Typhimurium 4/74 displays an extended lag time in M9+Succ medium. The growth curves of four independent cultures (I-IV) of 4/74 WT in M9+Succ are presented. Cultures were inoculated with bacteria grown beforehand to stationary phase in LB. the succinate evolved bacteria were harvested in ESP (50 μL of culture) and were grown twice in LB prior to re-inoculation in M9+Succ. (B) The succinate-evolved bacteria grow fast in M9+Succ in comparison with the WT strain. Growth curves of the succinate evolved cultures I-IV from (A) in M9+Succ are presented: the 4/74 WT and ΔrpoS (JH3674) strains were included as controls. (C) Succinate fast growing (Succ+) mutants were detected in liquid M9+Succ 4/74 cultures. Cultures from (A) were spread on LB plates at the indicated growth phase and growth with succinate was assessed for 15 isolated colonies per replicate on M9+Succ agar plates after 48 hours of incubation. The graph shows the proportion (%) of Succ+ clones. ND = not detected. (D) Succ+ spontaneous mutants emerge from 4/74 WT bacterial lawns on M9+Succ agar plates. 4/74 WT cultures (~107 CFU) were spread on a M9+Succ agar plates and the picture of a representative plate was taken after 3 days of incubation at 37°C. For the growth curves (A&B), bacteria were grown at 37°C with aeration in 25 ml of M9+Succ (in 250 ml conical flasks) with an initial inoculum of ~107 CFU/mL (OD600 = 0.01). Growth phases are indicated in (A&C): Lag phase (Lag); Early exponential phase (EEP); Mid-exponential phase (MEP); Early stationary phase (ESP).

https://doi.org/10.1371/journal.pgen.1011142.g002

To investigate heritability of the succinate growth phenotype, the initial M9+Succ cultures (Fig 2A) were spread on LB agar plates at different stages of growth, and isolated colonies were tested on M9+Succ plates. Of 60 colonies obtained from lag phase, none grew faster than the wild type (Fig 2C). However, fast growing Succ+ mutants harvested from early exponential, mid-exponential and early stationary phase culture were detected at a frequency of 20%, 78% and 90%, respectively (Fig 2C). When ~107 Colony Forming Units (CFU) of 4/74 wild-type (WT) were spread on M9+Succ plates, between 100 and 1000 Succ+ colonies grew in the bacterial lawn after 3 days of incubation (Fig 2D).

Collectively, these results indicated that Salmonella growth upon succinate was consistently inhibited in our experimental setup. The eventual initiation of exponential phase did not result from an orchestrated metabolic switch, but reflected the emergence of spontaneous mutants that efficiently utilised succinate, and proliferated to outcompete the WT bacteria. This Succ+ phenotype remained stable after two passages on LB medium, indicating that the trait was not a phase-variable phenomenon caused by epigenetic mechanisms, as has been observed for other reversible phenotypes in Salmonella [41,42]. Moreover, our data suggest that other pathogenic bacteria, such as Klebsiella also suppress succinate assimilation (Fig 1).

Here, we define “succinate utilisation” as the ability of Salmonella to grow with succinate as a sole carbon and energy source. We selected S. Typhimurium strain 4/74 for further study of the suppression of succinate utilisation because it is the prototrophic parent of strain SL1344, which has been used for a plethora of regulatory and infection studies in the past [43]. We aimed to identify novel genetic determinants involved in the control of the uptake and catabolism of this infection-relevant C-source.

Identification of novel mutations that abolish inhibition of succinate utilisation under aerobic conditions

To identify mutations that ablate the inhibition of succinate utilisation, we used three complementary unbiased approaches. We first screened a collection of published S. Typhimurium 4/74 mutants [44] that lacked key regulatory proteins (S1A Fig), and focused on mutations that both promoted growth on succinate agar plates and in liquid medium with aeration. This screen revealed that mutants lacking the RNA chaperone Hfq (Δhfq) and the polynucleotide phosphorylase (PNPase, mutant Δpnp) had a Succ+ phenotype. Complementation with low-copy plasmids carrying hfq+ or pnp+, restored the Succ- WT phenotype in the corresponding mutant (S3C and S3D Fig).

To explore metabolic suppression in more depth, we used global Tn5 transposon mutagenesis to generate insertions that promoted growth on succinate (Methods). RpoS has a key role in the inhibition of succinate utilisation [2931] and rpoS inactivation did cause the drastic shortening of the lag time of 4/74 in M9+Succ, similarly to the succinate evolved cultures (Figs 2B, S1B, and S3B). Therefore, we developed a strategy to avoid the selection of Succ+ rpoS mutants by constructing a strain that carried two chromosomal copies of rpoS (4/74 rpoS2X; S1B and S1C Fig). Following Tn5 mutagenesis of the rpoS2X strain, individual Succ+ Tn5 mutants were isolated. The Tn5 insertions were P22-transduced into 4/74 WT and the Succ+ phenotype of the transductants was confirmed (Table 1).

thumbnail
Table 1. Fifteen mutations that stimulated growth of S. Typhimurium upon succinate.

https://doi.org/10.1371/journal.pgen.1011142.t001

In parallel, we isolated spontaneous Succ+ mutants from either M9+Succ agar plates (Fig 2D) or from liquid cultures (Methods). We first verified the RpoS positive (rpoS+) status of each spontaneous Succ+ mutant (Methods), and then used whole genome-sequencing to identify relevant nucleotide changes, which were associated with seven genes (Table 1).

These complementary genetic screens identified Succ+ mutants that carried Tn5 insertions in iraP, cspC, rbsR and fliD or in the 5’ untranslated region (5’-UTR) of the yobJ-cspC operon (Table 1). In addition, a nonsense spontaneous mutation in iscR (yfhP) and a spontaneous in-frame insertion of 4 codons in rbsR were identified in spontaneous Succ+ mutants (Table 1). To independently confirm the function of these genes, λ red recombination was used to generate ΔiraP, ΔcspC, ΔrbsR and ΔiscR deletion mutants. Each of the four mutants had the Succ+ phenotype (Fig 3). We confirmed that the corresponding WT proteins could inhibit succinate utilisation by plasmid-borne complementation experiments (S3E–S3H Fig).

thumbnail
Fig 3. Growth phenotype of 5 genome-edited Succ+ mutants and 8 regulatory mutants on solid succinate minimal medium.

S. Typhimurium strains 4/74 WT, ΔrpoS (JH3674), Δhfq (JH3584), ΔiraP (SNW188), ΔfliST (SNW288), ΔcspC (SNW292), ΔiscR (SNW184), ΔrbsR (SNW294), Δpnp (JH3649), dctAmut1 (SNW160), dctAmut2 (SNW315), oxyRmut (SNW318), rrsAmut (SNW336), rrsHmut (SNW314), were spread on M9+Succ agar plates and incubated for 48 hours at 37°C. Experiments were carried out with biological triplicates and a representative picture is shown for each strain.

https://doi.org/10.1371/journal.pgen.1011142.g003

We found that inactivation of fliD alone did not cause a Succ+ phenotype. The fliD gene is co-transcribed with the downstream fliS and fliT genes [45]. Tn5 insertions may have polar effects on the expression of surrounding genes [46], raising the possibility that the fliD::Tn5 insertions (S2A Fig) modulated expression of fliS or fliT. Our genetic dissection of the fliDST operon revealed that inactivation of either fliS or fliT caused the Succ+ phenotype, suggesting that the two genes contribute to the inhibition of succinate utilisation (S2B Fig). The Succ+ phenotype of the ΔfliST mutant was confirmed, and complementation of the double mutation restored the WT Succ- phenotype (S3I Fig).

During the complementation experiments, we observed that the presence of chloramphenicol (Cm) mildly stimulated the growth of the WT strain (S3J Fig), an observation that we have investigated further (S2 Text).

Certain spontaneous mutations that stimulated growth with succinate did not reflect a typical loss-of-function scenario. For example, one of the spontaneous Succ+ mutants had an additional CTA codon resulting in an extra leucine between residues 239 and 240 of the transcriptional regulator OxyR (mutant oxyRmut) (Table 1). We also identified two Succ+ mutants with a single nucleotide polymorphism (SNP) located in the 5’-UTR of dctA (mutants dctAmut1 and dctAmut2, Table 1), that encodes for the aerobic succinate transporter DctA [47]. Finally, two Succ+ mutants carried a SNP in the anti-Shine-Dalgarno sequence of the rrsA and rrsH genes that encode two 16S ribosomal RNAs; mutants rrsAmut and rrsHmut (Table 1).

The function of the mutations associated with genes oxyR, dctA, rrsA and rrsH was confirmed by scarless genomic editing to generate exactly the same nucleotide changes in the WT background (Methods). All these reconstructed mutations caused the Succ+ phenotype (Fig 3).

In summary, we identified novel mutations that promote Salmonella growth upon succinate and shortened the lag time in comparison with the WT (S3 and S4 Figs). These included mutations that involved the succinate transporter DctA, the transcriptional regulators (RbsR, IscR, OxyR), RNA binding proteins (PNPase, Hfq and CspC), flagellar protein chaperones (FliS and FliT) and in ribosomal RNAs (RrsA and RrsH) (Fig 3). The eleven novel Succ+ mutations also promoted Salmonella growth upon fumarate or malate, suggesting that the regulatory systems play a general role in the de-inhibition of C4-dicarboxylate utilisation (S1D Fig).

The OxyS sRNA stimulates growth upon succinate by repressing expression of CspC

The spontaneous Succ+ mutants included an oxyR variant (oxyRmut) that encoded an extra leucine residue in the C-terminus domain of the OxyR transcriptional regulator (Table 1 and Fig 4A). OxyR senses oxidative stress and is activated by disulfide bond formation in the presence of reactive oxygen species [48,49]. In E. coli, the OxyR regulon includes about 40 genes, mainly associated with oxidative stress resistance [50,51]. In addition, the oxidized form of OxyR triggers the transcription of OxyS, an Hfq-binding small regulatory RNA (sRNA) [5254]. Previous studies reported the isolation of constitutively-active OxyR variants that carried mutations in the same region as the extra leucine residue of the oxyRmut variant [5557].

thumbnail
Fig 4. The Succ+ oxyRmut mutant expresses OxyS constitutively and stimulates Salmonella growth with succinate in an Hfq-dependent manner.

(A) Schematic representation of the oxyRS locus. The Succ+ mutation oxyRmut has an additional CTA codon encoding for an extra leucine in oxyR. (B) OxyS is constitutively expressed in the oxyRmut strain. Northern blot detection revealed OxyS expression in strains 4/74 WT, oxyRmut (SNW318) and ΔoxyR (SNW320) grown in LB and exposed or not to 2 mM H2O2 for 30 min (Methods). The detection of the 5S rRNA was used as a loading control. (C) OxyS stimulates Salmonella growth with succinate. Growth curves showed the fast growth of the oxyRmut strains in comparison with the WT and the oxyRmut ΔoxyS (SNW340) strains. Growth is restored in the complemented strain oxyRmut ΔoxyS oxySchr+ (SNW362), carrying a chromosomal oxyS copy. (D) The plasmid-borne expression of OxyS stimulates the growth of the WT but not of the ΔoxyR mutant. Growth was assessed for strains 4/74 WT and ΔoxyR (SNW320), carrying either the empty (vector, pPL) or the oxyS expressing (oxyS++, pNAW255) plasmids, schematised at the top of the figure. The bent arrows represent the constitutive promoter PL lacO-1 of the ApR pPL plasmid, carrying the oriColE1 replicon. (E) Hfq inactivation suppresses the fast growth of the oxyRmut mutant. Growth was assessed for strains 4/74 WT, Δhfq (JH3584), oxyRmut (SNW318) and oxyRmut Δhfq (SNW663). The medium used is indicated for each experiment. Growth curves were performed in 96-well plates, as specified in Methods. The details about the construction of strains SNW320, SNW340 and SNW362 are depicted in the S7 Fig.

https://doi.org/10.1371/journal.pgen.1011142.g004

Consequently, we investigated whether the oxyRmut allele drove constitutive transcription of the OxyS sRNA. Northern blot analysis revealed that in the oxyRmut strain, OxyS was strongly expressed both in the absence and in the presence of hydrogen peroxide, indicating that the OxyRmut protein is constitutively active (Fig 4B). We then determined whether constitutive expression of OxyS was responsible for the Succ+ phenotype of the oxyRmut mutant. The deletion of oxyS in the oxyRmut strain (oxyRmut ΔoxyS) totally abolished the Succ+ phenotype (Fig 4C). A complementation experiment was carried out by re-introducing a single copy of oxyS and its native promoter (oxySchr+, S7 Fig) into the chromosome of the oxyRmut ΔoxyS strain (Methods). This chromosomal complementation restored the fast growth of the oxyRmut ΔoxyS mutant in M9+Succ (Fig 4C). Furthermore, the plasmid-borne expression of OxyS boosted the growth of 4/74 WT in M9+Succ, confirming that high level expression of the OxyS sRNA stimulated growth with succinate (Fig 4D). The same plasmid did not stimulate the growth of the ΔoxyR strain indicating that other OxyR-dependent genes are required to grow under this condition.

We previously showed that Hfq inactivation boosted succinate utilisation, but in the oxyRmut genetic background the same Hfq inactivation dramatically reduced growth and extended the duration of lag time in M9+Succ (Fig 4E). Collectively, our findings show that the OxyS sRNA orchestrates the de-inhibition of succinate utilisation in concert with Hfq.

In E. coli, OxyS acts as an indirect repressor of RpoS expression, likely via the titration of Hfq [53]. OxyS also represses the expression of the yobF-cspC operon, probably by base-pairing near the SD motif of the yobF 5’-UTR [58,59]. Because RpoS, Hfq and CspC repress succinate utilisation (Fig 3), we tested the effects of the plasmid-borne overexpression of rpoS+, hfq+ or cspC+ on the growth of the oxyRmut strain. The overexpression of Hfq and RpoS slightly increased the lag time of oxyRmut strain, while the plasmid-borne expression of CspC totally abolished the Succ+ phenotype in this genetic background (Fig 5A).

thumbnail
Fig 5. The OxyS sRNA stimulates Salmonella growth with succinate by repressing expression of CspC.

(A) Overexpression of cspC inhibits the growth of the oxyRmut strain (SNW318), while hfq and rpoS overexpression have only a mild effect. The oxyRmut strain was carrying either the empty (vector, pNAW125), the prpoS (pNAW95), the phfq (pNAW45) or the pcspC (pNAW92) plasmids, schematised at the top of the figure. These low-copy CmR plasmids are carrying the ori101* replicon and the genes of interest are under the control of the strong constitutive promoter PL tetO-1 (bent arrows). As a control, the growth of 4/74 WT carrying the empty plasmid (vector) was also assessed. (B) Strategy to test whether the Salmonella sRNA OxyS represses the expression of the yobF-cspC mRNA at the post-transcriptional level, as previously reported in E. coli [58]. The plasmid-borne translational fusion yobF::sfgfp (pNAW258) is depicted and was constructed as described in Methods and in Corcoran et al., 2012 [99]. This fusion is under the control of the constitutive promoter PL tetO-1.SD = Shine-Dalgarno. (C) The yobF::sfgfp activity is reduced in the oxyRmut strain, that expresses constitutively OxyS: fluorescence GFP intensities were measured in strains 4/74 WT and oxyRmut carrying yobF::sfgfp. (D) Prediction of the RNA secondary structure of Salmonella OxyS and of yobF 5’-UTR, using Mfold [119]. The putative kissing complex between the two RNA molecules was predicted with IntaRNA [120] and the corresponding nucleotides are indicated in magenta for yobF 5’-UTR or in blue for OxyS. The mutation oxySGG is indicated and the corresponding nucleotides are framed. (E) The plasmid-borne overexpression of OxyS represses the yobF::sfgfp activity and the oxySGG mutation attenuates this repression: fluorescence GFP intensities were measured in the ΔoxyS strain (SNW338) carrying yobF::sfgfp and the empty (empty vector, pPL), the pPL-oxyS (oxySWT++, pNAW255) or the pPL-oxySGG (oxySGG++, pNAW259) plasmids. (F) The oxySGG mutation reduces the growth of the oxyRmut strain: strains 4/74 WT, oxyRmut and the oxyRmut mutant carrying the oxySGG mutation (SNW670) were grown in M9+Succ. (G) Effects of rpoS overexpression on the growth of the ΔcspC mutant: the strains 4/74 WT and ΔcspC (SNW292), carrying either the empty plasmid (vector, pNAW125) or the prpoS (pNAW95) plasmid were grown in M9+Succ. For A, F and G, growth curves were carried out in the indicated medium with 6 replicates grown in 96-well plates. For C & E, strains were grown to OD600 ~ 2 in LB, supplemented with the appropriate antibiotic(s). GFP fluorescence intensities were measured, as specified in Methods. The graphs represent the relative fluorescence intensities (%), in comparison with the indicated reference strain (100% of intensity). The same strains carrying GFP fusions were grown on LB agar plates and pictures were taken under blue light exposure. The data are presented as the average of biological triplicates ± standard deviation and the statistical significance is indicated, as specified in Methods.

https://doi.org/10.1371/journal.pgen.1011142.g005

To confirm that the OxyS-driven repression of the yobF-cspC was conserved in Salmonella we used a plasmid-borne yobF::sfgfp translational reporter (Fig 5B). In comparison with the WT, the yobF::sfgfp activity was significantly lower in the oxyRmut strain (~2-fold repression), confirming that OxyS represses the expression of the yobF-cspC operon in Salmonella (Fig 5C). Bioinformatic analyses identified putative secondary structures and a potential base-pairing interaction between OxyS and the yobF 5’-UTR (Fig 5D), which was predicted to be an 12 nucleotide-long kissing complex between OxyS and the yobF 5’-UTR, consistent with the proposed interaction in E. coli [59] (Fig 5D).

To assess the role of the kissing complex experimentally, we generated a mutated version of OxyS with a CC➔GG mutation in the loop of the first RNA hairpin (allele oxySGG, Fig 5D). This mutation was introduced into the chromosome of the oxyRmut strain (strain oxyRmut oxySGG) and the oxySGG gene was cloned into the pPL expression vector [60]. To test the impact of sRNAs OxyS and OxySGG upon yopF-cspC operon expression, the empty pPL vector, the pPL-oxyS or the pPL-OxySGG plasmids were transferred into the oxyS null strain carrying the yobF::sfgfp fusion and the GFP signal was measured. In this genetic background, oxyS expression (oxyS++) reduced the GFP fluorescence intensity by ~3-fold, but only by ~1.5-fold in the strains expressing oxySGG (oxySGG++) (Fig 5E). Consistent with the attenuated repression of yobF in the presence of OxySGG, the oxyRmut oxySGG strain had a longer lag time than the oxyRmut mutant, confirming that the mutated region of OxyS is involved in the de-inhibition of succinate utilisation (Fig 5F).

In E. coli, CspC stabilises rpoS mRNA and increases the cellular level of RpoS [61,62]. To investigate whether the Succ+ phenotype of the CspC null mutant was caused by changes in RpoS expression, we tested the effect of RpoS overexpression in the ΔcspC mutant (Fig 5G). The plasmid-encoded overexpression of RpoS only marginally extended lag time in the ΔcspC mutant, indicating that repression of succinate utilisation by CspC is RpoS-independent. A recent study corroborated this observation, as the CspC-mediated activation of RpoS did not occur in Salmonella [63].

Collectively, our results indicate that the OxyS sRNA is a key determinant in the de-inhibition of succinate utilisation by Salmonella. Despite the fact that OxyS can regulate RpoS expression levels, we propose that OxyS stimulates the Succ+ phenotype by repressing the expression of CspC via base-pairing in the vicinity of the yobF SD motif. CspC is an RNA binding protein, belonging to the cold shock protein family [64]. CspC and its paralog CspE often have redundant functions, being involved in biofilm formation, motility, stress resistance and virulence modulation in S. Typhimurium [61,63]. It remains unclear how the OxyS-driven inhibition of CspC expression impacts upon the catabolism of succinate. One possibility is that CspC directly represses succinate catabolic genes. In line with this hypothesis, a transcriptomic study in S. Typhimurium, revealed that the sdhC, sdhD and sdhA genes are moderately up-regulated in a ΔcspEC mutant [63].

Our study also revealed that two other RNA binding proteins, Hfq and PNPase in concert with their cognate sRNAs also contribute to the repression of Salmonella growth upon succinate (S1 Text and S5 Fig). We discovered that Hfq played a dual role on the regulation of succinate utilisation by stimulating succinate utilisation in concert with OxyS, whilst maintaining succinate growth inhibition, probably via the stimulation of rpoS expression [65].

In summary, these results demonstrate the crucial role of RNA-binding proteins and their associated sRNAs in the fine-tuning of Salmonella carbon metabolism.

The iron-sulphur cluster regulator IscR inhibits growth upon succinate by repressing DctA expression

In E. coli, the C4-dicarboxylate transporter DctA mediates succinate uptake under aerobic conditions [47]. DctA is also a C4-dicarboxylate co-sensor that modulates the expression of several genes, including dctA itself, in concert with the two-component system DcuR/S [66]. The transcription of dctA is controlled by catabolic repression and putative CRP binding sites have been identified in the dctA promoter region [67,68] and are conserved in Salmonella (Fig 6A).

thumbnail
Fig 6. IscR represses the expression of DctA and inhibits Salmonella growth upon succinate.

(A) Detailed schematic representation of the dctA promoter (PdctA) region in S. Typhimurium (S. Tm) 4/74 and in E. coli MG1655. Conserved nucleotides in 4/74 and MG1655 are highlighted in light gray. The promoter -35 and -10 boxes and the transcription start site (TSS, numbering +1) are indicated according to Davies et al., 1999 [67]. The DcuR binding site [121] and the putative CRP binding site [67,68] are indicated. A putative IscR binding site is depicted and the mutations identified in the Succ+ mutants dctAmut1 and dctAmut2 are indicated in red. In addition, the corresponding region of E. coli B strain REL606 is depicted: in this strain, a G➔A SNP (in red) causing dctA up-regulation and the stimulation of succinate utilisation was previously described [75]. SD = Shine-Dalgarno motif. Promoters (P) are represented by bent arrows. The symbol “*” denotes that the binding sites were not experimentally demonstrated. (B) The stimulation of DctA expression and the dctAmut1 and dctAmut2 mutations boost Salmonella growth with succinate. The AHT-inducible strain tetR-PtetA-dctA (SNW133) is depicted: the tetR repressor gene, the PtetR and the PtetA promoters are indicated. The residual FLP recognition target site sequence is denoted by”frt”. In the absence of AHT no growth was detected for both WT and tetR-PtetA-dctA strains, while AHT addition stimulated the growth of the tetR-PtetA-dctA strain. Similarly, the dctAmut1 (SNW160) and dctAmut2 (SNW315) strains displayed a fast growth in M9+Succ. (C & D) The SNP mutations dctAmut1 and dctAmut2 and the inactivation of IscR stimulate the expression of DctA. The chromosomal transcriptional/translation dctA::sfgfp fusion is depicted: sfgfp encodes for the superfolder GFP fused in frame to DctA C-term. The GFP fluorescence intensity was measured in strain WT dctA::sfgfp (SNW296), dctAmut1 dctA::sfgfp (SNW310), dctAmut2 dctA::sfgfp (SNW316), ΔiscR dctA::sfgfp (SNW329) and ΔrpoS dctA::sfgfp (SNW313).(E) Both Apo- and holo- forms of IscR repress Salmonella growth with succinate. The growth was assessed for strains 4/74 WT and ΔiscR (SNW184) carrying either the empty plasmid (vector, pXG10-SF), the piscR (pNAW96) or the piscR3CA (pNAW97) plasmids. The piscR3CA expressed the IscR3CA variant that prevents the binding of an iron-sulphur cluster, maintaining IscR in its apo-form (see text for details). Growth curves (B & E) were carried out in the indicated medium with 6 replicates grown in 96-well plates. For C & D, strains carrying the dctA::sfgfp fusion were grown in M9+Gly+Succ to OD600 0.5–1 and GFP fluorescence intensity was measured by fluorescence microscopy (C) or by flow cytometry (D), as specified in Methods. The graph (D) represents the relative fluorescence intensities of each strain (%), in comparison with the WT strain carrying dctA::sfgfp (SNW296, 100% of intensity). The data are presented as the average of biological triplicates ± standard deviation and the statistical significance is indicated, as specified in Methods. NS, not significant.

https://doi.org/10.1371/journal.pgen.1011142.g006

To determine whether DctA was required for Salmonella growth under our experimental conditions, we constructed a chromosomal inducible dctA construct, by replacing the dctA promoter with tetR (encoding the TetR repressor) and the tetA promoter (strain tetR-PtetA-dctA, Fig 6B). In the absence of anhydrotetracycline (AHT) inducer, the tetR-PtetA-dctA strain did not grow at all in M9+Succ. However, upon addition of AHT, the tetR-PtetA-dctA strain displayed a fast growth phenotype that was not observed with the WT (Fig 6B). A long incubation (>60 hours) confirmed that dctA was essential for Salmonella growth and the emergence of the Succ+ mutants, ruling out a role for other C4-dicarboxylate transporters under the conditions tested (S8 Fig). We conclude that Salmonella requires the DctA transporter to grow on succinate and that dctA expression is likely to be repressed in the WT, as previously hypothesised by Hersch and co-workers [16].

The two spontaneous Succ+ mutants dctAmut1 and dctAmut2 (Table 1) carry SNPs in the 5’-UTR of dctA (Fig 6A), and both promoted growth with succinate (Fig 6B). We reasoned that the Succ+ mutations could de-repress dctA expression, which we examined with a chromosomal dctA::sfgfp transcriptional/translational reporter fusion in the Succ+ mutant backgrounds (Methods and Fig 6C). To allow homogenous growth for all the strains (including the Succ- 4/74 WT), bacteria were grown in M9 supplemented with glycerol (40 mM) as the main C-source and the addition of 10 mM succinate, to stimulate expression of succinate-induced genes [69]. The GFP fluorescence intensity of single bacteria was measured for each strain by flow cytometry. In comparison with the low levels of GFP fluorescence seen in the 4/74 WT background, higher GFP levels were only observed in the presence of the dctAmut1, dctAmut2 and ΔiscR mutations (S9A–S9N Fig). The regulation was confirmed by fluorescence microscopy (Fig 6C) and flow cytometry, using biological triplicates (Fig 6D).

A recent report proposed that RpoS indirectly represses dctA in Salmonella [16], but such an increase of the dctA::sfgfp fusion activity was not observed in the ΔrpoS strain under our experimental conditions (Figs 6D and S9P). We found that the dctA::sfgfp-tagged ΔrpoS, dctAmut1, dctAmut2 and ΔiscR mutants grew much faster in M9+Succ than the isogenic WT strain (S9O Fig), with the same growth rate we observed previously with the corresponding untagged mutants (Figs 6B,6E, and S1B). This indicated that the C-terminal addition of sfGFP did not impede the function of DctA as a succinate transporter or co-sensor.

The up-regulation of dctA in the Salmonella IscR null mutant was consistent with the IscR-driven repression of dctA proposed in E. coli [70]. In most Gram-negative bacteria, the dual regulator IscR controls the transcription of the iron-sulphur (Fe-S) cluster biosynthesis operon iscRSUA and the Fe-S cluster assembly genes sufABCDSE [71,72]. The apoprotein form of IscR (apo-IscR) is matured by the Isc system into its [Fe2-S2]-containing holo-form. The resulting holo-IscR represses the expression of several genes including the isc operon. Under iron-limitation and in the presence of reactive oxygen species, the IscR apo-form predominates and stimulates the expression of the suf operon, in concert with OxyR [71,73]. IscR binds to two classes of DNA motifs: Type 1 motifs are only bound by holo-IscR, while Type 2 motifs are recognised by both holo- and apo- forms [70,74].

Analysis of the promoter region of the Salmonella dctA revealed the presence of a putative Type 2 IscR-binding site (ATAACCTTACAAGACCTGTGGTTTTT) [74] located 10 bp downstream of the transcription start site of dctA (Fig 6A). Both the Succ+ mutants dctAmut1, dctAmut2 carried SNPs within this DNA motif. This motif is also conserved in E. coli MG1655 (Fig 6A) and a similar SNP, that stimulated dctA transcription and succinate utilisation, has previously been identified in E. coli B [75].

To investigate which of the apo/holo-forms of IscR was repressing succinate utilisation in Salmonella, we constructed a plasmid expressing an IscR variant carrying three Cys➔Ala substitutions (IscR3CA,Cys92,98,104➔Ala92,98,104) that prevent the binding of [Fe2-S2] to IscR, and maintain the apo-form of the protein [76,77]. The plasmid-borne expression of both IscR and IscR3CA complemented the ΔiscR deletion and suppressed the Succ+ phenotype (Fig 6E), indicating that both apo- and holo- IscR repress succinate utilisation.

Collectively, these results demonstrated that IscR plays a critical role in the repression of dctA and in the inhibition of succinate utilisation. The apo-IscR represses succinate utilisation, suggesting that IscR represses dctA expression by binding the putative Type 2 DNA motif identified downstream of the dctA promoter. The finding of dctA-stimulating SNPs in this motif supports our hypothesis, and it is possible that the binding of IscR downstream of the promoter acts as a “roadblock” that inhibits dctA transcription, as proposed for the repression of mgtC by PhoP [78]. However, the repression of dctA by IscR could also be indirect with the dctAmut1 and dctAmut2 mutations stimulating dctA transcription by another mechanism. Further study is required to understand the IscR-driven repression of dctA at the mechanistic level.

Our data show that the de-repression of dctA expression is sufficient to stimulate Salmonella growth with succinate, consistent with a previous report [16]. However, dctA de-repression was only observed in the dctAmut1, dctAmut2 and ΔiscR mutants, raising the question of whether DctA-driven succinate uptake is the key limiting factor for rapid growth of Salmonella in M9+Succ media. In E. coli, the presence of fumarate stimulates the expression of sdhC and other TCA cycle genes in a DcuR-dependent manner [18]. Because DctA can act as a co-sensor of the DcuS/R transduction cascade [66], it is possible that the increase of DctA abundance impacts upon the expression of the sdh operon, stimulating growth upon succinate.

Succinate utilisation is inhibited by RbsR and FliST via RpoS

Factors that modulate RpoS expression, stability or activity are likely to control succinate utilisation in Salmonella. For example, the inactivation of the anti-adapter IraP stimulates succinate utilisation by increasing RssB-facilitated proteolysis of RpoS by the ClpXP protease [16,79,80]. Our genetic screens showed that succinate utilisation was stimulated by the absence of IraP, CspC, RbsR and FliST and by increased expression of OxyS.

To assess RpoS levels in the corresponding Succ+ mutants, we used Western blot detection (S10A Fig). The levels of RpoS protein in the ΔcspC, ΔfliST, oxyRmut mutants were similar to the WT strain, while lower levels of RpoS were observed in the ΔiraP and ΔrbsR mutants. We confirmed the role of RbsR in RpoS activation by complementing the ΔrbsR mutation with the prbsR plasmid, revealing that inactivation of RbsR reduced RpoS abundance in exponential and early stationary phases, but not in stationary phase (Figs 7A and S10B). During the characterisation of the Succ+ mutants, we noticed that the RbsR null strain was impaired in its capacity to form red, dry and rough colonies (RDAR), another RpoS-dependent phenotype of Salmonella [81]. The RDAR morphotype was restored in the ΔrbsR mutant by complementation with the plasmid prbsR (Fig 7B). We observed that the plasmid-borne overexpression of RpoS in the RbsR null mutant totally abolished the Succ+ phenotype, indicating that RbsR represses succinate utilisation via the activation of RpoS (Fig 7C).

thumbnail
Fig 7. RbsR and FliST inhibit Salmonella growth upon succinate via RpoS.

(A) RbsR inactivation reduces the cellular level of RpoS in exponential and early stationary phases. Strains 4/74 WT and ΔrbsR (SNW294) carrying the empty (vector, pXG10-SF) or the prbsR (pNAW93) plasmids were grown in LB (without Cm) to late exponential (OD600~1), early stationary (OD600~2) and stationary phase (OD600~4). The cellular levels of RpoS and DnaK (loading control) were assessed by Western blotting (Methods). As negative control, the ΔrpoS mutant (JH3674) carrying pXG10-SF was included. The experiment presented is representative of three independent experiments and two replicates are presented in S10B Fig. (B) RbsR inactivation reduces red, dry and rough (RDAR) morphotype, another RpoS-dependent phenotype. The RDAR phenotypic assays were carried out as specified in Methods with strain 4/74 WT, ΔrbsR and ΔrpoS carrying the indicated plasmids (vector = pXG10-SF). At least three independent experiments were performed, and representative RDAR colony pictures are presented. The plasmid-borne rpoS overexpression suppresses the Succ+ phenotype of the ΔrbsR (C) and ΔfliST (D) mutants. Growth was assessed in the indicated medium in a 96-well plate for strains ΔrbsR (SNW294) or ΔfliST (SNW288) carrying either the empty plasmid (vector, pNAW125) or the prpoS plasmid (pNAW95) and strains 4/74 WT carrying the empty plasmid.

https://doi.org/10.1371/journal.pgen.1011142.g007

RbsR is a LacI-type transcriptional regulator that inhibits the transcription of the ribose utilisation operon (rbsDACBKR) in the absence of ribose [82]. To investigate whether RbsR stimulated rpoS transcription, we used a chromosomal transcriptional GFP reporter fusion, with a gfp+ gene (including its SD sequence) inserted downstream of the main transcription start site of the rpoS locus (S10C Fig). Similar GFP levels were observed in the WT and the ΔrbsR strains grown to either exponential, early stationary or stationary phase (S10D Fig).

Taken together, our findings show that RbsR acts as a pleiotropic regulator in Salmonella, controlling growth with succinate and the RDAR morphotype, via the positive regulation of rpoS. RbsR does not directly stimulate rpoS promoter activity and we propose an indirect RbsR-driven activation of RpoS at the post-transcriptional or the post-translational level. In line with this hypothesis, it was recently observed that RpoS is repressed at the post-transcriptional level, when the RbsR-dependent rbsD gene is over-expressed in E. coli [83].

Our work also revealed that the two flagellar chaperones FliS [84] and FliT [85] control succinate utilisation (Figs 3 and S2B), suggesting a link between the control of the flagellar machinery and Salmonella central carbon metabolism. In the ΔfliST mutant, RpoS overexpression totally suppressed the Succ+ phenotype (Fig 7D) indicating that the regulation is RpoS-dependent. However, Western blots did not show reduced levels of RpoS in the FliST null mutant (S10A Fig). It remains unclear how these protein chaperones inhibit succinate utilisation, and whether FliS and FliT are capable of stimulating the expression or the activity of RpoS.

Perspective

Here, we have revealed that in addition to the RpoS-driven inhibition of succinate utilisation CspC and IscR systems control the utilisation of this infection-relevant C-source. The fact that mutations in the anti-Shine-Dalgarno sequence of two 16S ribosomal RNAs or the presence of ribosome-inhibiting antibiotic chloramphenicol boosted growth with succinate added extra complexity (S2 Text and S11 and S12 Figs). More work will be required to fully unravel the genetic mechanisms by which these redundant regulatory systems impact upon succinate catabolism and/or succinate uptake genes.

These distinct regulatory mechanisms are likely to adjust Salmonella metabolism during the journey of the pathogen through the host; from the colonisation of the gastrointestinal tract to intra-macrophage replication. We showed that the sRNA OxyS antagonises CspC-dependent inhibition. Because OxyS is induced by oxidative stress, our findings raise the possibility that the reactive oxygen species produced in the inflamed gut [86,87] can stimulate Salmonella growth upon microbiota-derived succinate in this niche [3,5]. Despite, the abundance of succinate within infected macrophages [12], the intracellular proliferation of Salmonella does not require succinate catabolic genes [15]. The high levels of intra-macrophage expression of the iscR [88] and rpoS [88,89] lead us to propose that the DctA-driven uptake and catabolism of succinate are repressed in the intra-macrophage niche.

The crucial open question that remains is why Salmonella and other enterobacteriaceae display an aversion to utilising succinate for growth. Because succinate is an important infection-relevant C-source, we hypothesise that the distinct regulatory mechanisms we have identified act in response to various environmental stimuli to fine-tune Salmonella succinate metabolism during the journey of the pathogen through the host. We propose a model that explains how the overlapping regulatory systems allow Salmonella to utilise succinate “at the right place and at the right time” during infection (Fig 8).

thumbnail
Fig 8. A model depicting the modulation of Salmonella succinate utilisation by multiple environmental stimuli under aerobic conditions.

(A) RpoS expression stimulation by Hfq, PNPase, sRNAs and RbsR inhibit Salmonella succinate utilisation. The sRNAs ArcZ, DsrA and RprA stimulate rpoS mRNA transcription elongation [122] and translation [123] in concert with Hfq. RpoS represses the transcription of several genes of the TCA cycle, including the sdh operon [24,25], inhibiting succinate catabolism and Salmonella growth with this C-source. PNPase presumably represses succinate utilisation indirectly by its role in the stabilization of several Hfq-associated sRNAs [124] and in the translational activation of rpoS [125]. The ribose sensor RbsR stimulates the expression of RpoS, presumably at the post-transcriptional level (Fig 7), repressing growth upon succinate. (B) The sRNAs RyhB-1, Spf and RybB repress sdh mRNA translation [126,127] in concert with Hfq and attenuate succinate dehydrogenase (SDH) synthesis, inhibiting Salmonella growth upon succinate. In addition, PNPase may promote the degradation of the sdhCD mRNA [128]. (C) Under aerobic conditions, succinate is mainly imported by DctA [47]. DctA interacts with the DcuS protein and acts as a co-sensor of C4-dicarboxylates [66,129]. In the presences of succinate, DcuS/DctA activates the response regulator DcuR, that stimulates the expression of several genes, including dctA [121]. De novo synthetised DctA accumulates and increases the uptake of succinate and DctA accumulation may also stimulate the transcription of succinate utilisation genes in concert with DcuS/R [18]. However, dctA expression is robustly repressed by the both halo- and apo-forms of the iron-sulphur cluster regulator IscR (Fig 6), which is up-regulated under iron limiting conditions [111]. Therefore, IscR plays a pivotal role in the succinate utilisation repression and blocks Salmonella growth with this C-source (D) OxyR is stimulated by oxidative stress and stimulates the expression of the sRNA OxyS. OxyS stimulates Salmonella succinate utilisation by repressing the small RNA binding protein CspC. CspC represses succinate utilisation by a still unknown mechanism.

https://doi.org/10.1371/journal.pgen.1011142.g008

Materials and methods

Bacterial strains and growth conditions

Precise details of all the chemicals, reagents, DNA oligonucleotides (primers), plasmids and bacterial strains used in this study are listed in S1 Table. The Salmonella mutant strains were all derivatives of Salmonella enterica serovar Typhimurium strain 4/74 [90]. Strain 4/74 is now available from the UK National Collection of Type Cultures (https://www.culturecollections.org.uk/products/bacteria/index.jsp) as NCTC 14672. All the nucleotide coordinates given for 4/74-derived strains correspond to the published genome: GenBank CP002487.1 [37]. Escherichia coli strains Top10 (Invitrogen) and S17-1 λpir [91] were used as hosts for the cloning procedures.

Unless otherwise specified, bacteria were grown at 37°C with aeration (orbital shaking 220 rpm) in Lennox Broth (LB: 10 g/L BD Tryptone, 5 g/L BD Yeast Extract, 5 g/L NaCl), LBO (10 g/L BD Tryptone, 5 g/L BD Yeast Extract) or in M9 minimal medium [92], prepared with M9 Salts, 2 mM MgSO4, 0.1 mM CaCl2 and 40 mM sodium succinate dibasic hexahydrate (succinate) or 40 mM glycerol + 10 mM Succinate, as sole C-sources (henceforth, media M9+Succ and M9+Gly+Succ). Agar plates were prepared with the same media, solidified with 1.5% BD Bacto agar. When required, 10–100 μM of FeCl3 were added to the M9 media.

To seed the M9-derived media of all the experiments, stationary phase pre-cultures were prepared by inoculating isolated colonies into 5 mL LB (in 30 mL Universal glass tubes) and the cultures were incubated for 6–20 hours at 37°C with aeration. Bacteria were harvested by centrifugation, washed once, and Optical Density at 600 nm (OD600) was adjusted to 1 with the minimal medium used for the cultures, or with 1 X Phosphate-Buffered Saline (PBS) to generate a standardised inoculum. Subsequently, bacteria were grown aerobically in conical flasks (topped with aluminium foil) or in Greiner 50 mL plastic tubes (with lids slightly open to allow gaseous exchange). Washed bacteria from the pre-cultures were inoculated as a 1:100 dilution to give a starting OD600 of 0.01 (~107 CFU/mL), in a final medium volume corresponding to 10% of the flask/tube capacity, to ensure optimal oxygenation by shaking.

For growth curves in 96-well microplates (Greiner #655180), bacteria grown beforehand for ~6 hours in LB, were washed with the minimal medium used for the cultures, or with PBS, and inoculated to give a starting OD600 of 0.01 in 200 μL of medium per well. The microplates were incubated at 37°C with orbital shaking (500 rpm), in a FLUOstar Omega plate reader (BMG Labtech), and the OD600 was monitored every 15–30 min, using the appropriate growth medium as blank.

When required, antibiotics were added as follows: 50 μg/mL kanamycin monosulfate (Km), 100 μg/mL Ampicillin sodium (Ap), 25 μg/mL tetracycline hydrochloride (Tc), 20 μg/mL gentamicin sulfate (Gm) and 25 μg/mL chloramphenicol (Cm).

For strains carrying the tetR-PtetA module, the PtetA promoter was induced by adding 500 ng/mL of anhydrotetracycline hydrochloride (AHT, from a 1 mg/mL stock solubilised in methanol). The same volume of methanol was added to the mock-induced cultures. To stimulate expression of genes controlled by the PBAD promoter (e.g. in plasmid pWRG99), 0.2% L-(+)-arabinose was added to the culture. For the strains carrying the plasmid pSW-2 [93], the Pm promoter was induced by adding 1 mM m-toluate (500 mM stock titrated with NaOH to pH 8.0).

Bacterial transformation and Tn5 mutagenesis

Chemically-competent E. coli were prepared and transformed as previously published [94]. Electrocompetent cells were prepared with Salmonella cultures grown in salt free LBO medium and were electroporated, as described previously [95]. After recovery in LB at 37°C (30°C for temperature-sensitive plasmids) transformation reactions were spread on selective LB agar plates and transformants were obtained after incubation at 30°C or 37°C.

For the Tn5 transposon mutagenesis, ultra-competent Salmonella were prepared from LBO cultures grown at 45°C, as previously reported [95,96]. The λpir-dependent plasmid pRL27 [97], that encodes the Tn5 transposase gene (tnp) and the mini Tn5-oriR6K-KmR transposon (Tn5), was used to generate the Salmonella Tn5 libraries, as follows: 50 μL of ultra-competent Salmonella were electroporated with 500 ng of the non-replicating Tn5 delivery plasmid pRL27. After 1 hour recovery in LB the transformation reactions containing the Tn5-carrying Salmonella were washed in PBS or minimal media and 1% of transformations was spread on LB Km plates to estimate the size of the resulting Tn5 library. The remainder of the Tn5 libraries were stored for further experiments.

Cloning procedures

Enzymes, buffer and kits used are listed in S1 Table. DNA manipulations were carried out according to standard protocols [92]. DNA fragment were purified from enzymatic reactions or from agarose gel using the Bioline ISOLATE II PCR and Gel Kit. Plasmids were extracted with the Bioline ISOLATE II Plasmid Mini Kit. Genomic DNA (gDNA) was isolated from 0.5–1 mL of stationary phase cultures with the Zymo Quick DNA Universal Kit.

For PCR, DNA was amplified with Phusion High Fidelity DNA polymerase, template DNA and 0.5 μM primers, in the presence of 3% Dimethyl Sulfoxide and 1 M betaine. For plasmid/strain verifications by Sanger sequencing, PCR reactions were carried out from bacterial colonies with MyTaq Red Mix 2 X and PCR fragment were Sanger sequenced with the appropriate primers (Lightrun service, Eurofins Genomics).

DNA digestion/ligation procedures or the restriction-free “PCR cloning” technique [98] were used to insert DNA fragments into the plasmids pEMG [93], pJV300 (pPL) [60] and pXG10-SF [99]. E. coli Top10 was used as host for the construction of pXG10-SF and pPL derived plasmids, while S17-1 λpir was used for pEMG derivatives.

The construction of each plasmid is detailed in S1 Table. For complementation experiments, the genes of interest (including their native ribosome binding site) were PCR-amplified from 4/74 gDNA and were cloned between the NsiI and XbaI sites of the low copy plasmid pXG10-SF, resulting in plasmids phfq (pNAW45), prpoS (pNAW95), piraP (pNAW98), pfliST (pNAW94), prbsR (pNAW93), piscR (pNAW96) and pcspC (pNAW92). For the construction of the piscR3CA plasmid (pNAW97, carrying three Cys➔Ala substitution at positions 92, 98 and 104 in IscR), two fragments carrying the appropriate mutations were first amplified with primer pairs NW_461/NW_467 and NW_462/NW_466 and the resulting amplicons were fused by overlap extension PCR [100] before insertion into pXG10-SF. In all these plasmids, the genes of interest were under the control of the strong constitutive promoter PL tetO-1 [101]. For the construction of ppnp (pNAW256), the pnp gene and its promoter region (including sraG) were amplified and inserted between the XhoI and XbaI sites of pXG10-SF. In the resulting plasmid, pnp expression was controlled by its native promoter. For the construction of poxyS (pPL-oxyS, pNAW255), oxyS was PCR amplified and inserted by PCR cloning downstream of constitutive PL lacO-1 promoter of the pPL vector, as described earlier [102]. The plasmid-borne translational fusion yobF::sfgfp (pyobF::sfgfp, pNAW258) was constructed by cloning the 5’-UTR of yobF and its 31 first codons in frame with the sfgfp gene of pXG10-SF, as previously described [99].

Plasmid site-directed mutagenesis [103] was used to introduce the CC➔GG mutation in the pPL-oxyS: complementary primers NW_1022 and NW_1023, carrying the mutations were annealed and elongated by PCR for 10 cycles. Fifty nanograms of plasmid poxyS (pPL-oxyS, pNAW255) were added to the reaction and the PCR reaction was resumed for 25 cycles. After DpnI treatment and transformation in E. coli Top10, the mutated plasmids pPL-oxySGG (pNAW259) was obtained.

For the construction of the empty vector pNAW125 (pXG10-SF lacking the lacZ186::sfgfp fragment), a 3.5 kb fragment was PCR-amplified from pXG10-SF with the primer pair NW_348/NW_565, the resulting fragment was digested with NsiI and self-ligated.

All the pXG10-SF, pEMG and pPL derived plasmids were verified by Sanger sequencing: primers pXG10_R2 and pZE-CAT were used for plasmid pXG10-SF, primers M13_-40_long and M13_Rev_long for plasmid pEMG and primer pZE-A for plasmid pPL.

Genome editing techniques

The λ red recombination methodology was used to insert or delete genes in the Salmonella chromosome, using the heat-inducible λ red plasmid pSIM5-tet [104]. Salmonella carrying the pSIM5-tet were grown in LBO at 30°C and electrocompetent bacteria were prepared after a 15 minute heat shock at 42°C, as previously described [95,105]. PCR fragments carrying a resistance gene were PCR-amplified from the template plasmids pKD4, pKD3, pNAW52, pNAW55 or pNAW62 [95,106]. Electrocompetent bacteria (50 μL) were transformed with 500–3000 ng of the PCR fragments, and recombinants were selected on LB agar plates that contained the appropriate antibiotics.

The deletions/insertions linked to a selective marker and the Tn5 insertions were transduced to S. Typhimurium strains using the P22 HT 105/1 int-201 (P22 HT) transducing phage, as previously reported [107,108].

When required, antibiotic resistance cassettes flanked by FLP recognition target sites (frt), were removed from the Salmonella chromosome with the FLP recombinase-expressing plasmid pCP20 [109]. Subsequently, the temperature sensitive plasmid pCP20 plasmid was eliminated by a single passage at 42°C.

The construction of each strain is detailed in S1 Table. The marked mutations ΔrpoS::aph, Δhfq::aph, ΔrybB::aph, Δpnp::cat (pnp-539) and ΔarcA::aph were transduced from published 4/74 and SL1344 derivative strains. For the transduction of the Δhfq::aph mutation, the donor strain JH3584 was first complemented with the plasmid phfq (pNAW45), because Hfq is required for P22 transduction [60].

The 4/74 tetR-PtetA-dctA strain was constructed according to the principle described by Schulte and colleagues [110]. The promoter and the 5’-UTR of dctA (coordinates 3812883–3812972) were replaced by a frt-aph-frt-tetR-PtetA module from plasmid pNAW55. In the resulting strain, dctA has the tetA ribosome binding site and is controlled by the AHT-inducible PtetA promoter.

To measure rpoS expression, a chromosomal transcriptional rpoS-gfp+ fusion was constructed as follows: the gfp+-frt-aph-frt module of pNAW52 (including the gfp+ SD) was inserted after the main rpoS transcription start site (TSS), previously mapped at the coordinate 3089613 on the 4/74 chromosome [111].

To measure dctA expression, a chromosomal transcriptional/translational dctA::sfgfp fusion was constructed: the flexible amino acid linker GSAGSAAGSGEF and the sequence encoding for superfolder GFP (sfgfp) were fused to the dctA C-term, using a sfgfp-frt-aph-frt module amplified from pNAW62 [95].

For complementation in oxyS and rpoS null mutants, a single copy of these genes was inserted in a non-transcribed chromosomal locus in two steps: first, an antibiotic cassette was inserted upstream of the oxyS or rpoS promoters, coordinates 4364521 and 3089777, respectively. Then, the aph-oxyS and the cat-rpoS modules were PCR-amplified from the resulting strains and inserted into the non-transcribed pseudogene STM474_1565 (between coordinates 1585970–1586170 on the 4/74 chromosome) [95]. The STM474_1565 gene is also known as STM1553 or SL1344_1483.

For scarless transfer of rrsAmut and rrsHmut mutations located in the 16S rRNA genes, the two step λ red recombination-based methodology described by Blank and colleagues was used [112]. The rrsA and rrsH genes (including their promoters) were replaced by a I-SceI-aph module, amplified from plasmid pKD4-I-SceI [107]. The mutations were transduced in 4/74 WT, yielding the strains ΔrrsA::(I-SceI-aph) and ΔrrsA::(I-SceI-aph). The two strains were transformed with the λ red plasmid pWRG99, expressing the I-SceI nuclease in the presence of AHT [112]. Electro-competent cells were prepared with the two Salmonella+pWRG99 strains in the presence of arabinose, as described above. Competent bacteria were electroporated with 2 μg of the rrsAmut or the rrsHmut fragments, obtained by PCR from spontaneous mutants SNW245 (rrsAmut) and SNW246 (rrsHmut). Replacement of the I-SceI-aph module by the corresponding rrs mutated PCR fragment was selected on LB agar supplemented with Ap and AHT at 30°C. The rrsAmut or rrsHmut insertions were confirmed by PCR and Sanger sequencing and the temperature-sensitive pWRG99 plasmid was eliminated by a passage at 42°C.

For other scarless genome editing procedures, the pEMG-based allele exchange technique [93] was used as previously described [107,113]. For the transfer of the dctAmut1, dctAmut2 and oxyRmut mutations, fragments encompassing the mutations and the ~500 bp flanking regions were PCR-amplified from the corresponding spontaneous Succ+ mutants. For the construction of the oxySGG mutant, two fragments flanking the mutations were PCR amplified with primer pairs NW_1021/NW_1022 and NW_1023/NW_1024. The two amplicons carrying the CC➔GG mutation on one of their extremities were fused by overlap extension PCR. All the PCR fragments carrying the mutations were inserted by digestion/ligation between the EcoRI and BamHI sites of the suicide plasmid pEMG and E. coli S17-1 λpir was transformed with the resulting ligation reactions. The resulting suicide plasmids were mobilised into the recipient Salmonella by conjugation and recombinants were selected on M9-glucose agar plates containing Km. Merodiploid resolution was carried out with the pSW-2 plasmid, as previously described [107]. The relevant mutations were confirmed in KmS candidates by PCR and Sanger sequencing. Finally, the unstable pSW-2 plasmid was eliminated by 2–3 passages on LB.

Experimental evolution to select Succ+ mutants

A 4/74 strain with an additional rpoS copy inserted in the STM474_1565 pseudogene (strain rpoS2X, SNW226, CmR) was transformed with pRL27 and ten libraries of approximately 10,000 Tn5 mutants were grown aerobically in 25 mL of M9+Succ containing Cm and Km in ten 250 mL conical flasks. After 48 hours incubation at 37°C, the cultures were spread on LB + Km plates and isolated colonies were passaged twice on LB agar. Growth on M9+Succ plates was assessed for >10 isolates per library. The Tn5 insertions from fast-growing colonies (Succ+ phenotype) were P22-transduced into 4/74 WT, and the Succ+ status of the transductants was verified.

For the isolation of Succ+ spontaneous 4/74 mutants, bacteria obtained from stationary phase LB cultures were washed with PBS and the OD600 was adjusted to 0.1 (~108 CFU/mL). Approximately 107 CFU (100 μL) were spread on M9+Succ agar and the plates were incubated at 37°C, until Succ+ large colonies were visible (3–4 days). Alternatively, spontaneous mutants were obtained from liquid M9+Succ cultures (25 mL), inoculated with ~100 Salmonella. The cultures were grown aerobically at 37°C, until substantial growth was observed (typically after 3 days incubation) and the cultures were spread on LB agar plates. All the presumed Succ+ spontaneous mutants were passaged twice on LB plates before confirming the Succ+ phenotype on M9+Succ agar plates. The rpoS positive status (rpoS+) of each mutant was tested by phenotypic assays (see below) and confirmed by PCR and Sanger sequencing, using primers NW_403, NW_252 and NW_252.

The genomes of a collection of rpoS+ Succ+ mutants were sequenced by the Illumina whole genome sequencing service of MicrobesNG (Birmingham, UK). Mutations were identified using the VarCap workflow [114] available on Galaxy (http://galaxy.csb.univie.ac.at:8080), using the published 4/74 genome as reference. The identified mutations were confirmed by PCR and Sanger sequencing, and were transferred into 4/74 WT using the two scarless genome editing techniques described above. After transfer, the Succ+ phenotype of all genome-edited mutants was confirmed.

Phenotypic characterisation of the Succ+ mutants

Phenotypes linked to the rpoS status were tested for each of the Succ+ Tn5 or Succ+ spontaneous mutants. The RpoS-dependent catalase activity was assessed with hydrogen peroxide directly on colonies or with stationary phase LB cultures, as described earlier [115,116]. The RpoS-dependent RDAR (red, dry and rough) morphotype [81] was tested by adding 2 μL of a stationary phase LB cultures on LBO agar plates containing 40 μg/mL of Congo Red. The RDAR morphotype was observed after at least 3 days of incubation at room temperature.

Mapping of Tn5 insertion sites

The Tn5 insertion sites of the Succ+ mutants were mapped by an arbitrary PCR approach [117]. For each Tn5 mutant, the arbitrary PCRs were carried out directly from colonies with the primer pair NW_319/NW_320 (0.5 μM each) and MyTaq Red Mix 2 X in a final volume of 20 μL. The arbitrary PCR conditions were: 95°C 120 sec; 6 X [95°C 15 sec 30°C; 30 sec; 72°C 90 sec]; 30 X [95°C 15 sec; 50°C 30 sec; 72°C 90 sec]; 72°C 300 sec; 4°C. The amplicons were purified on column and eluted in 20 μL of water. For the nested PCRs, 2 μL of the arbitrary PCR products were used as template and mixed with primer pair NW_318/NW_321 (0.5 μM each) and MyTaq Red Mix 2 X in a final volume of 40 μL. The second PCR conditions were: 95°C 120 sec; 30 X [95°C 15 sec; 50°C 30 sec; 72°C 90 sec]; 72°C 300 sec. The PCR products were separated by electrophoresis on a 1% agarose gel containing Midori Green for DNA UV-visualization. The most prominent DNA bands were excised, the DNA was purified and Sanger-sequenced with primer NW_318. Insertions were mapped by BLAST, using the 4/74 genome as reference.

Quantification of GFP fluorescence intensity

Strains carrying the chromosomal fusion rpoS-gfp+ or the plasmid-borne yobF::sfgfp translational fusion were grown in the indicated conditions in biological triplicates and bacteria were harvested by centrifugation and re-suspended in the same volume of PBS. The GFP signal was measured with a FLUOStar Omega plate reader (BMG Labtech) with 200 μL of bacterial suspension per well in black microplates (Greiner #655090). For each strain, the PBS background fluorescence was subtracted from the GFP signal (in arbitrary unit [a.u.]). The fluorescence values were divided by the OD600 of the cell suspensions. The fluorescence background of a WT unlabeled strain (carrying the empty plasmid pNAW125, when required) was measured similarly, and was subtracted from the fluorescence signal of the GFP-labelled strain. For each strain the GFP fluorescence intensity is represented as absolute values (GFP fluorescence intensity/OD600 [a.u.]) or as a relative GFP fluorescence intensity (in %).

To measure the activity of the dctA::sfgfp chromosomal fusion, bacteria were grown in M9+Gly+Succ to OD600 0.5–1. Cells were harvested and re-suspended in PBS, prior to fixation with 4% paraformaldehyde and washes with PBS [118]. To measure the GFP fluorescence intensity in all the strains carrying dctA::sfgfp, the IntelliCyt iQue Screener PLUS (Sartorius) was used. To quantify the dctA::sfgfp activity more precisely in the WT, dctAmut1, dctAmut2, ΔiscR and ΔrpoS genetic backgrounds, bacteria were grown in biological triplicates and were fixed with paraformaldehyde. The FITC-H GFP fluorescence intensity (median of the population) was measured using a FACSCanto II flow cytometer (BD Biosciences). The fluorescence background of a WT unlabelled strain was subtracted from the fluorescence intensity of each dctA::sfgfp carrying strain. The data are represented as the GFP fluorescence intensity of each mutant, relative to the intensity of the WT isogenic strain (%). All flow cytometry data were analysed using the FlowJo software (BD Biosciences).

For fluorescence microscopy, the dctA::sfgfp-carrying strains were grown in M9+Succ or M9+Gly+Succ and bacteria were immobilised in PBS solidified with 0.75% low melting point agarose. Pictures were taken with the EVOS FL cell imaging system (Thermo Fisher), as previously described [102].

RpoS detection by Western blotting

The strains of interest were grown in the indicated condition, and bacteria (~109 CFU, estimated by OD600) were pelleted by centrifugation and stored at -80°C. Bacteria were re-suspended in 100 μL of PBS and 10 μL of the cell suspensions were mixed with 990 μL of PBS to measure the OD600 1/100 of each suspension. Bacteria were lysed by adding 100 μL Laemmli Buffer 2 X [120 mM Tris-HCl pH 6.8, 4% (wt/vol) SDS, 20% (vol/vol) glycerol, Bromophenol blue 0.02% (wt/vol)] and 10 μL β-mercaptoethanol (5% vol/vol final). The lysates were boiled for 15 min, chilled on ice for 1 min and spun down for 5 min at 4°C (14,000 rpm). Bacterial extracts were separated by SDS polyacrylamide gel electrophoresis and the proteins RpoS and DNaK were detected by Western Blotting, as described earlier [105]. The volume of protein extract loaded (~10 μL) on the SDS 10% polyacrylamide gel was normalised by the OD600 of the bacterial suspensions prior to lysis.

Catalogue numbers of all antibodies are listed in S1 Table. The primary antibodies, Anti-E. coli RNA Sigma S Antibody (diluted 1:5,000) and anti-DnaK mAb 8E2/2 (diluted 1:10,000), were used for the detection of RpoS and DnaK (loading control), respectively. For detection, the secondary antibody Goat anti-mouse IgG (H + L)-HRP (diluted 1:2,500) and the Pierce ECL Western blotting substrate were used. The chemiluminescent reaction was detected with the ImageQuant LAS 4000 imaging system (GE Healthcare Life Sciences).

OxyS detection by Northern blotting

The WT, ΔoxyR and oxyRmut strains were grown in 25 mL of LB to OD600 = 1 and the cultures were split in two 10 mL subcultures. For each strain, hydrogen peroxide (H2O2, 2 mM) was added to one of the subcultures. After 30 min at 37°C, cellular RNA transcription and degradation processes were stopped by adding 4 mL of ice-cold STOP solution (95% ethanol + 5% acid phenol) to the 10 mL cultures. After a 30 min incubation on ice, bacteria were pelleted by centrifugation and total RNA was extracted with Trizol, as described previously [102].

Probe synthesis and OxyS sRNA detection by Digoxigenin (DIG)-based Northern blotting were carried out with the DIG Northern Starter Kit, according to the DIG Application Manual for Filter Hybridization (Roche) and a previous study [102]. Briefly, heat-denatured RNA (2.5 μg) was separated on an 8.3 M urea, 7% polyacrylamide gel in TBE 1X. RNA was transferred to a positively charged nylon membrane with the Bio-Rad Semi Dry transfer system (#170–3940). RNA was UV-crosslinked to the membrane before hybridization with the DIG-anti-OxyS probe in DIG Easy Hyb buffer at 68°C for 20 hours. The membrane was washed and the OxyS transcripts were detected using the Anti-Digoxigenin antibody and the CDP-Star substrate. Finally, the chemiluminescent reaction was visualised using the ImageQuant LAS 4000 imager. After OxyS detection, the membrane was stripped and re-probed with the DIG-anti-5S probe to detect the 5S ribosomal RNA, used as a loading control. The ssRNA DIG-labelled probes DIG-anti-OxyS and DIG-anti-5S were synthesised with the T7 polymerase and DNA templates obtained by PCR with template 4/74 gDNA and primer pairs NW_485/NW_485 and DH58/DH59, respectively.

Quantification and statistical analysis

Numerical data were plotted and analysed using Microsoft Excel (version 16.46) and GraphPad Prism 9.3.1. Data are presented as the mean of three to six biological replicates ± standard deviation, as indicated in the figures. The unpaired t-test was used to compare the groups and statistical significance is indicated on the figures. P values (two-tail) are reported using the following criteria: 0.0001 to 0.001 = ***, 0.001 to 0.01 = **, 0.01 to 0.05 = *, ≥ 0.05 = NS 

Supporting information

S1 Text. Hfq, PNPase and their cognate sRNAs maintain the inhibition of succinate utilisation.

https://doi.org/10.1371/journal.pgen.1011142.s001

(DOCX)

S2 Text. Anti-Shine-Dalgarno mutations and sub-inhibitory concentration of chloramphenicol boost succinate utilisation.

https://doi.org/10.1371/journal.pgen.1011142.s002

(DOCX)

S1 Fig. Growth assessment of Salmonella mutants with diverse C-sources involving a chromosomal construct to complement the rpoS mutation.

(A) the growth of a collection of S. Typhimurium 4/74 mutants lacking regulatory proteins was assessed on solidified M9+Succ agar and in liquid M9+Succ medium (in microplates), revealing the fast growth of mutants ΔrpoS (JH3674), Δhfq (JH3584) and Δpnp (JH3649). (B&C) Chromosomal complementation of the ΔrpoS mutation: a copy of rpoS (including its native promoter, bent arrow), linked to the cat Cm resistance gene was inserted in the non-transcribed pseudogene STM474_1565 of 4/74 WT (strain 4/74 rpoS2X, SNW226) and of ΔrpoSrpoS + rpoS, JH4160). The STM474_1565 gene is also known as STM1553 or SL1344_1483. RpoS-dependent phenotypes were assessed for each strain: growth was tested in M9+Succ (B) and RDAR phenotype was tested on Congo Red agar plates, confirming the Succ- RDAR+ of the complemented strain ΔrpoS + rpoS. (D) The growth of the novel Succ+ mutants identified (presented in Fig 3) was tested on solidified M9 minimal medium supplemented with 40 mM citrate, malate or fumarate. The growth was assessed with biological triplicates after the indicated incubation time (37°C) and the growth of each mutant was compared to 4/74 WT (Succ-) and ΔrpoS (Succ+).

https://doi.org/10.1371/journal.pgen.1011142.s003

(TIF)

S2 Fig. Genetic dissection of the fliDST operon reveals that fliS and fliT can inhibit succinate utilisation.

(A) Schematic representation of the fliDST operon. The transcription start site (+1 TSS) and the two Tn5 transposon insertions causing Succ+ phenotype are depicted (Table 1). (B) The inactivation of the flagellar chaperones FliS and FliT stimulates Salmonella growth with succinate. The growth of 4/74 WT and of mutants ΔfliD (SNW278), ΔfliS (SNW280), ΔfliT (SNW282), ΔfliDST (SNW284), ΔfliDS (SNW286), ΔfliST (SNW288) was assessed on M9+Succ agar plates after 48 hours of incubation at 37°C.

https://doi.org/10.1371/journal.pgen.1011142.s004

(TIF)

S3 Fig. Genetic complementation of eight Succ+ regulatory mutations.

(A) The plasmids used for the complementation experiments are depicted and were constructed (Methods) using the backbone of pXG10-SF [99], a low-copy plasmid encoding for the cat resistance gene and carrying the ori101* replicon. Each plasmid carries the gene(s) of interest under the control of the strong constitutive promoter PL tetO-1 [101], except for pnp, that is controlled by its native promoter. For each growth curve (B-I), the plasmid pXG10-SF (“vector”) expressing the lacZ186::sfgfp fusion was used as a negative control. The strains 4/74 WT, ΔrpoS (JH3674), Δhfq (JH3584), Δpnp (JH3649), ΔiraP (SNW188), ΔiscR (SNW184), ΔrbsR (SNW294), ΔcspC (SNW292) and ΔfliST (SNW288) carrying the indicated plasmids were grown in M9+Succ, supplemented with 25 μg/mL Cm. (J) The presence of Cm (25 μg/mL) in M9+Succ stimulates mildly the growth of 4/74 WT carrying pXG10-SF. The growth curves were carried out with 6 replicates in 96-well plates.

https://doi.org/10.1371/journal.pgen.1011142.s005

(TIF)

S4 Fig. Succ+ mutations reduce the lag time in M9+Succ.

(A) The lag time was defined as the time required to reach an OD600 of 0.1. The lag time was determined for the regulatory mutants, ΔrpoS, Δhfq, ΔiraP, ΔfliST, ΔcspC, ΔiscR ΔrbsR, and Δpnp and for the corresponding complemented strains, based on the growth curves presented in S3 Fig. (B) The lag time was determined for the mutants oxyRmut, dctAmut1, dctAmut2, rssHmut and rrsAmut, based on the growth curves presented in Figs 5, 6, and S11.

https://doi.org/10.1371/journal.pgen.1011142.s006

(TIF)

S5 Fig. Hfq, PNPase and sRNAs inhibit Salmonella growth with succinate.

(A) Hfq and PNPase inactivation boosts Salmonella growth on succinate. (B) The co-inactivations of the rpoS activating sRNAs ArcZ, DsrA and RprA or of the sdh repressing sRNAs RyhB-1/2, Spf and RybB did not stimulate Salmonella growth with succinate. (C) Successive inactivations of ArcZ, DsrA and RprA in the ΔryhB-1 ryhB-2 rybB spf genetic background stimulate gradually the growth with succinate. (D) The overexpression of rpoS abolishes totally the Succ+ phenotype of the Δpnp mutant, lacking PNPase: growth was assessed for strains 4/74 WT and Δpnp, carrying the empty plasmid (vector, pNAW125) or the prpoS (pNAW95) plasmid, overexpressing rpoS. The strains used were all 4/74 derivatives: Δhfq (JH3584), Δpnp (JH3649), ΔarcZ dsrA rprA (JH4385), ΔryhB-1 ryhB-2 rybB spf (SNW630), ΔryhB-1 ryhB-2 rybB spf arcZ (SNW639), ΔryhB-1 ryhB-2 rybB spf arcZ dsrA (SNW640) and ΔryhB-1 ryhB-2 rybB spf arcZ dsrA rprA (SNW641). The medium used is indicated for each experiment. Growth curves were carried out with 6 replicates grown in 96-well plates, as specified in Methods.

https://doi.org/10.1371/journal.pgen.1011142.s007

(TIF)

S6 Fig. Salmonella growth with succinate is not stimulated by iron supplementation or the inactivation of sRNAs RyhB-1 and RyhB-2.

Strains 4/74 and ΔryhB-1 ryhB-2ryhB-1/2, JH4390) were grown in M9+Succ medium supplemented or not with iron (FeCl3) at the indicated concentration. The growth curves were carried out with 6 replicates in 96-well plates in the indicated medium.

https://doi.org/10.1371/journal.pgen.1011142.s008

(TIF)

S7 Fig. Detailed schematic representation of the oxyS-oxyR intergenic region and of the ΔoxyS, ΔoxyR and oxyRchr+ constructs.

The intergenic region sequence is depicted and the -35 and -10 boxes of the PoxyS and PoxyR promoters are highlighted in blue, according to the corresponding locus of E. coli K-12 [52]. The ΔoxyS and ΔoxyR mutation are indicated. The transcription start sites (TSS) are indicated, according to the SalcomMac transcriptomic database [88,111]. For the complementation of the ΔoxyS mutation in strain SNW362 (oxySchr+), the oxyS gene, its native promoter and a KmR cassette were inserted into the non-transcribed pseudogene STM474_1565 [95].

https://doi.org/10.1371/journal.pgen.1011142.s009

(TIF)

S8 Fig. DctA is essential for long-term aerobic growth and emergence of Succ+ mutants in M9+Succ medium.

Growth of S. Typhimurium 4/74 WT displays an extended lag time, while ΔdctA does not grow at all in M9+Succ medium. The growth curves of three independent cultures (I-III) of 4/74 WT and of ΔdctA in M9+Succ are presented. bacteria were grown at 37°C with aeration in 25 ml of M9+Succ (in 250 ml conical flasks) with an initial inoculum of ~107 CFU/mL (OD600 = 0.01).

https://doi.org/10.1371/journal.pgen.1011142.s010

(TIF)

S9 Fig. Stimulation of dctA expression in the dctAmut1, dctAmut2 and ΔiscR mutants.

(A-N) Strains carrying the chromosomal transcriptional/translational fusion dctA::sfgfp were grown in M9+Gly+Succ minimal medium (OD600~1) and the GFP fluorescence intensity was measured with the IntelliCyt iQue Screener PLUS (Sartorius) after bacteria fixation with formaldehyde. The 4/74 WT (untagged strain) was used as a negative control (A). Each Succ+ mutants carrying dctA::sfgfp was compared with the “WT” strain carrying the same fusion (SNW296, in grey). (O) The dctA::sfgfp tagged strain ΔiscR, dctAmut1, dctAmut2 and ΔrpoS grow fast in M9+Succ in comparison with the dctA::sfgfp tagged WT strain, showing that the fusion of sfGFP to the C-term of DctA does not impede the DctA-driven uptake of succinate. (P) The same strains were grown in M9+Succ (OD600~1) and the dctA::sfgfp induction was observed by fluorescence microscopy, as specified in Methods. The dctA::sfgfp tagged Succ+ mutants used for these experiments were: ΔrpoS (SNW313), Δhfq (SNW309), ΔiraP (SNW423), ΔfliST (SNW330), ΔcspC (SNW424), ΔiscR (SNW329), ΔrbsR (SNW425), Δpnp (SNW437), dctAmut1 (SNW310), dctAmut2 (SNW316), oxyRmut (SNW426), rrsAmut (SNW374) and rrsHmut (SNW331).

https://doi.org/10.1371/journal.pgen.1011142.s011

(TIF)

S10 Fig. RbsR stimulates rpoS expression at the protein level but not at the transcriptional level.

(A) Western blot detection of RpoS and DnaK (loading control) in 4/74 WT and mutants ΔrpoS (JH3674), ΔiraP (SNW188), ΔcspC (SNW292), ΔfliST (SNW288), ΔrbsR (SNW294) and oxyRmut (SNW318) grown in LB to OD600 1 and 2.5. (B) Two independent replicates of the Western blot analyses presented in Fig 7A confirmed the down-regulation of rpoS in the ΔrbsR mutant (see Fig 7A legend). (C) Schematic representation of chromosomal PrpoS-gfp transcriptional fusion. The gfp+ gene and its Shine-Dalgarno (SD) were inserted downstream of the main promoter of rpoS (PrpoS, bent arrow), interrupting the nlpD gene. The residual FLP recognition target site sequence is denoted by”frt”. The PrpoS-gfp fusion was inserted in 4/74 WT and in ΔrbsR, resulting in strain SNW367 and SNW368, respectively. (D) The PrpoS-gfp fusion activity was measured in the WT and ΔrbsR genetic background in bacteria grown in LB to late exponential phase (LEP, OD600~1), early stationary phase (ESP, OD600~2) and stationary phase (SP, OD600~4). The GFP fluorescence intensity (absolute values) were measured, as specified in Methods. The data are presented as the average of biological triplicates ± standard deviation. The difference of fluorescence intensities between the WT and the ΔrbsR strains were not significant (NS) in the three conditions tested, as defined in Methods.

https://doi.org/10.1371/journal.pgen.1011142.s012

(TIF)

S11 Fig. Mutation of the 16S rRNA aSD motifs and sub-inhibitory concentrations of chloramphenicol stimulate Salmonella growth upon succinate.

(A) Schematic representation of the Salmonella rrnA and rrnH ribosomal RNA (rRNA) operons. The 23S (rrl), 16S (rrs) and 5S (rrf) rRNAs and the ileT, ileV, alaT and alaU tRNAs are represented, according to the annotation of the corresponding loci of S. Typhimurium LT2 (Genbank AE006468.2) [130]. The bent arrows represent the ribosomal promoter. The replacement of the full rrsA and rrsH loci (promoters included) with an I-SceI-Km cassette (Methods) in strains ΔrrsA (SNW335) and ΔrrsH (SNW311) is represented by the “Δ” symbol. The SNP mutations in the anti-shine-Dalgarno (aSD) motifs of mutant rrsAmut (SNW336) and rrsHmut (SNW314) are indicated in red. (B) The aSD mutations rrsAmut and rrsHmut stimulate Salmonella growth with succinate, while the full inactivation of the rrsA and rrsH loci (strains ΔrrsA and ΔrrsH) did not affect the growth (C). The plasmid borne overexpression of rpoS has moderate effects on the growth of the rrsAmut (D) and rrsAmut (E) mutants with succinate. The 4/74 WT and the rrs mutants carried the empty plasmid (Vector, pNAW125) or the prpoS (pNAW95) plasmid. (F) Subinhibitory concentrations of chloramphenicol (Cm) stimulate Salmonella growth with succinate. All the growth curves were carried out with 6 replicates in 96-well plates with the indicated medium.

https://doi.org/10.1371/journal.pgen.1011142.s013

(TIF)

S12 Fig. Sub-inhibitory concentrations of chloramphenicol, but not of tetracycline, stimulate Salmonella growth with succinate.

(A) Low concentration of chloramphenicol (Cm) stimulates the growth of S. Enteritidis strain P125109 and of S. Typhimurium strain 4/74 with succinate, while tetracycline (Tc) does not affect the growth profile (B).

https://doi.org/10.1371/journal.pgen.1011142.s014

(TIF)

S1 Table. Key resources, DNA oligonucleotides, plasmids and bacterial strains used in this study.

https://doi.org/10.1371/journal.pgen.1011142.s015

(XLSX)

S1 Data. Numerical values and statistical analysis underlying each figure.

https://doi.org/10.1371/journal.pgen.1011142.s016

(XLSX)

Acknowledgments

We are very grateful to Aoife Colgan for the construction of several mutants, to Stefano Marzi and Laurent Aussel for helpful advice, and to Paul Loughnane for his expert technical assistance. We also thank Celeste Peterson and Susan Gottesman for sharing preliminary data about the Rbs-driven regulation of RpoS. Finally, we thank all the present and former members of the Hinton Lab for helpful and productive discussions (in the AJ and elsewhere).

References

  1. 1. Rohmer L, Hocquet D, Miller S I. Are pathogenic bacteria just looking for food? Metabolism and microbial pathogenesis. Trends Microbiol. 2011;19: 341–348. pmid:21600774
  2. 2. Schubert C, Unden G. C4-dicarboxylates as growth substrates and signaling molecules for commensal and pathogenic enteric bacteria in mammalian intestine. J Bacteriol. 2022;Jan 3. pmid:34978458
  3. 3. De Vadder F, Kovatcheva-Datchary P, Zitoun C, Duchampt A, Bäckhed F, Mithieux G. Microbiota-Produced Succinate Improves Glucose Homeostasis via Intestinal Gluconeogenesis. Cell Metab. 2016;24: 151–157. pmid:27411015
  4. 4. Petersen E, Miller S I. The cellular microbiology of Salmonellae interactions with macrophages. Cell Microbiol. 2019;21. pmid:31509644
  5. 5. Spiga L, Winter M G, Furtado de Carvalho T, Zhu W, Hughes E R, Gillis C C, et al. An Oxidative Central Metabolism Enables Salmonella to Utilize Microbiota-Derived Succinate. Cell Host Microbe. 2017. pmid:28844888
  6. 6. Haraga A, Ohlson M B, Miller S I. Salmonellae interplay with host cells. Nat Rev Microbiol. 2008;6: 53–66. pmid:18026123
  7. 7. Fàbrega A, Vila J. Salmonella enterica Serovar Typhimurium Skills To Succeed in the Host: Virulence and Regulation. Clin Microbiol Rev. 2013;26: 308–341. pmid:23554419
  8. 8. Murphy M P O’Neill L A J. Krebs Cycle Reimagined: The Emerging Roles of Succinate and Itaconate as Signal Transducers. Cell. 2018. pmid:30096309
  9. 9. Mitosch K, Beyß M, Phapale P, Drotleff B, Nöh K, Alexandrov T, et al. A pathogen-specific isotope tracing approach reveals metabolic activities and fluxes of intracellular Salmonella. PLOS Biol. 2023;21: 1–31. pmid:37594988
  10. 10. Mills E O’Neill L A J. Succinate: A metabolic signal in inflammation. Trends in Cell Biology. 2014. pmid:24361092
  11. 11. Ryan D G O’Neill L A J. Krebs Cycle Reborn in Macrophage Immunometabolism. Annual Review of Immunology. 2020. pmid:31986069
  12. 12. Jiang L, Wang P, Song X, Zhang H, Ma S, Wang J, et al. Salmonella Typhimurium reprograms macrophage metabolism via T3SS effector SopE2 to promote intracellular replication and virulence. Nat Commun. 2021;12: 879. pmid:33563986
  13. 13. Bowden S D, Rowley G, Hinton J C D, Thompson A. Glucose and Glycolysis Are Required for the Successful Infection of Macrophages and Mice by Salmonella enterica Serovar Typhimurium. Infect Immun. 2009;77: 3117–3126. pmid:19380470
  14. 14. Rosenberg G, Yehezkel D, Hoffman D, Ciolli Mattioli C, Fremder M, Ben-Arosh H, et al. Host succinate is an activation signal for Salmonella virulence during intracellular infection. Science (80-). 2021. pmid:33479153
  15. 15. Bowden S D, Ramachandran V K, Knudsen G M, Hinton J C D, Thompson A. An Incomplete TCA Cycle Increases Survival of Salmonella Typhimurium during Infection of Resting and Activated Murine Macrophages. PLoS One. 2010. pmid:21079785
  16. 16. Hersch SJ, Radan B, Ilyas B, Lavoie P, Navarre W W. A stress-induced block in dicarboxylate uptake and utilization in Salmonella. J Bacteriol. 2021; JB.00487–20. pmid:33593945
  17. 17. Cecchini G, Schröder I, Gunsalus R P, Maklashina E. Succinate dehydrogenase and fumarate reductase from Escherichia coli. Biochim Biophys Acta—Bioenerg. 2002;1553: 140–157. https://doi.org/10.1016/S0005-2728(01)00238-9
  18. 18. Surmann K, Stopp M, Wörner S, Dhople V M, Völker U, Unden G, et al. Fumarate dependent protein composition under aerobic and anaerobic growth conditions in Escherichia coli. J Proteomics. 2020;212: 103583. pmid:31734389
  19. 19. Fang F C, Libby S J, Buchmeier N A, Loewen P C, Switala J, Harwood J, et al. The alternative σ factor KatF (RpoS) regulates Salmonella virulence. Proc Natl Acad Sci U S A. 1992. pmid:1465428
  20. 20. Soo Lee I, Lin J, Hall H K, Bearson B, Foster J W. The stationary-phase sigma factor σS (RpoS) is required for a sustained acid tolerance response in virulent Salmonella typhimurium. Mol Microbiol. 1995;17. pmid:7476202
  21. 21. Knudsen G M, Nielsen M-B, Grassby T, Danino-Appleton V, Thomsen L E, Colquhoun I J, et al. A third mode of surface-associated growth: immobilization of Salmonella enterica serovar Typhimurium modulates the RpoS-directed transcriptional programme. Environ Microbiol. 2012;14: 1855–1875.
  22. 22. Lago M, Monteil V, Douche T, Guglielmini J, Criscuolo A, Maufrais C, et al. Proteome remodelling by the stress sigma factor RpoS/σS in Salmonella: Identification of small proteins and evidence for post-transcriptional regulation. Sci Rep. 2017. pmid:28522802
  23. 23. Robbe-Saule V, Lopes M D, Kolb A, Norel F. Physiological effects of crl in Salmonella are modulated by σS level and promoter specificity. J Bacteriol. 2007. pmid:17293430
  24. 24. Lévi-Meyrueis C, Monteil V, Sismeiro O, Dillies M A, Monot M, Jagla B, et al. Expanding the RpoS/σS-network by RNA sequencing and identification of σS-controlled small RNAs in Salmonella. PLoS One. 2014. pmid:24810289
  25. 25. Lévi-Meyrueis C, Monteil V, Sismeiro O, Dillies M A, Kolb A, Monot M, et al. Repressor activity of the RpoS/σS-dependent RNA polymerase requires DNA binding. Nucleic Acids Res. 2015. pmid:25578965
  26. 26. Blount Z D, Lenski R E, Losos J B. Contingency and determinism in evolution: Replaying life’s tape. Science (80-). 2018;362: eaam5979. pmid:30409860
  27. 27. Warsi O, Lundin E, Lustig U, Näsvall J, Andersson D I. Selection for novel metabolic capabilities in Salmonella enterica. Evolution (N Y). 2019;73: 990–1000. pmid:30848832
  28. 28. Unden G, Strecker A, Kleefeld A, Bin Kim O. C4-Dicarboxylate Utilization in Aerobic and Anaerobic Growth. EcoSal Plus. 2016. pmid:27415771
  29. 29. Robbe-Saule V, Algorta G, Rouilhac I, Norel F. Characterization of the RpoS status of clinical isolates of Salmonella enterica. Appl Environ Microbiol. 2003.
  30. 30. Dong T, Schellhorn H E. Control of RpoS in global gene expression of Escherichia coli in minimal media. Mol Genet Genomics. 2009. pmid:18843507
  31. 31. Chiang S M, Dong T, Edge T A, Schellhorn H E. Phenotypic diversity caused by differential RpoS activity among environmental Escherichia coli isolates. Appl Environ Microbiol. 2011. pmid:21948830
  32. 32. Kingsley R A, Msefula C L, Thomson N R, Kariuki S, Holt K E, Gordon M A, et al. Epidemic multiple drug resistant Salmonella Typhimurium causing invasive disease in sub-Saharan Africa have a distinct genotype. Genome Res. 2009. pmid:19901036
  33. 33. Feasey N A, Hadfield J, Keddy K H, Dallman T J, Jacobs J, Deng X, et al. Distinct Salmonella Enteritidis lineages associated with enterocolitis in high-income settings and invasive disease in low-income settings. Nat Genet. 2016. pmid:27548315
  34. 34. Perez-Sepulveda B M, Predeus A V, Fong W Y, Parry C M, Cheesbrough J, Wigley P, et al. Complete Genome Sequences of African Salmonella enterica Serovar Enteritidis Clinical Isolates Associated with Bloodstream Infection. Bruno V, editor. Microbiol Resour Announc. 2021;10. pmid:33766909
  35. 35. Pulford C V, Wenner N, Redway M L, Rodwell E V, Webster H J, Escudero R, et al. The diversity, evolution and ecology of Salmonella in venomous snakes. PLoS Negl Trop Dis. 2019. pmid:31163033
  36. 36. den Bakker H C, Moreno Switt A I, Govoni G, Cummings C A, Ranieri M L, Degoricija L, et al. Genome sequencing reveals diversification of virulence factor content and possible host adaptation in distinct subpopulations of Salmonella enterica. BMC Genomics. 2011. pmid:21592414
  37. 37. Richardson E J, Limaye B, Inamdar H, Datta A, Manjari K S, Pullinger G D, et al. Genome sequences of Salmonella enterica serovar Typhimurium, Choleraesuis, Dublin, and Gallinarum strains of well-defined virulence in food-producing animals. Journal of Bacteriology. 2011. pmid:21478351
  38. 38. Bialek-Davenet S, Criscuolo A, Ailloud F, Passet V, Jones L, Delannoy-Vieillard A S, et al. Genomic definition of hypervirulent and multidrug-resistant klebsiella pneumoniae clonal groups. Emerg Infect Dis. 2014. pmid:25341126
  39. 39. Schauer D B, Zabel B A, Pedraza I F, O’Hara C M, Steigerwalt A G, Brenner D J. Genetic and biochemical characterization of Citrobacter rodentium sp. nov. J Clin Microbiol. 1995.
  40. 40. Wilmes-Riesenberg M R, Foster J W, Curtiss R. An altered rpoS allele contributes to the avirulence of Salmonella typhimurium LT2. Infect Immun. 1997.
  41. 41. Van Der Woude M W, Bäumler A J. Phase and antigenic variation in bacteria. Clinical Microbiology Reviews. 2004. pmid:15258095
  42. 42. Sánchez-Romero M A, Casadesús J. The bacterial epigenome. Nature Reviews Microbiology. 2020. pmid:31728064
  43. 43. Kröger C, Dillon S C, Cameron A D S, Papenfort K, Sivasankaran S K, Hokamp K, et al. The transcriptional landscape and small RNAs of Salmonella enterica serovar Typhimurium. Proc Natl Acad Sci U S A. 2012. pmid:22538806
  44. 44. Colgan A M, Kröger C, Diard M, Hardt W D, Puente J L, Sivasankaran S K, et al. The Impact of 18 Ancestral and Horizontally-Acquired Regulatory Proteins upon the Transcriptome and sRNA Landscape of Salmonella enterica serovar Typhimurium. PLoS Genet. 2016. pmid:27564394
  45. 45. Kutsukake K, Ikebe T, Yamamoto S. Two novel regulatory genes, fliT and fliZ, in the flagellar regulon of Salmonella. Genes Genet Syst. 1999. pmid:10791024
  46. 46. De Bruijn F J, Lupski J R. The use of transposon Tn5 mutagenesis in the rapid generation of correlated physical and genetic maps of DNA segments cloned into multicopy plasmids—a review. Gene. 1984;27: 131–149.
  47. 47. Janausch I G, Zientz E, Tran Q H, Kröger A, Unden G. C4-dicarboxylate carriers and sensors in bacteria. Biochimica et Biophysica Acta—Bioenergetics. 2002. pmid:11803016
  48. 48. Storz G. New perspectives: Insights into oxidative stress from bacterial studies. Arch Biochem Biophys. 2016. pmid:27095210
  49. 49. Zheng M, Åslund F, Storz G. Activation of the OxyR transcription factor by reversible disulfide bond formation. Science (80-). 1998. pmid:9497290
  50. 50. Storz G, Tartaglia L A, Ames B N. The OxyR regulon. Antonie Van Leeuwenhoek. 1990. pmid:2256675
  51. 51. Chiang S M, Schellhorn H E. Regulators of oxidative stress response genes in Escherichia coli and their functional conservation in bacteria. Archives of Biochemistry and Biophysics. 2012. pmid:22381957
  52. 52. Altuvia S, Weinstein-Fischer D, Zhang A, Postow L, Storz G. A small, stable RNA induced by oxidative stress: Role as a pleiotropic regulator and antimutator. Cell. 1997. pmid:9230301
  53. 53. Zhang A, Altuvia S, Tiwari A, Argaman L, Hengge-Aronis R, Storz G. The OxyS regulatory RNA represses rpoS translation and binds the Hfq (HF-I) protein. EMBO J. 1998. pmid:9774349
  54. 54. Zhang A, Wassarman K M, Ortega J, Steven A C, Storz G. The Sm-like Hfq protein increases OxyS RNA interaction with target mRNAs. Mol Cell. 2002. pmid:11804582
  55. 55. Christman M F, Morgan R W, Jacobson F S, Ames B N. Positive control of a regulon for defenses against oxidative stress and some heat-shock proteins in Salmonella typhimurium. Cell. 1985. pmid:2988786
  56. 56. Christman M F, Storz G, Ames B N. OxyR, a positive regulator of hydrogen peroxide-inducible genes in Escherichia coli and Salmonella typhimurium, is homologous to a family of bacterial regulatory proteins. Proc Natl Acad Sci U S A. 1989. pmid:2471187
  57. 57. Kullik I, Toledano M B, Tartaglia L A, Storz G. Mutational analysis of the redox-sensitive transcriptional regulator OxyR: Regions important for oxidation and transcriptional activation. J Bacteriol. 1995. pmid:7868602
  58. 58. Tjaden B, Goodwin S S, Opdyke J A, Guillier M, Fu D X, Gottesman S, et al. Target prediction for small, noncoding RNAs in bacteria. Nucleic Acids Res. 2006. pmid:16717284
  59. 59. Hemm M R, Paul B J, Miranda-Ríos J, Zhang A, Soltanzad N, Storz G. Small stress response proteins in Escherichia coli: Proteins missed by classical proteomic studies. J Bacteriol. 2010. pmid:19734316
  60. 60. Sittka A, Pfeiffer V, Tedin K, Vogel J. The RNA chaperone Hfq is essential for the virulence of Salmonella typhimurium. Mol Microbiol. 2007. pmid:17163975
  61. 61. Phadtare S, Inouye M. Role of CspC and CspE in regulation of expression of RpoS and UspA, the stress response proteins in Escherichia coli. J Bacteriol. 2001.
  62. 62. Cohen-Or I, Shenhar Y, Biran D, Ron E Z. CspC regulates rpoS transcript levels and complements hfq deletions. Res Microbiol. 2010. pmid:20633642
  63. 63. Michaux C, Holmqvist E, Vasicek E, Sharan M, Barquist L, Westermann A J, et al. RNA target profiles direct the discovery of virulence functions for the cold-shock proteins CspC and CspE. Proc Natl Acad Sci U S A. 2017. pmid:28611217
  64. 64. Keto-Timonen R, Hietala N, Palonen E, Hakakorpi A, Lindström M, Korkeala H. Cold Shock Proteins: A Minireview with Special Emphasis on Csp-family of Enteropathogenic Yersinia. Front Microbiol. 2016. pmid:27499753
  65. 65. Battesti A, Majdalani N, Gottesman S. The RpoS-Mediated General Stress Response in Escherichia coli. Annu Rev Microbiol. 2011. pmid:21639793
  66. 66. Steinmetz P A, Wörner S, Unden G. Differentiation of DctA and DcuS function in the DctA/DcuS sensor complex of Escherichia coli: Function of DctA as an activity switch and of DcuS as the C4-dicarboxylate sensor. Mol Microbiol. 2014. pmid:25135747
  67. 67. Davies S J, Golby P, Omrani D, Broad S A, Harrington V L, Guest J R, et al. Inactivation and regulation of the aerobic C4-dicarboxylate transport (dctA) gene of Escherichia coli. sc. 1999.
  68. 68. Amin M R, Korchinski L, Yoneda J K, Thakkar R, Sanson C L A, Fitzgerald S F, et al. A mutation in the putative CRP binding site of the dctA promoter of Salmonella enterica serovar Typhimurium enables growth with low orotate concentrations. Can J Microbiol. 2022.
  69. 69. Zientz E, Bongaerts J, Unden G. Fumarate regulation of gene expression in Escherichia coli by the DcuSR (dcuSR genes) two-component regulatory system. J Bacteriol. 1998.
  70. 70. Giel J L, Rodionov D, Liu M, Blattner F R, Kiley P J. IscR-dependent gene expression links iron-sulphur cluster assembly to the control of O2-regulated genes in Escherichia coli. Mol Microbiol. 2006. pmid:16677314
  71. 71. Py B, Barras F. Building Feg-S proteins: Bacterial strategies. Nature Reviews Microbiology. 2010. pmid:20467446
  72. 72. Schwartz C J, Giel J L, Patschkowski T, Luther C, Ruzicka F J, Beinert H, et al. IscR, an Fe-S cluster-containing transcription factor, represses expression of Escherichia coli genes encoding Fe-S cluster assembly proteins. Proc Natl Acad Sci U S A. 2001. pmid:11742080
  73. 73. Yeo W S, Lee J H, Lee K C, Roe J H. IscR acts as an activator in response to oxidative stress for the suf operon encoding Fe-S assembly proteins. Mol Microbiol. 2006. pmid:16824106
  74. 74. Nesbit A D, Giel J L, Rose J C, Kiley P J. Sequence-Specific Binding to a Subset of IscR-Regulated Promoters Does Not Require IscR Fe-S Cluster Ligation. J Mol Biol. 2009. pmid:19361432
  75. 75. Quandt E M, Deatherage D E, Ellington A D, Georgiou G, Barrick J E. Recursive genomewide recombination and sequencing reveals a key refinement step in the evolution of a metabolic innovation in Escherichia coli. Proc Natl Acad Sci U S A. 2014. pmid:24379390
  76. 76. Vergnes A, Viala J P M, Ouadah-Tsabet R, Pocachard B, Loiseau L, Méresse S, et al. The iron–sulfur cluster sensor IscR is a negative regulator of Spi1 type III secretion system in Salmonella enterica. Cell Microbiol. 2017. pmid:27704705
  77. 77. Rajagopalan S, Teter S J, Zwart P H, Brennan R G, Phillips K J, Kiley P J. Studies of IscR reveal a unique mechanism for metal-dependent regulation of DNA binding specificity. Nat Struct Mol Biol. 2013. pmid:23644595
  78. 78. Zwir I, Latifi T, Perez J C, Huang H, Groisman E A. The promoter architectural landscape of the Salmonella PhoP regulon. Mol Microbiol. 2012. pmid:22435712
  79. 79. Bougdour A, Wickner S, Gottesman S. Modulating RssB activity: IraP, a novel regulator of σS stability in Escherichia coli. Genes Dev. 2006. pmid:16600914
  80. 80. Bougdour A, Cunning C, Baptiste P J, Elliott T, Gottesman S. Multiple pathways for regulation of σS (RpoS) stability in Escherichia coli via the action of multiple anti-adaptors. Mol Microbiol. 2008. pmid:18383615
  81. 81. Robbe-Saule V, Jaumouillé V, Prévost M C, Guadagnini S, Talhouarne C, Mathout H, et al. Crl activates transcription initiation of RpoS-regulated genes involved in the multicellular behavior of Salmonella enterica serovar Typhimurium. J Bacteriol. 2006. pmid:16707690
  82. 82. Mauzy C A, Hermodson M A. Structural and functional analyses of the repressor, RbsR, of the ribose operon of Escherichia coli. Protein Sci. 1992. pmid:1304369
  83. 83. Badaoui A, Smith I, Balla M, Cavanaugh M, Lockart J, Parsons D, et al. Ribose lowers RpoS translation through RbsD mRNA. bioRxiv. 2023.
  84. 84. Auvray F, Thomas J, Fraser G M, Hughes C. Flagellin polymerisation control by a cytosolic export chaperone. J Mol Biol. 2001. pmid:11327763
  85. 85. Bennett J C Q, Thomas J, Fraser G M, Hughes C. Substrate complexes and domain organization of the Salmonella flagellar export chaperones FlgN and FliT. Mol Microbiol. 2001. pmid:11169117
  86. 86. Faber F, Bäumler A J. The impact of intestinal inflammation on the nutritional environment of the gut microbiota. Immunology Letters. 2014. pmid:24803011
  87. 87. Bäumler A J, Sperandio V. Interactions between the microbiota and pathogenic bacteria in the gut. Nature. 2016;535: 85–93. pmid:27383983
  88. 88. Srikumar S, Kröger C, Hébrard M, Colgan A, Owen S V, Sivasankaran S K, et al. RNA-seq Brings New Insights to the Intra-Macrophage Transcriptome of Salmonella Typhimurium. PLoS Pathog. 2015. pmid:26561851
  89. 89. Chen C Y, Eckmann L, Libby S J, Fang F C, Okamoto S, Kagnoff M F, et al. Expression of Salmonella typhimurium rpoS and rpoS-dependent genes in the intracellular environment of eukaryotic cells. Infect Immun. 1996;64: 4739–4743.
  90. 90. Wray C, Sojka W J. Experimental Salmonella Typhimurium infection in calves. Res Vet Sci. 1978.
  91. 91. Simon R, Priefer U, Pühler A. A broad host range mobilization system for in vivo genetic engineering: Transposon mutagenesis in gram negative bacteria. Bio/Technology. 1983.
  92. 92. Sambrook J, Russell D W. Molecular Cloning: A Laboratory Manual, Third Edition. Molecular Cloning: a laboratory a manual,. 2001.
  93. 93. Martínez-García E, de Lorenzo V. Engineering multiple genomic deletions in Gram-negative bacteria: Analysis of the multi-resistant antibiotic profile of Pseudomonas putida KT2440. Environ Microbiol. 2011. pmid:21883790
  94. 94. Green R, Rogers E J. Chemical Transformation of E. coli. Methods Enzym. 2014; 329–336.
  95. 95. Owen S V, Wenner N, Dulberger C L, Rodwell E V, Bowers-Barnard A, Quinones-Olvera N, et al. Prophages encode phage-defense systems with cognate self-immunity. Cell Host Microbe. 2021;29. pmid:34597593
  96. 96. Edwards R A, Allen Helm R, Maloy S R. Increasing DNA transfer efficiency by temporary inactivation of host restriction. Biotechniques. 1999. pmid:10337482
  97. 97. Larsen R A, Wilson M M, Guss A M, Metcalf W W. Genetic analysis of pigment biosynthesis in Xanthobacter autotrophicus Py2 using a new, highly efficient transposon mutagenesis system that is functional in a wide variety of bacteria. Arch Microbiol. 2002. pmid:12189420
  98. 98. Van Den Ent F, Löwe J. RF cloning: A restriction-free method for inserting target genes into plasmids. J Biochem Biophys Methods. 2006. pmid:16480772
  99. 99. Corcoran C P, Podkaminski D, Papenfort K, Urban J H, Hinton J C D, Vogel J. Superfolder GFP reporters validate diverse new mRNA targets of the classic porin regulator, MicF RNA. Mol Microbiol. 2012. pmid:22458297
  100. 100. Heckman K L, Pease L R. Gene splicing and mutagenesis by PCR-driven overlap extension. Nat Protoc. 2007. pmid:17446874
  101. 101. Lutz R, Bujard H. Independent and tight regulation of transcriptional units in Escherichia coli via the LacR/O, the TetR/O and AraC/I1-I2 regulatory elements. Nucleic Acids Res. 1997;25: 1203–1210. pmid:9092630
  102. 102. Owen S V, Canals R, Wenner N, Hammarlöf D, Kröger C, Hinton J. A window into lysogeny: Revealing temperate phage biology with transcriptomics. Microb Genomics. 2020.
  103. 103. Hemsley A, Arnheim N, Toney M D, Cortopassi G, Galas D J. A simple method for site-directed mutagenesis using the polymerase chain reaction. Nucleic Acids Res. 1989;17: 6545–6551. pmid:2674899
  104. 104. Koskiniemi S, Pränting M, Gullberg E, Näsvall J, Andersson D I. Activation of cryptic aminoglycoside resistance in Salmonella enterica. Mol Microbiol. 2011. pmid:21507083
  105. 105. Hammarlöf D L, Kröger C, Owen S V., Canals R, Lacharme-Lora L, Wenner N, et al. Role of a single noncoding nucleotide in the evolution of an epidemic African clade of Salmonella. Proc Natl Acad Sci U S A. 2018. pmid:29487214
  106. 106. Datsenko K A, Wanner B L. One-step inactivation of chromosomal genes in Escherichia coli K-12 using PCR products. Proc Natl Acad Sci U S A. 2000. pmid:10829079
  107. 107. Owen S V, Wenner N, Canals R, Makumi A, Hammarlöf D L, Gordon M A, et al. Characterization of the prophage repertoire of African Salmonella Typhimurium ST313 reveals high levels of spontaneous induction of novel phage BTP1. Front Microbiol. 2017. pmid:28280485
  108. 108. Schmieger H. Phage P22-mutants with increased or decreased transduction abilities. MGG Mol Gen Genet. 1972. pmid:4564719
  109. 109. Cherepanov P P, Wackernagel W. Gene disruption in Escherichia coli: TcR and KmR cassettes with the option of Flp-catalyzed excision of the antibiotic-resistance determinant. Gene. 1995. pmid:7789817
  110. 110. Schulte M, Sterzenbach T, Miskiewicz K, Elpers L, Hensel M, Hansmeier N. A versatile remote control system for functional expression of bacterial virulence genes based on the tetA promoter. Int J Med Microbiol. 2019. pmid:30501934
  111. 111. Kröger C, Colgan A, Srikumar S, Händler K, Sivasankaran S K, Hammarlöf D L, et al. An infection-relevant transcriptomic compendium for Salmonella enterica serovar Typhimurium. Cell Host Microbe. 2013. pmid:24331466
  112. 112. Blank K, Hensel M, Gerlach R G. Rapid and highly efficient method for scarless mutagenesis within the Salmonella enterica chromosome. PLoS One. 2011. pmid:21264289
  113. 113. Canals R, Hammarlöf D L, Kröger C, Owen S V., Fong W Y, Lacharme-Lora L, et al. Adding function to the genome of African Salmonella Typhimurium ST313 strain D23580. PLoS Biol. 2019. pmid:30645593
  114. 114. Zojer M, Schuster L N, Schulz F, Pfundner A, Horn M, Rattei T. Variant profiling of evolving prokaryotic populations. PeerJ. 2017. pmid:28224054
  115. 115. Notley-McRobb L, King T, Ferenci T. rpoS mutations and loss of general stress resistance in Escherichia coli populations as a consequence of conflict between competing stress responses. J Bacteriol. 2002.
  116. 116. Ashton P M, Owen S V, Kaindama L, Rowe W P M, Lane C R, Larkin L, et al. Public health surveillance in the UK revolutionises our understanding of the invasive Salmonella Typhimurium epidemic in Africa. Genome Med. 2017. pmid:29084588
  117. 117. O’Toole G A, Kolter R. Initiation of biofilm formation in Pseudomonas fluorescens WCS365 proceeds via multiple, convergent signalling pathways: A genetic analysis. Mol Microbiol. 1998. pmid:9632250
  118. 118. Bongaerts R J M, Hautefort I, Sidebotham J M, Hinton J C D. Green fluorescent protein as a marker for conditional gene expression in bacterial cells. Methods Enzymol. 2002. pmid:12474378
  119. 119. Zuker M. Mfold web server for nucleic acid folding and hybridization prediction. Nucleic Acids Res. 2003. pmid:12824337
  120. 120. Mann M, Wright P R, Backofen R. IntaRNA 2.0: Enhanced and customizable prediction of RNA-RNA interactions. Nucleic Acids Res. 2017. pmid:28472523
  121. 121. Abo-Amer A E, Munn J, Jackson K, Aktas M, Golby P, Kelly D J, et al. DNA Interaction and Phosphotransfer of the C4-Dicarboxylate-Responsive DcuS-DcuR Two-Component Regulatory System from Escherichia coli. J Bacteriol. 2004.
  122. 122. Sedlyarova N, Shamovsky I, Bharati B K, Epshtein V, Chen J, Gottesman S, et al. sRNA-Mediated Control of Transcription Termination in E. coli. Cell. 2016. pmid:27662085
  123. 123. Soper T, Mandin P, Majdalani N, Gottesman S, Woodson S A. Positive regulation by small RNAs and the role of Hfq. Proc Natl Acad Sci U S A. 2010. pmid:20457943
  124. 124. Bandyra K J, Sinha D, Syrjanen J, Luisi B F, De Lay N R. The ribonuclease polynucleotide phosphorylase can interact with small regulatory RNAs in both protective and degradative modes. RNA. 2016. pmid:26759452
  125. 125. De Lay N, Gottesman S. Role of polynucleotide phosphorylase in sRNA function in Escherichia coli. RNA. 2011. pmid:21527671
  126. 126. Massé E, Escorcia F E, Gottesman S. Coupled degradation of a small regulatory RNA and its mRNA targets in Escherichia coli. Genes Dev. 2003. pmid:12975324
  127. 127. Desnoyers G, Massé E. Noncanonical repression of translation initiation through small RNA recruitment of the RNA chaperone Hfq. Genes Dev. 2012. pmid:22474262
  128. 128. Nurmohamed S, Vincent H A, Titman C M, Chandran V, Pears M R, Du D, et al. Polynucleotide phosphorylase activity may be modulated by metabolites in Escherichia coli. J Biol Chem. 2011. pmid:21324911
  129. 129. Witan J, Monzel C, Scheu P D, Unden G. The sensor kinase DcuS of Escherichia coli: Two stimulus input sites and a merged signal pathway in the DctA/DcuS sensor unit. Biological Chemistry. 2012. pmid:23109544
  130. 130. McClelland M, Sanderson K E, Spieth J, Clifton S W, Latreille P, Courtney L, et al. Complete genome sequence of Salmonella enterica serovar Typhimurium LT2. Nature. 2001. pmid:11677609