Skip to main content
Advertisement
  • Loading metrics

Signal transduction pathway mediated by the novel regulator LoiA for low oxygen tension induced Salmonella Typhimurium invasion

  • Lingyan Jiang ,

    Contributed equally to this work with: Lingyan Jiang, Lu Feng

    Affiliations TEDA Institute of Biological Sciences and Biotechnology, Nankai University, Tianjin, P. R. China, College of Environmental Science and Engineering, Nankai University, Tianjin, P. R. China

  • Lu Feng ,

    Contributed equally to this work with: Lingyan Jiang, Lu Feng

    Affiliations TEDA Institute of Biological Sciences and Biotechnology, Nankai University, Tianjin, P. R. China, The Key Laboratory of Molecular Microbiology and Technology, Ministry of Education, Tianjin, P. R. China, Tianjin Key Laboratory of Microbial Functional Genomics, Tianjin, P. R. China

  • Bin Yang,

    Affiliations TEDA Institute of Biological Sciences and Biotechnology, Nankai University, Tianjin, P. R. China, The Key Laboratory of Molecular Microbiology and Technology, Ministry of Education, Tianjin, P. R. China, Tianjin Key Laboratory of Microbial Functional Genomics, Tianjin, P. R. China

  • Wenwen Zhang,

    Affiliations TEDA Institute of Biological Sciences and Biotechnology, Nankai University, Tianjin, P. R. China, The Key Laboratory of Molecular Microbiology and Technology, Ministry of Education, Tianjin, P. R. China, Tianjin Key Laboratory of Microbial Functional Genomics, Tianjin, P. R. China

  • Peisheng Wang,

    Affiliations TEDA Institute of Biological Sciences and Biotechnology, Nankai University, Tianjin, P. R. China, The Key Laboratory of Molecular Microbiology and Technology, Ministry of Education, Tianjin, P. R. China, Tianjin Key Laboratory of Microbial Functional Genomics, Tianjin, P. R. China

  • Xiaohan Jiang,

    Affiliations TEDA Institute of Biological Sciences and Biotechnology, Nankai University, Tianjin, P. R. China, The Key Laboratory of Molecular Microbiology and Technology, Ministry of Education, Tianjin, P. R. China, Tianjin Key Laboratory of Microbial Functional Genomics, Tianjin, P. R. China

  • Lei Wang

    wanglei@nankai.edu.cn

    Affiliations TEDA Institute of Biological Sciences and Biotechnology, Nankai University, Tianjin, P. R. China, The Key Laboratory of Molecular Microbiology and Technology, Ministry of Education, Tianjin, P. R. China, Tianjin Key Laboratory of Microbial Functional Genomics, Tianjin, P. R. China

Correction

12 Aug 2019: Jiang L, Feng L, Yang B, Zhang W, Wang P, et al. (2019) Correction: Signal transduction pathway mediated by the novel regulator LoiA for low oxygen tension induced Salmonella Typhimurium invasion. PLOS Pathogens 15(8): e1007997. https://doi.org/10.1371/journal.ppat.1007997 View correction

Abstract

Salmonella enterica serovar Typhimurium (S. Typhimurium) is a major intestinal pathogen of both humans and animals. Salmonella pathogenicity island 1 (SPI-1)-encoded virulence genes are required for S. Typhimurium invasion. While oxygen (O2) limitation is an important signal for SPI-1 induction under host conditions, how the signal is received and integrated to the central SPI-1 regulatory system in S. Typhimurium is not clear. Here, we report a signal transduction pathway that activates SPI-1 expression in response to low O2. A novel regulator encoded within SPI-14 (STM14_1008), named LoiA (low oxygen induced factor A), directly binds to the promoter and activates transcription of hilD, leading to the activation of hilA (the master activator of SPI-1). Deletion of loiA significantly decreased the transcription of hilA, hilD and other representative SPI-1 genes (sipB, spaO, invH, prgH and invF) under low O2 conditions. The response of LoiA to the low O2 signal is mediated by the ArcB/ArcA two-component system. Deletion of either arcA or arcB significantly decreased transcription of loiA under low O2 conditions. We also confirmed that SPI-14 contributes to S. Typhimurium virulence by affecting invasion, and that loiA is the virulence determinant of SPI-14. Mice infection assays showed that S. Typhimurium virulence was severely attenuated by deletion of either the entire SPI-14 region or the single loiA gene after oral infection, while the virulence was not affected by either deletion after intraperitoneal infection. The signal transduction pathway described represents an important mechanism for S. Typhimurium to sense and respond to low O2 conditions of the host intestinal tract for invasion. SPI-14-encoded loiA is an essential element of this pathway that integrates the low O2 signal into the SPI-1 regulatory system. Acquisition of SPI-14 is therefore crucial for the evolution of S. Typhimurium as an intestinal pathogen.

Author summary

Salmonella enterica serovar Typhimurium (S. Typhimurium) is a major intestinal pathogen of both humans and animals. Salmonella pathogenicity island 1 (SPI-1) is required for host cell invasion by S. Typhimurium. Expression of SPI-1 genes is induced by low oxygen (O2) tension under host conditions, but the relevant regulatory mechanisms are not clear. Here, we report a low O2-induced signal transduction pathway for the activation of SPI-1 expression in S. Typhimurium. A novel regulator, STM14_1008 (named LoiA), encoded within SPI-14 directly activates hilD, which in turn activates hilA (the master activator of SPI-1), and thus other SPI-1 genes under O2-limited conditions. The response of LoiA to the low O2 signal is mediated by the ArcB/ArcA two-component system. We also confirmed that SPI-14 contributes to S. Typhimurium virulence by affecting invasion, with loiA as the virulence determinant. This novel SPI-1 activation pathway can be used by S. Typhimurium to sense and respond to low O2 conditions of the host intestinal tract for invasion. Acquisition of SPI-14 is therefore very important for the evolution of S. Typhimurium virulence by providing an essential component of this pathway, loiA.

Introduction

Salmonella infects both humans and animals, resulting in a variety of diseases ranging from mild self-limiting gastroenteritis to severe systemic illness depending on serovar/host combination. Salmonella enterica serovar Typhimurium (S. Typhimurium) is a leading cause of human gastroenteritis, but can induce systemic disease in mice that resembles human typhoid fever [1,2]. S. Typhimurium pathogenicity and virulence mechanisms have been extensively studied using murine systems [1,3]. S. Typhimurium infection starts by ingestion of contaminated food or water. After surviving acidic challenge of the host stomach, S. Typhimurium enters the small intestine where it manages to invade and penetrate intestinal epithelium, leading to either inflammatory diarrhea confined to the intestinal tract or systemic spread of bacteria that are taken up by macrophages [35].

Salmonella pathogenesis is largely dependent on virulence genes encoded by Salmonella pathogenicity islands (SPIs). 23 SPIs have been identified so far, and 12 are present in S. Typhimurium (SPIs-1 to 6, 9, 11 to 14 and 16) [6,7]. SPI-1 and SPI-2, which encode separate type III secretion systems (T3SS-1 and T3SS-2, respectively) and effectors that are present in all S. enterica serovars, are the most important SPIs required for Salmonella pathogenicity. SPI-1 is required for Salmonella invasion into host intestinal epithelium, whereas SPI-2 is necessary for Salmonella survival and replication within macrophages and progress to systemic infection [1,8,9]. Virulence contributions of other SPIs in S. Typhimurium have also been investigated except for SPI-9 and SPI-14. SPI-3, SPI-6 and SPI-11 to 13 are involved in intracellular survival, contributing to S. Typhimurium systemic infection of mice [1014]; SPI-4 and SPI-5 contribute to enteropathogenicity of S. Typhimurium by affecting bacterial adhesion or invasion, respectively [15,16]; SPI-16 is required for the long-term intestinal persistence of S. Typhimurium in mice [17]. Although not investigated in S. Typhimurium, SPI-9, encoding a type I secretion system, was reported to contribute to the adhesion of S. Typhi to epithelial cells [18], and biofilm formation and invasion in S. Enteritidis [19].

SPI-14 corresponds to a 9-kb region in S. Typhimurium [20]. It consists of eight open reading frames from STM14_1001 to STM14_1008 (S1 Fig), of which, STM14_1008 encodes a putative LysR family transcriptional regulator, STM14_1004 is a pseudogene and the other six encode putative cytoplasmic proteins of unknown function. The contribution of SPI-14 to Salmonella virulence has been investigated in S. Gallinarum and S. Enteritidis [20,21]. By using a PCR-based signature-tagged mutagenesis, two mutants of SPI-14 from S. Gallinarum (SGA-8 and SGC-8, homologs of S. Typhimurium STM14_1002 and STM14_1008, respectively) showed attenuated virulence in chickens [20]. How those two genes contribute to S. Gallinarum virulence was not investigated further. Using a similar approach, a mutant of SPI-14 from S. Enteritidis (SEN0803, homolog to S. Typhimurium STM14_1005) showed reduced invasion of Caco-2 cells and reduced invasiveness to chicken liver cells [21]. Therefore, it is most likely that SPI-14 also contributes to virulence in S. Typhimurium.

Invasion into host intestinal epithelial cells is the first crucial step in S. Typhimurium pathogenesis and is dependent on functions encoded by SPI-1 [1,2]. Effectors encoded within and outside SPI-1, and translocated by the SPI-1-encoded T3SS-1 into the intestinal epithelial cells of the eukaryotic host, are known to induce actin cytoskeletal rearrangements, leading to membrane ruffling and uptake of the invading Salmonella by the epithelial cells [3,22]. Expression of SPI-1 genes is induced when S. Typhimurium reaches the distal ileum of the host intestinal tract, the preferred invasion site, in response to combined environmental signals, including low O2, high osmolarity (salt), near neutral pH, and high acetate and high iron concentrations [2326]. Low O2 tension is an important signal for invasion, as the expression of SPI-1 genes are highly induced under low O2 conditions [23,24,27]. However, regulatory mechanisms for low O2-induced SPI-1 activation are not clear.

A complex regulatory network has been reported for controlling SPI-1 expression. HilA encoded within SPI-1 is the master activator of SPI-1, which activates prg and inv/spa operons of SPI-1 directly and the sip operon of SPI-1 by acting through InvF, another SPI-1-encoded activator [28]. Expression of hilA is directly controlled by three AraC-like transcriptional regulators, including SPI-1-encoded HilC and HilD, and RtsA encoded outside of SPI-1, which constitute a feed-forward regulatory loop [29]. HilD is the dominant regulator of the system, as a hilD mutant had almost no hilA expression, whereas deletion of either hilC or rtsA decreased hilA expression less significantly [29,30]. Many other regulators encoded outside of SPI-1 affect SPI-1 genes and/or HilA expression through HilD, including HilE (repression of hilA by binding to and preventing HilD function) [31], FliZ (activation of hilA via post-translational regulation of HilD) [32], DNA adenine methylase (activation of hilA via post-translational regulation of HilD) [33], EnvZ-OmpR (activation of hilA via HilD) [34], CpxA/R (repression of hilA by decreasing the stability of HilD) [35], BarA/SirA (activation of hilA via activation of csrB/csrC to block CsrA repression of hilD) [24] and Fur (activation of hilA via an unknown regulation of HilD) [26]. In addition, Mlc, PhoP/Q, PhoB/R and FimZ/Y repress expression of SPI-1/HilA by acting through HilE [31,36,37].

The ability of S. Typhimurium to sense and respond to environmental signals at the distal ileum to induce SPI-1 expression is essential for invasion. Of the known regulators, the BarA/SirA two-component system mediates the response to high osmolarity for salt-induced SPI-1 activation while another osmo-regulator system, EnvZ/OmpR, is not involved [38]. SirA, independently of BarA, also activates hilD in response to the high acetate concentration (10–30 mM) found in the distal ileum [24]. As for low O2-induced SPI-1 activation, it was found that expression of rtsA, hilC and hilD genes is induced under low O2 conditions [39], indicating the involvement of those direct SPI-1 regulators. However, how this signal is received and integrated to the central SPI-1 regulatory system is not clear [25,39,40].

In this study, we investigated the role of SPI-14 in S. Typhimurium pathogenicity using the murine model. We found that deletion of SPI-14 significantly reduced invasion of S. Typhimurium into Caco-2 epithelial cells, and the BALB/c mice orally infected with the mutant contained fewer bacteria in the ileum, liver and spleen, and survived better, in comparison with the mice infected with the wild-type strain. Further mutation and complementation analysis revealed that the virulence determinant in SPI-14 is STM14_1008 which we named LoiA (low oxygen induced factor A), encoding a putative LysR family transcriptional regulator. Effects of LoiA on S. Typhimurium invasion were further investigated using in vitro and in vivo experiments. It was found that LoiA positively regulates the expression of hilA (the master activator of SPI-1) through direct activation of HilD under low O2 conditions, leading to the activation of SPI-1 genes for invasion. Finally, the response to low O2 conditions by LoiA was mediated by the ArcB/ArcA two-component regulatory system. Thus this study reports a novel low O2 signal transduction pathway, with SPI-14 encoded LoiA provided an essential element, for the activation of SPI-1 genes to facilitate S. Typhimurium invasion.

Results

SPI-14 deletion attenuated S. Typhimurium virulence in mice by affecting invasion

To investigate the effect of SPI-14 in S. Typhimurium pathogenicity, the SPI-14 mutant of S. Typhimurium strain 14028 (wild-type) was generated. Two groups of 6- to 8-week-old female BALB/c mice were orally infected with approximately 5×106 CFU of wild-type or SPI-14 mutant, respectively, and monitored for survival over a 30-day period. When challenged with the wild-type strain, the BALB/c mice started to die from day 3, and all had died within 16 days. In contrast, the mice infected with the SPI-14 mutant strain started to die from day 9, and only 25% had died after 16 days, with 75% surviving for the duration of the experiment (30 days) (Fig 1A). In addition, mice infected with the SPI-14 mutant had fewer bacteria in ileum (87-fold), livers (234-fold) and spleens (263-fold) than the mice infected with the wild-type strain, 5 days post-infection (Fig 1B). These results demonstrated that the deletion of SPI-14 severely attenuated the virulence of S. Typhimurium in mice after oral infection.

thumbnail
Fig 1. Lack of SPI-14 decreased S. Typhimurium virulence in mice by affecting intestinal invasion.

(A) Survival plots of BALB/c mice for a 30-day period after orally infected with ~5×106 CFU of wild-type or SPI-14 mutant. Data presented are the combination of three independent experiments, ***P<0.001 by log-rank curve comparison test. (B) Bacterial counts recovered from ileum, liver and spleen of the orally infected mice. At day 5 post-infection, mice organs were harvested and homogenized for colony enumeration. Data are combined from three independent experiments. Bars represent mean CFU of all mice, with P value determined by the Mann-Whitney U test (**P<0.01). (C) Survival plots of BALB/c mice after inoculation intraperitoneally (i.p.) with 1×104 CFU of wild-type or SPI-14 mutant. Data presented are the combination of two independent experiments, with P value determined by log-rank curve comparison test (ns, not significant). (D) Bacterial counts recovered from liver and spleen of the i.p. infected mice. At day 3 post-infection, mice organs were harvested and homogenized for colony enumeration. Data are combined from two independent experiments. Bars represent mean CFU of all mice, with P value determined by the Mann-Whitney U test (ns, not significant).

https://doi.org/10.1371/journal.ppat.1006429.g001

To determine whether the attenuated virulence caused by the deletion of SPI-14 in orally infected mice was due to its effect on invasion and/or systemic infection, we repeated mice infection assays by intraperitoneal (i.p.) injection, which allows Salmonella to directly disseminate to the systemic sites via the lymphatic and bloodstream system, bypassing the need for invasion of the intestine as in oral infection [41,42]. In contrast to the severely attenuated virulence of the SPI-14 mutant in orally infected mice, virulence in i.p. infected mice was not affected by deletion of SPI-14, as indicated by similar death rates and similar numbers of bacteria recovered from systemic organs (livers and spleens) between mice infected with the SPI-14 mutant and with the wild-type strain (Fig 1C and 1D). These results indicate that SPI-14 is not involved in systemic infection, and the attenuated virulence of the SPI-14 mutant observed in orally infected mice was likely due to the defect in intestinal invasion.

SPI-14 deletion reduced S. Typhimurium invasion of epithelial cells

How SPI-14 contributes to S. Typhimurium pathogenicity was further investigated by examining the effect of SPI-14 deletion on the ability of S. Typhimurium to adhere to and/or invade Caco-2 epithelial cells, and to replicate in murine RAW264.7 macrophages. As indicated by gentamicin protection assays, SPI-14 mutant and wild-type strains showed similar adherence ability to Caco-2 cells (Fig 2A); however, the invasion ability of the SPI-14 mutant to Caco-2 cells was decreased 4.8-fold compared with the wild-type strain (Fig 2B), indicating that SPI-14 contributes to S. Typhimurium invasion of, but not adherence to, epithelial cells. The replication fold in RAW264.7 macrophages 16 h post-infection was similar for the SPI-14 mutant and wild-type strains (Fig 2C), indicating that SPI-14 does not have a role in intracellular replication of S. Typhimurium within murine macrophages. The effect of SPI-14 deletion on S. Typhimurium invasion was further confirmed by immunofluorescence microscopy examination. By examining at least 50 Caco-2 cells for each strain in random fields, we found 64% of Caco-2 cells infected by the wild-type strain contained a single bacterium and 26% contained two or more bacteria, while 10% of the infected cells did not contain bacteria. In contrast, 70% of the Caco-2 cells infected by the SPI-14 mutant strain did not contain bacteria, while only 28% of cells contained one bacterium and 2% of cells contained two bacteria (S2A Fig). On average, Caco-2 cells infected with the wild-type strain contained 1.42 bacteria, while Caco-2 cells infected with the SPI-14 mutant contained 0.27 bacteria (S2B Fig). The wild-type and SPI-14 mutant exhibited similar growth rates in vitro (S3A and S3B Fig), indicating that the decreased ability of the SPI-14 mutant to invade Caco-2 cells was not due to a growth defect.

thumbnail
Fig 2. Lack of SPI-14 decreased S. Typhimurium invasion of Caco-2 epithelial cells.

(A, B, C) Adhesion (A), invasion (B) and macrophage replication assays (C) of wild-type and SPI-14 mutant. Caco-2 cells or RAW264.7 cells were infected with wild-type S. Typhimurium or SPI-14 mutant at a multiplicity of infection (MOI) of 10. Adhesion and invasion ability of SPI-14 mutant were reported as percentage relative to the wild-type. Replication in RAW264.7 cells was determined by the ratio of the number of intracellular bacteria at 16 h post-infection to the number of bacteria at 2 h post-infection. Data are representative of at least three independent experiments and are presented as mean ± SD; P value was determined by student’s t test (***P<0.001).

https://doi.org/10.1371/journal.ppat.1006429.g002

Collectively, this result demonstrates that deletion of SPI-14 attenuates S. Typhimurium virulence by decreasing bacterial invasion of epithelial cells.

The gene loiA is the virulence determinant in SPI-14

To identify the virulence determinants in SPI-14 affecting S. Typhimurium invasion, mutant strains lacking the left (STM14_1001 to STM14_1004) and right (STM14_1005 to STM14_1008) regions of SPI-14 were generated and tested for invasion abilities into Caco-2 cells. Deletion of STM14_1001 to STM14_1004 had no effect on invasion, while deletion from STM14_1005 to STM14_1008 reduced bacterial invasion to the same level as the deletion of the entire SPI-14 region (Fig 3A). Mutants lacking each of the four genes from STM14_1005 to STM14_1008 were then generated and tested for invasion ability. Mutation in genes STM14_1005, STM14_1006 and STM14_1007 had no effect on invasion, while mutation in gene STM14_1008 (named loiA) significantly reduced bacterial invasion (4.5-fold) (S2C Fig), similar to the level of reduction caused by the deletion of the entire SPI-14 region (4.8-fold) (Fig 3A and 3B). Mice infection assays showed that the survival rates of mice orally infected with the loiA mutant and the SPI-14 mutant were enhanced to similar levels, and the numbers of bacteria recovered from ileum, livers and spleens, respectively, were also reduced to similar levels between the two mutants, in comparison with the values obtained with the wild-type strain (Fig 3C and 3D). Complementation of either mutant by the expression of a functional loiA gene restored the virulence capacity of both loiA mutant and SPI-14 mutant strains to wild-type level (Fig 3C and 3D). Mice i.p. infection assays also found that the loiA mutant did not influence systemic infection (S4A and S4B Fig), in line with the result obtained with the SPI-14 mutant.

thumbnail
Fig 3. The gene loiA (STM14_1008) is the virulence determinant in SPI-14 influencing S. Typhimurium invasion.

(A) Invasion assays of wild-type, SPI-14 mutant, STM14_1001-STM14_1004 mutant, STM14_1005-STM14_1008 mutant, STM14_1005 mutant, STM14_1006 mutant, STM14_1007 mutant and loiA (STM14_1008) mutant. (B) Invasion assays of wild-type, SPI-14 mutant, loiA (STM14_1008) mutant and complemented strains. For (A) and (B), Caco-2 cells were infected with bacteria at an MOI of 10. The invasion ability of mutants is reported as percentages relative to the wild-type strain. Data are representative of at least three independent experiments and are presented as mean ±SD. P values were determined by student’s t test (***P<0.001). (C) Survival plots of BALB/c mice over a 30-day period after orally infected with ~5×106 CFU of indicated bacterial strains. Data presented are the combination of three independent experiments, ***P<0.001 by log-rank curve comparison test. (D) Bacterial counts recovered from ileum, liver and spleen of the orally infected mice. At day 5 post-infection, mice organs were harvested and homogenized for colony enumeration. Data are combined from three independent experiments. Bars represent mean CFU of all mice, with P value determined by the Mann-Whitney U test (**P<0.01; ns, not significant).

https://doi.org/10.1371/journal.ppat.1006429.g003

Collectively, these results demonstrate that the gene loiA is the virulence determinant in SPI-14 affecting S. Typhimurium invasion into epithelial cells.

loiA deletion reduced the expression of hilA and SPI-1 genes

Considering the essential role of SPI-1 for Salmonella invasion, whether loiA has a role in regulating SPI-1 genes was investigated by quantitative real-time PCR (qRT-PCR) analysis. Expression of the SPI-1 master regulatory gene hilA and five HilA-regulated SPI-1 genes (sipB, spaO, invH, prgH and invF) was tested in the loiA mutant, in comparison with the wild-type strain, grown under SPI-1-inducing conditions (low O2, high salt). Expression of the five HilA-regulated SPI-1 genes was downregulated 4- to 10-fold in the mutant, while expression of hilA was downregulated 2.2-fold, in comparison with the expression of respective genes in the wild-type strain (Fig 4A). Complementation of a functional loiA gene to the loiA mutant restored the expression of all SPI-1 genes tested to wild-type levels (Fig 4A). These results indicate that LoiA functions as a positive regulator of hilA to activate SPI-1 genes.

thumbnail
Fig 4. LoiA contributes to invasion through activating SPI-1.

(A) Wild-type, loiA mutant and complemented strain were grown under SPI-1-inducing conditions (high salt, low O2) to late-exponential phase. Transcript levels of the SPI-1 genes in different strains were assessed by qRT-PCR. 16S rRNA gene was used as the internal control. Data are representative of three independent experiments and are presented as mean ±SD. P values were determined by student’s t test (**P<0.01; *** P<0.001). (B) Bacterial counts recovered from ileum, liver and spleen of the BALB/c mice orally infected with 5×106 CFU of wild-type, loiA mutant, SPI-1 mutant or SPI-1/loiA double mutant at day 5 post-infection. Data are combined from three independent experiments. Bars represent mean CFU of all mice, with P values determined by the Mann-Whitney U test (**P<0.01; ns, not significant).

https://doi.org/10.1371/journal.ppat.1006429.g004

To investigate whether LoiA affects invasion by regulating other virulence genes apart from SPI-1, a SPI-1 mutant (lacking the sit operon, which is required for systemic infection [43]) and a SPI-1/loiA double mutant were generated and used for mice infection assays. Mice orally infected with the SPI-1 mutant and SPI-1/loiA double mutant had similar numbers of bacteria in the organs tested (the ileums, livers and spleens), and the values from both mutant strains were significantly lower than those of loiA mutant infected mice (Fig 4B). These results indicate that the virulence defect of the SPI-1/loiA double mutant is caused by SPI-1 deletion, not loiA deletion, and thus the effect of loiA on virulence is due to its effect on SPI-1. In addition, through i.p infection of mice, we found that loiA did not confer an additional virulence defect by influencing systemic infection in a SPI-1 mutant background, since mice i.p. infected with the SPI-1 mutant or SPI-1/loiA double mutant also resulted in similar bacterial burdens in the livers and spleens (S5A and S5B Fig). Taken together, these findings confirmed that loiA affects S. Typhimurium virulence by affecting invasion via SPI-1.

LoiA activates expression of hilA gene through activating hilD

As the expression of hilA is directly controlled by HilD, HilC and RtsA, we tested whether LoiA regulates HilA through any of these three regulators. Using qRT-PCR analysis, hilD gene expression was significantly reduced in the loiA mutant compared with the wild-type strain in vitro under SPI-1-inducing conditions (2.4-fold) (Fig 5A), while the expression of hilC and rtsA in the LoiA mutant were not significantly affected (Fig 5A). These results indicate that LoiA regulates hilD positively, but has no direct effect on hilC or rtsA.

thumbnail
Fig 5. LoiA regulates hilA and other SPI-1 genes through HilD.

(A) qRT-PCR analysis of changes in the expression of three direct higher regulators of hilA (hilD, hilC, and rtsA) in wild-type, ΔloiA and ΔloiA+pLoiA. Strains were grown under SPI-1-inducing conditions. (B) qRT-PCR analysis of changes in hilA expression in wild-type, ΔloiA, ΔhilD, ΔloiAΔhilD, ΔloiAΔhilD+pHilD and ΔloiAΔhilD+pLoiA. Strains were grown under SPI-1-inducing conditions. (C) Western blot analysis of HilA in wild-type, ΔloiA, ΔhilD, ΔloiAΔhilD, ΔloiAΔhilD+pHilD and ΔloiAΔhilD+pLoiA. Strains were grown under SPI-1-inducing conditions. Expression of the tagged HilA protein was determined using that of DnaK as the internal control. (D) Invasion assays of wild-type, ΔloiA, ΔhilD, ΔloiAΔhilD, ΔloiAΔhilD+pHilD and ΔloiAΔhilD+pLoiA. Caco-2 cells were infected with bacteria at an MOI of 10. The invasion ability of mutants is reported as percentages relative to the wild-type strain. (E) EMSAs of hilD promoter DNA fragment with purified LoiA-His6 protein (0, 0.1, 0.2, 0.4, 0.8 and 1.6 μM). hilC promoter is used as a negative control. (F) Fold enrichment of the hilD promoter in ChIP samples, as measured via ChIP-qPCR. hilC promoter is used as negative control. ΔloiA, loiA mutant; ΔloiA +pLoiA, ΔloiA complemented strain; ΔhilD, hilD mutant; ΔloiAΔhilD, loiA/hilD double mutant; ΔloiAΔhilD+pHilD, loiA/hilD double mutant complemented with HilD; ΔloiAΔhilD+pLoiA, loiA/hilD double mutant complemented with LoiA. Data are representative of at least three independent experiments and are presented as mean ±SD. P values were determined by student’s t test (**P<0.01; *** P<0.001).

https://doi.org/10.1371/journal.ppat.1006429.g005

To further confirm that LoiA regulates hilA through hilD, a hilD mutant strain △hilD, the loiA and hilD double mutant strain ΔloiAΔhilD, and the corresponding complemented strains ΔloiAΔhilD +pHilD and ΔloiAΔhilD +pLoiA, were constructed. Expression of hilA in the loiA/hilD double mutant was equivalent to that in the hilD mutant, with both mutants showing much lower levels of hilA expression than the reduced level in the loiA mutant, as evidenced by qRT-PCR analysis at the transcription level and western blotting at the protein level (Fig 5B and 5C), indicating that the presence of HilD is necessary for the activity of LoiA. The stronger reduction of hilA expression by the deletion of hilD is in line with the fact that hilD is regulated by many other regulators apart from LoiA. Complementation of the loiA/hilD double mutant with a functional hilD gene restored hilA expression, while complementation with a functional loiA gene could not restore hilA expression (Fig 5B and 5C). These results indicate that LoiA activates hilA via activating hilD. As expected, in invasion assays the loiA/hilD double mutant and the hilD mutant strains showed reduced invasion at similar levels, and complementation of the loiA/hilD double mutant with a functional hilD gene restored the invasion ability of this double mutant, while complementation of the double mutant with a functional loiA gene could not restore its invasion ability (Fig 5D). This indicates that LoiA activates invasion through hilD.

Electrophoretic mobility shift assays (EMSAs) were performed with purified LoiA-His6 protein to test whether LoiA directly binds the hilD promoter. As shown in Fig 5E, with increasing concentrations of LoiA protein, slowly migrating bands were observed for the hilD promoter, while no retarded bands were observed for hilC promoter (negative control), which indicated that LoiA can bind to the hilD promoter in vitro. Chromatin immunoprecipitation-quantitative PCR (ChIP-qPCR) analysis further demonstrated the binding of LoiA with the hilD promoter. The hilD promoter was exceedingly enriched in LoiA-ChIP samples, and the relative quantity was 3.8-fold higher than in the untagged control sample, while the control hilC promoter was not enriched in the LoiA-ChIP sample (Fig 5F).

Collectively, these results demonstrate that LoiA directly binds to the hilD promoter to activate hilD expression; and HilD in turn activates hilA expression, leading to the activation of SPI-1 genes.

LoiA expression is activated under low O2 conditions

Signals for LoiA-induced SPI-1 activation were investigated. qRT-PCR analysis showed that loiA gene expression was 7.2-fold higher under the in vitro SPI-1-inducing conditions than under non-inducing conditions (low salt, high O2) (S6 Fig), indicating LoiA responds to high osmolarity or low O2 or both to regulate SPI-1 expression via HilD and HilA.

Further examinations revealed that loiA gene expression was 6.8-fold higher under low O2 and low salt conditions than under high O2 and low salt conditions in LB medium (0.17M NaCl) as indicated by qRT-PCR (Fig 6A) and western blot analysis (Fig 6B). In contrast, expression levels of LoiA were similar under both high salt (0.3M NaCl) and low salt (0.17M NaCl) conditions in the presence of a high level of O2 (Fig 6A and 6B). Thus, these results demonstrate that LoiA expression is activated by low O2 conditions, but not high osmolarity (high salt) conditions.

thumbnail
Fig 6. Expression of LoiA was induced by low O2 concentration.

(A) qRT-PCR analysis of loiA gene expression under high O2 or low O2 (LB, low salt), and high salt (0.3M NaCl) or low salt (0.17M NaCl) conditions (LB, high O2). Data are representative of at least three independent experiments and are presented as mean ±SD. P values were determined by student’s t test (**P<0.01; ns, not significant). (B) Western blot analysis of LoiA protein level under high O2 or low O2 (LB, low salt), and high salt or low salt conditions (LB, high O2). The expression of the tagged LoiA protein was determined using that of DnaK as the internal control.

https://doi.org/10.1371/journal.ppat.1006429.g006

Activation of LoiA expression in response to low O2 conditions is mediated by the ArcB/ArcA two-component regulatory system

Fnr (fumarate nitrate reduction regulator) and the ArcB/ArcA (aerobic respiratory control) two-component regulation system are well-known global regulators, which can sense and respond to changes in O2 availability and regulate gene expression in low O2 and anaerobic conditions. Whether LoiA responds to low O2 conditions through either of these two systems was investigated.

Mutant strains for fnr, arcA encoding the cognate response regulator and arcB encoding the membrane-bound sensor kinase were constructed, and expression levels of loiA in mutant strains were compared by qRT-PCR with that of the wild-type strain. The fnr mutant did not affect loiA gene expression in both low O2 and high O2 conditions (S7A and S7B Fig). Mutation of arcA or arcB significantly reduced loiA gene expression under low O2 conditions. Complementation of arcA and arcB mutants with the corresponding functional genes restored loiA gene expression to wild-type level (Fig 7A). Positive regulation of LoiA by the ArcB/ArcA system is dependent on low O2 conditions as the expression level of loiA in all mutant strains is comparable with the wild-type level under high O2 conditions (Fig 7B). These results showed that LoiA is able to sense and respond to low O2 via the ArcB/ArcA system for SPI-1 activation.

thumbnail
Fig 7. The activation of LoiA gene function by low O2 conditions is mediated by ArcAB.

(A, B) qRT-PCR analysis of loiA gene expression in wild-type, arcA mutant, arcB mutant, and complemented strains for ArcA and ArcB. Bacteria were grown in LB medium (0.17 M NaCl) either with low O2 (A) or high O2 (B). (C, D) qRT-PCR analysis of hilD gene expression in wild-type, arcA mutant, arcB mutant, loiA mutant, arcA/loiA double mutant, arcB/loiA double mutant or complemented strains. Bacteria were grown in LB medium (0.17 M NaCl) either with low O2 (C) or high O2 (D). Data from graphs (A) to (D) are representative of at least three independent experiments and are presented as mean ±SD. P values were determined by student’s t test (*P<0.05; **P<0.01). (E) Bacterial counts recovered from ileum of the BALB/c mice orally infected with 5×106 CFU of wild-type, loiA mutant, arcA mutant or arcA/loiA double mutant at day 5 post-infection. Data are combined from three independent experiments. Bars represent mean CFU of all mice, with significance determined by the Mann-Whitney U test (*P<0.05; **P<0.01; ns, not significant).

https://doi.org/10.1371/journal.ppat.1006429.g007

The connection between the ArcB/ArcA system and LoiA on SPI-1 regulation was further investigated by examing the expression levels of hilD in arcA, arcB, loiA, arcA/loiA and arcB/loiA double mutants grown under high and low O2 conditions. Expression of hilD was significantly reduced in all of the mutants when grown under low O2 conditions, in comparison to the wild-type strain (Fig 7C). However, the loiA, arcA/loiA and arcB/loiA double mutants showed similar levels of reduced hilD expression (2.8-, 2.9- and 2.7-fold, respectively), which were lower than the values of the arcA and arcB mutants (2.1- and 2.1-fold, respectively) (Fig 7C). Complementation of the arcA/loiA double mutant with LoiA restored hilD expression to the level of the arcA mutant, while there was no change in hilD expression level when complemented with arcA (Fig 7C). In addition, no significant reductions in hilD expression were detected in any of the tested mutants compared to the wild-type strain when bacteria were grown under high O2 conditions (Fig 7D). These results indicate that both ArcA and ArcB are involved in the regulation of hilD in response to low O2 conditions, and the ArcAB regulation on hilD is mediated by LoiA. Activation of the ArcAB system through LoiA during invasion was also confirmed by comparing the invasion phenotypes of the loiA mutant, the arcA mutant and arcA/loiA double mutant through orally infected mice. The bacterial burden in the ileum of arcA/loiA double mutant infected mice was equivalent to that of loiA mutant infected mice, and both were lower than that of arcA mutant infected mice (Fig 7E).

In addition, ArcA (the regulator gene of the ArcAB system) binding to the loiA promoter under low oxygen conditions was tested by EMSAs, with the ArcA-dependent cydA promoter used as a positive control [44,45]. Retardation of the cydA promoter (positive control) was clearly observed as the ArcA protein concentration increased, while no retardation could be detected of the loiA promoter (S8 Fig). This suggested that the positive regulation of ArcA on the loiA gene is indirect, with an unknown intermediate regulator.

Discussion

In this study, we characterize and report a signal transduction regulatory pathway that S. Typhimurium uses to sense low O2 conditions of the host intestinal tract to activate expression of SPI-1 genes for invasion. A model for this regulatory pathway is proposed (Fig 8). Briefly, ArcB responds to low oxygen conditions when bacteria reach the distal ileum and undergoes autophosphorylation, following which the phosphate group is transferred to ArcA; phosphorylated ArcA (ArcA-P) activates loiA gene expression indirectly through unknown regulator(s); LoiA activates HilD directly through binding the hilD promoter; HilD then activates hilA and other SPI-1 genes, and thus facilitates the invasion process. Both SPI-1 expression and intestinal invasion ability was severely affected when the pathway was blocked by the deletion of loiA, indicating that the ability to sense and respond to low O2 using this pathway is very important for S. Typhimurium invasion, further confirming the signalling role of low O2 for SPI-1 induction. Clearly, LoiA plays an essential role by receiving the low O2 signal and passes it to the central SPI-1 regulation system (HilD). The production of HilD was suggested as an integration point for particular environmental signals and regulatory elements that then switches to turn on/off SPI-1 [46], and this also applied to the O2-responsive SPI-1 regulatory system described here. HilD, HilC and RtsA are known to form a feed-forward loop for SPI-1 regulation [29]. Therefore, regulation of one of those regulators is expected to have effects on the other two. However, significant reduction of hilD expression (2.4-fold) in a loiA mutant seemed to have little effect on hilC and rtsA expression (1.1- and 1.3-fold reduction, respectively). The lack of feed-forward effects of those three regulators were also reported previously [36,47]. While this cannot be explained clearly, one possibility is that HilD had a major effect on the regulation of downstream SPI-1 genes.

thumbnail
Fig 8. Model of the LoiA regulatory network in S. Typhimurium invasion.

Under low O2 concentrations, the phosphorylated ArcA after sensing of low O2 signal by ArcB can indirectly activate loiA gene transcription, LoiA can then activate SPI-1 T3SS through directly activating HilD, thus promoting epithelial cell invasion.

https://doi.org/10.1371/journal.ppat.1006429.g008

In addition to low O2 signal, the contributions of other site-specific signals or factors are necessary for S. Typhimurium to invade at the preferred site (ileum), as low O2 is a general condition of intestinal tracts of hosts. Several ileum-specific signals for SPI-1 activation have been reported including being rich in acetate and iron, near central pH, and low in concentration of propionate, butyrate and long-chain fatty acids, which can facilitate invasion at this site [23,24,26,48]. On the other hand, SPI-1 expression is repressed by the acidic pH in the stomach [27,49], by the high bile concentration in the proximal small intestine [50,51], and by the high concentration of long-chain fatty acids in the upper and middle sections of the small intestine [48], preventing invasion by S. Typhimurium at those sites prior to reaching the distal ileum. The concentration of those molecules decreased by absorption along the long length of the small intestine, and thus in the distal small intestine, the repression of SPI-1 may be relieved. In case of passing the preferred site of invasion, upon entering the large intestine, SPI-1 expression would be shut down by the increased concentration of propionate and butyrate synthesized by the resident microbiota [24,47]. Those hypotheses need to be confirmed experimentally in the future.

The ArcB/ArcA system is a well-known global regulator that controls expression of a large number of genes for growth under O2 limitation conditions, including genes for aerobic respiration and central metabolisms, as well as activation of the gene for the F-pilus [52,53]. Lim et al. also reported that deletion of arcA reduced hilD expression in S. Typhimurium strain SL1344 grown under aerobic conditions to early stationary phase, although no further information was given [54]. In this study, we clearly demonstrated the involvement of this system in a low O2-stimulated virulence regulation pathway. Activation of LoiA by the Arc system is indirect, and is mediated by unknown regulators that are under control of the Arc system, which remains to be investigated. Although the deletion of either the arc gene or loiA resulted in significantly decreased expression of hilD under low O2 conditions, the fact that hilD expression was still detected at lower levels (2.2- to 2.5-fold higher than that under high O2 conditions) (S9 Fig) indicated the presence of other low O2-signalling regulatory system(s) for SPI-1 activation.

Our study showed that Fnr, another well-known global regulator that senses and responds to O2 limitation, is not involved in the low O2-induced LoiA-mediated activation of SPI-1 genes as the expression of loiA was not affected by the deletion of fnr. Fnr has been previously reported to be a negative regulator of hilA in S. Typhimurium, as loss of Fnr caused a 1.5- to 2-fold increase of hilA expression under in vitro SPI-1-inducing conditions (low O2 and high salt) [46,55]. However, whether the repression of hilA by Fnr was induced by low O2 was not clarified in those studies. Furthermore, Fnr was reported not to influence low O2-mediated Salmonella invasion, as a fnr mutant did not differ in the ability to invade Madin-Darby canine kidney epithelial cells compared with wild-type S. Typhimurium, while both the fnr mutant and wild-type exhibited a significant increase in invasiveness when grown in low O2 vs. aerobic conditions [40]. On the other hand, Hassan et al. found that many SPI-1 genes (prgKJIH, iagB, sicA, spaPO, invJICBAEGF) had lower levels of expression in the fnr mutant than in wild-type S. Typhimurium (growing anaerobically) using microarray analysis [56]. Clearly, whether and how Fnr is involved in low O2-responsive SPI-1 regulation in S. Typhimurium need to be further clarified. Fur, the primary iron regulatory protein in Salmonella, can activate SPI-1 through HilD in response to the high concentration of Fe2+ in the small intestine [26]. Considering the low O2 conditions in the small intestine may contribute to the stability of Fe2+, whether low O2 indirectly influences Fur regulation on SPI-1 needs to be further investigated. Other unknown regulators may also be involved in low O2-mediated SPI-1 regulation. Taken together, these data and questions reflect the complicated nature of low O2-induced regulation of Salmonella invasion.

SPIs are regions of the bacterial chromosome harbouring virulence genes that are obtained via horizontal transmission from other bacteria at some point during evolution. Acquiring SPIs are considered to be ‘quantum leaps’ in Salmonella virulence evolution as critical virulence traits of Salmonella are directly linked to SPIs. SPI-1 is present in Salmonella bongori and all subspecies and serotypes of S. enterica analysed to date, and thus SPI-1 is suggested to be a rather ancient acquisition gained at the separation of the genera Escherichia and Salmonella from a common ancestor, and, relatively, the acquisition of other SPIs was later than that of SPI-1 [8]. Sequential acquisition of SPIs is important during S. Typhimurium evolution. While the acquisition of SPI-1 is a key event during the evolution of Salmonella virulence by providing the pathogen the ability for invasion, the acquisition of SPI-14 was crucial for the invasion to be induced under O2-limited conditions, in the host intestinal tract in this case, therefore representing another important event for the evolution of S. Typhimurium as an intestinal pathogen.

Genome analysis revealed that SPI-14 is present in most commonly reported non-typhoidal S. enterica serovars, including S. Typhimurium, S. Enteritidis, S. Gallinarum, S. Pullorum, S. Choleraesuis, and S. Dublin, while it is absent in the human-restricted serovars S. Typhi and S. Paratyphi A (the genome sequences of 104 representative Salmonella enterica strains for this analysis are available at ftp://ftp.ncbi.nlm.nih.gov/genomes/genbank/bacteria/Salmonella_enterica/). This indicates the presence of the LoiA-mediated low O2-responsive regulatory system in other SPI-14-containing serovars, and the absence of this pathway in the latter two serovars. The expression of SPI-1 genes and invasive phenotype of S. Paratyphi A are induced under low O2 conditions, but not under aerobic conditions [57], indicating that the bacterium uses other low O2-responsive regulatory mechanisms for the invasion of human intestinal epithelial cells. The same may apply to S. Typhi. Both await future studies. SGA-8 (homolog to STM14_1002) and SGC-8 (homolog to loiA) encoded by SPI-14 in S. Gallinarum and SEN0803 (homolog to STM14_1005) encoded by SPI-14 in S. Enteritidis were previously reported as virulence-related genes [20,21]. While SGC-8 is expected to play the same role as loiA in S. Gallinarum, homologs of SGA-8 and SEN0803 (STM14_1002 and STM14_1005) were not related to virulence in S. Typhimurium. Due to the undefined role of SGA-8 in virulence [20], STM14_1002 may contribute to virulence of S. Typhimurium in other ways, such as against the acid shock of the stomach, downregulation of inflammation after invasion and production of biofilm in the gall bladder to elicit chronic infections [1]; this needs further investigation. The contrary results for SEN0803 and STM14-1005 may be due to the different cell lines used for invasion assays (Caco-2 vs. LMH) and different hosts for virulence assays (chicken vs. mice), or a possibility of different functions or different regulation of this homologous gene in different serovars.

The novel low-O2 signal transduction pathway reported here constitutes both global regulators (ArcB/ArcA) and a specifically acquired regulator (LoiA), and direct regulators of SPI-1 (HilA, HilD), indicating the complex mechanisms of SPI-1 regulation. This study further enhanced our understanding on how Salmonella utilizes environmental cues to facilitate invasion.

Materials and methods

Ethics statement

All animal experiments were performed in accordance to the standards established in the Guide for the Care and Use of Laboratory Animals published by the Institute of Laboratory Animal Resources of the National Research Council (United States). The animal research procedures were approved by the Institutional Animal Care Committee at Nankai University and Tianjin Institute of Pharmaceutical Research New Drug Evaluation Co. Ltd (IACUC number: 2016032102), Tianjin, China. Every effort was made to minimize animal suffering and to reduce the number of animals used.

Strains and plasmids

Bacterial strains and plasmids used in this study are listed in S1 Table. Oligonucleotides used in this study are listed in S2 Table. S. Typhimurium strain ATCC 14028 was used as wild-type strain throughout this study.

Mutant strains and the 3×FLAG-tagged strains were generated by the λ Red recombinase system as reported previously [58]. Briefly, PCR products for construction of mutants were generated from the chloramphenicol or kanamycin resistance genes of pKD3 or pKD4, respectively, using primers carrying at their 5’ ends 38–40 bp of homology to the regions flanking the start and stop codons of the gene to be deleted. Primers for 3×FLAG-tagged loiA or hilA alleles were designed to amplify the FLAG epitope coding sequence and chloramphenicol resistance gene using plasmid pWSK-FLAG (carrying 3×FLAG sequence and chloramphenicol resistance gene sequence) as the template. The resulting PCR products were electroporated into strain ATCC 14028 carrying the plasmid pKD46 for homologous recombination. The mutants and 3×FLAG-tagged bacteria undergoing homologous recombination were selected by their resistance to chloramphenicol or kanamycin and then verified by PCR amplification and sequencing. When required, the antibiotic resistance cassette was removed by FLP-mediated recombination with introduction of pCP20 plasmid [59].

Complementation mutants were generated by the expression of the corresponding functional genes (cloned from wild-type ATCC 14028) on a low-copy-number plasmid, pWSK129 [60]. To generate plasmids pLoiA, pHilD, pArcA and pArcB, the ORF and the upstream promoter sequence of the corresponding genes were amplified by PCR from genomic DNA of the wild-type strain. The resulting DNA fragment and pWSK129 vector were digested with the corresponding restriction enzymes (XbaI and BamHI for construction of pLoiA and pHilD; BamHI and EcoRI for construction of pArcB and pArcA). After DNA purification, the amplified PCR fragments were ligated into pWSK129 to give the recombinant plasmids and then transformed into corresponding mutant strains to give complemented strains. The pET-LoiA and pET-ArcA plasmids used for purification of the LoiA-His6 and ArcA-His6 proteins were generated by cloning the loiA and arcA gene sequence into the HindIII and BamHI sites downstream of the His-tag element in plasmid pET-28a, respectively. All the resulting clones were verified by DNA sequencing.

Bacterial growth conditions and cell culture

Bacteria were routinely grown in Luria–Bertani (LB) medium containing 1% tryptone, 0.5% yeast extract and 0.17 M NaCl at 37°C except for strains containing the temperature-sensitive plasmids pKD46 or pCP20, which were grown at 30°C. Under SPI-1-inducing conditions, bacteria were grown in 5 ml high salt LB medium (0.3 M NaCl); incubation was carried out in tightly closed 15 ml Falcon tubes without shaking (low O2). Under SPI-1 non-inducing conditions, bacteria were grown in 5 ml low salt LB medium (0.17 M NaCl); incubation was carried out in a 13 mm test tube with a loose cap with shaking at 200 rpm (high O2) [29,36,39]. Antibiotics were used at the following concentrations: ampicillin (Ap) 100 μg/ml, kanamycin (Km) 50 μg/ml, chloramphenicol (Cm) 20 μg/ml, streptomycin (Sm) 200 μg/ml and gentamicin (Gm) 10 or 100 μg/ml.

The human colon adenocarcinoma (Caco-2) cell line and murine macrophage cell line RAW264.7 were purchased from the Shanghai Institute of Biochemistry and Cell Biology of the Chinese Academy of Sciences (Shanghai, China). Cells were grown in RPMI-1640 medium (Gibco) supplemented with 10% fetal bovine serum (FBS, Gibco) and incubated at 37°C in 5% CO2. At 48 h before infection, cells were seeded into 12-well tissue culture plates, with or without coverslips, at the concentration of 1×105 cells per well and maintained as differentiated monolayers.

Mice infection

Laboratory animals Female BALB/c mice (6–8 weeks old) were purchased from Beijing Vital River Laboratory Animal Technology Co. Ltd (Beijing, China). All mice were maintained in a specific pathogen-free environment. Mice infections were performed as previously described [29,42]. Overnight bacterial cultures were 1:100 diluted and cultured under SPI-1-inducing conditions (high salt, low O2) to an OD600 of 0.6 (late-exponential phase). The collected bacteria were serially diluted to either 5×107 CFU/ml in PBS (for oral infection) or 1×105 CFU/ml in 0.9% NaCl (for i.p. infection). For oral infections, mice were given a single dose of 20 mg streptomycin 24 h prior to infection, followed by oral gavage with 5×106 CFU of indicated S. Typhimurium strains in 0.1 ml PBS. For systemic infections, mice were given 1×104 CFU of indicated S. Typhimurium strains in 0.1 ml 0.9% NaCl by i.p. injection.

The infected mice were monitored daily and the survival rates were recorded. For CFU enumeration experiments, the infected mice were euthanized 5 days post-infection (oral infection) or 3 days post-infection (i.p.). Ileum, spleen and liver were harvested, homogenized in PBS, diluted and then plated on LB plates for determination of CFU. Experiments were repeated two or three times.

Adhesion and invasion assays

Adhesion and invasion assays were performed in the Caco-2 cell line as described previously with some modifications to the protocol [40,61]. Briefly, overnight bacteria were 1:100 sub-cultured into fresh LB medium (0.3 M NaCl) and incubated in tightly closed Falcon tubes without shaking until an OD600 of 0.6. Bacteria were pelleted, washed twice with PBS and re-suspended in RPMI medium. For the adhesion assays, both 12-well cell plates and bacterial suspensions were placed on ice for 15 min before infection. Ice-cold bacterial cells (500 μl per well) were added to Caco-2 cell monolayers at a multiplicity of infection (MOI) of 10 and incubated for 30 min on ice. For the invasion assays, bacterial cells were added to the Caco-2 cell monolayers at a MOI of 10 and incubated for 60 min at 37°C and 5% CO2. After washing three times with PBS, 500 μl RPMI medium containing 100 μg/ml gentamicin was added to the infected cells and incubated for an additional 60 min to kill extracellular bacteria. Following incubation, cells were washed three times with PBS, lysed with 500 μl 0.1% Triton X-100, and suitable dilutions were plated on LB agar plates containing appropriate antibiotics. Adhesion and invasion were calculated as percentages of the number of bacteria recovered from the total bacteria inoculated. All assays were performed with at least three independent biological replicates.

Macrophage replication assays

The macrophage replication assays were conducted with stationary phase bacterial cells as reported previously [62]. Overnight bacteria were 1:100 sub-cultured into fresh LB medium (0.17 M NaCl) until an OD600 of 2.0. Bacteria were pelleted and opsonized in 10% normal mice serum for 20 min and then added to macrophage cell monolayers at an MOI of 10. The plates were centrifuged at 1,000g for 5 min to synchronize the infection. After incubation for 30 min at 37°C and 5% CO2, the infected cells were washed three times with PBS and incubated with medium containing 100 μg/ml gentamicin for 1 h, followed by medium containing 10 μg/ml gentamicin for the remaining time of infection. At 2 h and 16 h post-infection, the supernatant was removed and cells were washed three times with PBS and lysed with 0.1% Triton X-100. Serial dilutions of the lysates were plated onto LB agar to enumerate intracellular bacteria. The fold intracellular replication was calculated by dividing the intracellular bacterial load at 16 h by the bacterial load at 2 h. At least three independent biological replicates were performed.

Immunofluorescence microscopy

For immunofluorescence microscopy, Caco-2 cells were seeded onto glass coverslips and infected as described above. After infection, cells were fixed with 3% paraformaldehyde (PFA) for 15 min at room temperature and washed three times with PBS. Cells were permeabilized for 20 min in 0.1% Triton X-100 and blocked with 5% BSA in PBS for 30 min. Mouse anti-Salmonella LPS (Abcam) was diluted 1:100 in PBS and applied for 1 h. Cells were washed three times with PBS, and then the secondary antibody goat anti-mouse IgG (FITC) (Abcam), diluted 1:200 in PBS, was applied for 1 h. Cells were washed again with PBS and incubated with DAPI (Invitrogen) for 2 min. After washing with PBS, cells were overlaid with 200 μl mounting medium. The cells were inspected for intracellular bacteria using a fluorescence microscope (Olympus) or a confocal laser scanning microscope (Leica) using filter sets for FITC (510 nm excitation, 530 nm emission). Images were further processed using the Leica TCS software package and Adobe Photoshop CS3.

qRT-PCR

qRT-PCR was performed using the 7500 Real-Time PCR system (Applied Biosystems). To test the influence of loiA on expression of SPI-1 genes, the loiA mutant, complemented strain and wild-type strain were grown under SPI-1-inducing conditions for 4 h to late-exponential phase (OD600 ~0.6). To test changes in hilA expression in wild-type, ΔloiA, ΔhilD, ΔloiAΔhilD, ΔloiAΔhilD+pHilD and ΔloiAΔhilD+pLoiA, strains were also grown under SPI-1-inducing conditions for 4 h to late-exponential phase (OD600 ~0.6). To test the influence of osmolarity and O2 level on loiA gene expression, the wild-type strain was grown under low O2, high O2, low osmolarity or high osmolarity conditions. For growth under high O2 and low O2 conditions, overnight bacterial cultures were 1:100 sub-cultured into fresh LB broth and then divided into two groups. One group was shaken at 200 rpm for an additional 4 h to late-exponential phase (OD600 ~1.2) with good aeration (high O2, control); the other group was transferred to a tightly closed Falcon tube and incubated standing for an additional 4 h to late-exponential phase (OD600 ~0.6; low O2). For growth under low osmolarity and high osmolarity conditions, overnight bacterial cultures were 1:100 sub-cultured into LB broth with 0.17 M NaCl (low salt) or 0.3 M NaCl (high salt) and shaken at 200 rpm for an additional 4 h. To test the influence of the fnr mutant and arcA/arcB mutant on loiA gene expression and the influence of the arcA or arcB mutants, and arcA/loiA or arcB/loiA double mutants on hilD gene expression, overnight bacterial cultures were sub-cultured 1:100 into fresh low-salt LB medium (0.17 M NaCl) and grown under high O2 or low O2 conditions for an additional 4 h to late-exponential phase (high O2: OD600 ~1.2; low O2: OD600 ~0.6).

Bacteria were pelleted by centrifugation, RNA samples were isolated using Trizol (Invitrogen), purified by the RNeasy Mini Kit (QIAGEN), DNase I treated (QIAGEN), reverse transcribed using random hexamers (Sigma) and processed for qRT-PCR. Each qRT-PCR reaction was carried out in a total volume of 20 μl in a 96-well optical reaction plate (Applied Biosystems) containing 10 μl FastStart Universal SYBR Green Master (ROX) mix, 1 μl cDNA, and two gene-specific primers with a final concentration of 0.3 mM each. The fold change in target gene relative to the housekeeping gene (16S rRNA) was determined by the 2-ΔΔCt method. At least three biological replicates were performed for each qRT-PCR analysis.

Expression and purification of LoiA-His6 and ArcA-his6

LoiA-His6 and ArcA-His6 fusion proteins were expressed in Escherichia coli BL21 containing pET-LoiA or pET-ArcA and purified from a soluble extract by using a HiTrap Ni2+-chelating column as previously described [63]. Protein concentration was determined by the Bradford procedure. Aliquots of the purified protein were stored at -70°C.

Electrophoretic mobility shift assays (EMSAs)

EMSAs were performed as described previously with some modifications to the protocol [64,65]. PCR fragments encompassing the regulatory regions of hilC, hilD, loiA and cydA were amplified using genomic DNA of S. Typhimurium 14028 as a template. The DNA fragments were gel-purified. LoiA gel shift assays were performed by incubating the purified hilD and hilC promoter fragments (100 ng) at 37°C for 20 min with various concentrations of purified LoiA-His6 protein (0–1.6 μM) in a 20 μl solution containing the band-shift buffer (20 mM Tris-HCl (pH 7.5), 80 mM NaCl, 0.1 mM EDTA and 1 mM DTT). For ArcA O2 limitation gel shift assays, various concentrations of purified ArcA-His6 protein (0–1.6 μM) were incubated under low O2 for 20 min at 37°C with the amplified loiA and cydA promoter fragments (100 ng) in a 20 μl solution containing the above band-shift buffer. Samples were loaded with native binding buffer on a 6% polyacrylamide gel in 0.5× Tris-borate-EDTA. The DNA fragments were stained with ethidium bromide.

ChIP-qPCR

The 3×FLAG-tagged strain (WT loiA-FLAG) was grown under SPI-1-inducing conditions and then pelleted by centrifugation. ChIP was performed based on established methods as reported previously [64,66]. Formaldehyde was added to bacterial cells (1% final concentration) for cross-linking and then incubated at room temperature for 25 min. Reactions were quenched with 0.5 M glycine, and samples were pelleted and washed three times with PBS. The samples were then used for ChIP following the Chromatin Immunoprecipitation kit (Millipore) protocol. The antibody used was the anti-FLAG mouse monoclonal antibody (Sigma). For ChIP-qPCR experiments, untreated chromatin was de-cross-linked by boiling for 10 min and purified for use as the “input” control. The relative enrichment of candidate gene promoters was performed with qRT-PCR and represents the value of the immunoprecipitated DNA divided by the input unprecipitated DNA. These values were normalized to the values obtained for each promoter precipitated using untagged wild-type in order to account for non-specific enrichment. The results represent the mean enrichment measured via qPCR in at least three biological replicate experiments.

Western blot assays

To analyse HilA protein levels in wild-type, ΔloiA, ΔhilD, ΔloiAΔhilD, ΔloiAΔhilD+pHilD and ΔloiAΔhilD+pLoiA strains, the corresponding 3×FLAG-tagged strains were grown in SPI-1-inducing conditions and then collected. To analyse the influence of osmolarity and O2 on the production of LoiA protein, the 3×FLAG-tagged strain (WT loiA-FLAG) was grown under the conditions indicated as described above (high O2 or low O2; high salt or low salt). Bacteria were pelleted by centrifugation, washed with PBS, resuspended in 100 μl SDS-polyacrylamide gel electrophoresis solubilization buffer (normalized for OD600 to ensure equivalent bacterial numbers) and lysed at 100°C for 10 min. Proteins were separated via 12% SDS-polyacrylamide gel electrophoresis and transferred to polyvinylidene difluoride membranes. The membranes were treated with 5% nonfat milk for 1 h to block non-specific binding and incubated with primary antibody raised in mice against the FLAG-tag (1:2,500 dilution, Sigma) or DnaK (1:5,000 dilution, Abcam) for 1 h, followed by washing in TBST. The blots were further incubated with the secondary antibody goat anti-mouse-HRP (1:5,000 dilution, CWBIO) for 1 h. Blots were washed in TBST followed by detection with the ECL enhanced chemiluminescence reagent.

Statistical analysis

Statistical analysis was conducted using the software GraphPad Prism (v7.0). The mean ± SD from three independent experiments is shown in figures. Differences between two mean values were evaluated by two-tailed Student’s t test or Mann-Whitney U test. Multiple groups were compared by one-way ANOVA with Bonferroni’s post hoc analysis. To compare survival curves, the log-rank test was used. A P value <0.05 was considered to indicate statistical significance.

Supporting information

S1 Fig. Gene cluster of SPI-14 in S. Typhimurium strain 14028, with STM14_1008 gene named loiA (low oxygen induced factor A).

https://doi.org/10.1371/journal.ppat.1006429.s001

(TIF)

S2 Fig. Immunofluorescence analysis showed the invasion defect of SPI-14 mutant and loiA mutant to Caco-2 cells.

(A) The percentages of infected cells containing 0, 1, 2 or >2 bacteria. Caco-2 cells were seeded on coverslips and infected with bacteria at an MOI of 10. Infected cells were fixed 1 h post-infection and stained for immunofluorescence microscopy. The number of bacteria per Caco-2 cell was counted for at least 50 cells. Data are representative of at least three independent experiments and are presented as mean ±SD. (B) Number of intracellular bacteria per Caco-2 cell. Intracellular bacteria per cell were counted in random fields at 1 h post-infection. Bars show the mean number of bacteria contained in the infected Caco-2 cells that were counted. Data are representative of three independent experiments, with P values determined by student’s t test (***P<0.001; ns, not significant). (C) Representative images of infected Caco-2 cells by wild-type strain and loiA mutant. Bacteria were labelled with anti-Salmonella LPS antibody (green), and cell nuclei were counterstained with DAPI (blue).

https://doi.org/10.1371/journal.ppat.1006429.s002

(TIF)

S3 Fig. Growth curves of S. Typhimurium strains in vitro.

(A, B) Overnight culture of wild-type, SPI-14 mutant or loiA mutant were sub-cultured 1:100 into fresh high salt (0.3 M NaCl) LB medium and cultured for additional 24 h at 37°C with high O2 (A) or low O2 concentrations (B). The absorbance at 600 nm (OD600) of 2 ml aliquots of culture was measured regularly over this period. Data are representative of at least three independent experiments and are presented as mean ±SD.

https://doi.org/10.1371/journal.ppat.1006429.s003

(TIF)

S4 Fig. Lack of loiA did not influence S. Typhimurium systemic infection of BALB/c mice.

(A) Survival plots of BALB/c mice after inoculation intraperitoneally (i.p.) with 1×104 CFU of wild-type, loiA mutant, or SPI-14 mutant. Data presented are the combination of two independent experiments, with P value determined by log-rank curve comparison test (ns, not significant). (B) Bacterial counts recovered from liver and spleen of the BALB/c mice i.p. infected with wild-type, loiA mutant or SPI-14 mutant at day 3 post-infection. Data are combined from two independent experiments. Bars represent mean CFU of all mice, with P value determined by the Mann-Whitney U test (ns, not significant).

https://doi.org/10.1371/journal.ppat.1006429.s004

(TIF)

S5 Fig. loiA mutant did not confer additional virulence defect in systemic infection in a SPI-1 mutant background.

Bacterial counts recovered from liver (A) and spleen (B) of the BALB/c mice i.p. infected with 1×104 CFU of wild-type, loiA mutant, SPI-1 mutant or SPI-1/loiA double mutant at day 3 post-infection. Data are combined from two independent experiments. Bars represent mean CFU of all mice, with P value determined by the Mann-Whitney U test (ns, not significant).

https://doi.org/10.1371/journal.ppat.1006429.s005

(TIF)

S6 Fig. Expression of LoiA was increased under SPI-1-inducing conditions.

qRT-PCR analysis of loiA gene expression under SPI-1-inducing conditions (low O2, high salt) and non-inducing conditions (high O2, low salt; control) to late-exponential phase. Data are representative of at least three independent experiments and are presented as mean ±SD. P values were determined by student’s t test (***P<0.001).

https://doi.org/10.1371/journal.ppat.1006429.s006

(TIF)

S7 Fig. fnr mutant did not influence LoiA expression.

qRT-PCR analysis of loiA gene expression in wild-type and fnr mutant. Bacteria were grown in LB medium (0.17 M NaCl) either with low O2 (A) or high O2 (B). Data are representative of at least three independent experiments and are presented as mean ±SD. P values were determined by student’s t test.

https://doi.org/10.1371/journal.ppat.1006429.s007

(TIF)

S8 Fig. ArcA does not bind to loiA promoter.

EMSAs of loiA promoter DNA fragment with purified ArcA-His6 protein (0, 0.1, 0.2, 0.4, 0.8 and 1.6 μM). cydA promoter is used as a positive control.

https://doi.org/10.1371/journal.ppat.1006429.s008

(TIF)

S9 Fig. Mutation of arcAB and loiA cannot fully eliminate the induction of hilD by low O2 concentration.

qRT-PCR analysis of hilD gene expression in wild-type strain grown with high O2, and wild-type, loiA mutant, arcA mutant, and arcB mutant strains grown with low O2. hilD expression levels are collected from the data of Fig 6A and Fig 7C. hilD gene expression level in wild-type strain under high O2 concentration was used as a control. Data are representative of at least three independent experiments and are presented as mean ±SD. P values were determined by student’s t test (*P<0.05; ***P<0.001).

https://doi.org/10.1371/journal.ppat.1006429.s009

(TIF)

S1 Table. Strains and plasmids used in this study.

https://doi.org/10.1371/journal.ppat.1006429.s010

(DOCX)

Acknowledgments

We are grateful for technical support provided by Dr. Hengchun Cao and Dr. Zhiqiu Yin. We thank Prof. Bin Liu for helpful suggestions.

Author Contributions

  1. Conceptualization: LW LJ LF.
  2. Data curation: LW LJ LF.
  3. Formal analysis: LW LJ LF.
  4. Funding acquisition: LW LJ LF.
  5. Investigation: LJ BY WZ PW XJ.
  6. Methodology: LJ LF LW.
  7. Project administration: LW LF.
  8. Supervision: LW LF.
  9. Visualization: LJ LF BY LW.
  10. Writing – original draft: LJ LF.
  11. Writing – review & editing: LF LW LJ.

References

  1. 1. Fabrega A, Vila J. Salmonella enterica serovar Typhimurium skills to succeed in the host: virulence and regulation. Clin Microbiol Rev 2013; 26 (2): 308–341. pmid:23554419
  2. 2. Coburn B, Grassl GA, Finlay BB. Salmonella, the host and disease: a brief review. Immunol Cell Biol 2007; 85 (2): 112–118. pmid:17146467
  3. 3. LaRock DL, Chaudhary A, Miller SI. Salmonellae interactions with host processes. Nat Rev Microbiol 2015; 13 (4): 191–205. pmid:25749450
  4. 4. Rollenhagen C, Bumann D. Salmonella enterica highly expressed genes are disease specific. Infect Immun 2006; 74 (3): 1649–1660. pmid:16495536
  5. 5. Ochman H, Soncini FC, Solomon F, Groisman EA. Identification of a pathogenicity island required for Salmonella survival in host cells. Proc Natl Acad Sci U S A 1996; 93 (15): 7800–7804. pmid:8755556
  6. 6. Hayward MR, Jansen V, Woodward MJ. Comparative genomics of Salmonella enterica serovars Derby and Mbandaka, two prevalent serovars associated with different livestock species in the UK. BMC Genomics 2013; 14: 365. pmid:23725633
  7. 7. Sabbagh SC, Forest CG, Lepage C, Leclerc JM, Daigle F. So similar, yet so different: uncovering distinctive features in the genomes of Salmonella enterica serovars Typhimurium and Typhi. FEMS Microbiol Lett 2010; 305 (1): 1–13. pmid:20146749
  8. 8. Hensel M. Evolution of pathogenicity islands of Salmonella enterica. Int J Med Microbiol 2004; 294 (2–3): 95–102. pmid:15493819
  9. 9. Schlumberger MC, Hardt WD. Salmonella type III secretion effectors: pulling the host cell's strings. Curr Opin Microbiol 2006; 9 (1): 46–54. pmid:16406778
  10. 10. Blanc-Potard AB, Groisman EA. The Salmonella selC locus contains a pathogenicity island mediating intramacrophage survival. EMBO J 1997; 16 (17): 5376–5385. pmid:9311997
  11. 11. Gunn JS, Alpuche-Aranda CM, Loomis WP, Belden WJ, Miller SI. Characterization of the Salmonella typhimurium pagC/pagD chromosomal region. J Bacteriol 1995; 177 (17): 5040–5047. pmid:7665482
  12. 12. Haneda T, Ishii Y, Danbara H, Okada N. Genome-wide identification of novel genomic islands that contribute to Salmonella virulence in mouse systemic infection. FEMS Microbiol Lett 2009; 297 (2): 241–249. pmid:19583791
  13. 13. Shi L, Adkins JN, Coleman JR, Schepmoes AA, Dohnkova A, Mottaz HM, et al. Proteomic analysis of Salmonella enterica serovar typhimurium isolated from RAW 264.7 macrophages: identification of a novel protein that contributes to the replication of serovar typhimurium inside macrophages. J Biol Chem 2006; 281 (39): 29131–29140. pmid:16893888
  14. 14. Mulder DT, Cooper CA, Coombes BK. Type VI secretion system-associated gene clusters contribute to pathogenesis of Salmonella enterica serovar Typhimurium. Infect Immun 2012; 80 (6): 1996–2007. pmid:22493086
  15. 15. Morgan E, Bowen AJ, Carnell SC, Wallis TS, Stevens MP. SiiE is secreted by the Salmonella enterica serovar Typhimurium pathogenicity island 4-encoded secretion system and contributes to intestinal colonization in cattle. Infect Immun 2007; 75 (3): 1524–1533. pmid:17220313
  16. 16. Wood MW, Jones MA, Watson PR, Hedges S, Wallis TS, Galyov EE. Identification of a pathogenicity island required for Salmonella enteropathogenicity. Mol Microbiol 1998; 29 (3): 883–891. pmid:9723926
  17. 17. Bogomolnaya LM, Santiviago CA, Yang HJ, Baumler AJ, Andrews-Polymenis HL. 'Form variation' of the O12 antigen is critical for persistence of Salmonella Typhimurium in the murine intestine. Mol Microbiol 2008; 70 (5): 1105–1119. pmid:18826410
  18. 18. Velasquez JC, Hidalgo AA, Villagra N, Santiviago CA, Mora GC, Fuentes JA. SPI-9 of Salmonella enterica serovar Typhi is constituted by an operon positively regulated by RpoS and contributes to adherence to epithelial cells in culture. Microbiology 2016; 162 (8): 1367–1378. pmid:27260307
  19. 19. Latasa C, Roux A, Toledo-Arana A, Ghigo JM, Gamazo C, Penades JR, et al. BapA, a large secreted protein required for biofilm formation and host colonization of Salmonella enterica serovar Enteritidis. Mol Microbiol 2005; 58 (5): 1322–1339. pmid:16313619
  20. 20. Shah DH, Lee MJ, Park JH, Lee JH, Eo SK, Kwon JT, et al. Identification of Salmonella Gallinarum virulence genes in a chicken infection model using PCR-based signature-tagged mutagenesis. Microbiology 2005; 151 (Pt 12): 3957–3968. pmid:16339940
  21. 21. Shah DH, Zhou X, Kim HY, Call DR, Guard J. Transposon mutagenesis of Salmonella enterica serovar Enteritidis identifies genes that contribute to invasiveness in human and chicken cells and survival in egg albumen. Infect Immun 2012; 80 (12): 4203–4215. pmid:22988017
  22. 22. Lostroh CP, Lee CA. The Salmonella pathogenicity island-1 type III secretion system. Microbes Infect 2001; 3 (14–15): 1281–1291. pmid:11755416
  23. 23. Altier C. Genetic and environmental control of salmonella invasion. J Microbiol 2005; 43 Spec No: 85–92. pmid:15765061
  24. 24. Lawhon SD, Maurer R, Suyemoto M, Altier C. Intestinal short-chain fatty acids alter Salmonella typhimurium invasion gene expression and virulence through BarA/SirA. Mol Microbiol 2002; 46 (5): 1451–1464. pmid:12453229
  25. 25. Jones BD. Salmonella invasion gene regulation: a story of environmental awareness. J Microbiol 2005; 43 Spec No: 110–117. pmid:15765064
  26. 26. Ellermeier JR, Slauch JM. Fur regulates expression of the Salmonella pathogenicity island 1 type III secretion system through HilD. J Bacteriol 2008; 190 (2): 476–486. pmid:17993530
  27. 27. Bajaj V, Lucas RL, Hwang C, Lee CA. Co-ordinate regulation of Salmonella typhimurium invasion genes by environmental and regulatory factors is mediated by control of hilA expression. Mol Microbiol 1996; 22 (4): 703–714. pmid:8951817
  28. 28. Lostroh CP, Bajaj V, Lee CA. The cis requirements for transcriptional activation by HilA, a virulence determinant encoded on SPI-1. Mol Microbiol 2000; 37 (2): 300–315. pmid:10931326
  29. 29. Ellermeier CD, Ellermeier JR, Slauch JM. HilD, HilC and RtsA constitute a feed forward loop that controls expression of the SPI1 type three secretion system regulator hilA in Salmonella enterica serovar Typhimurium. Mol Microbiol 2005; 57 (3): 691–705. pmid:16045614
  30. 30. Saini S, Ellermeier JR, Slauch JM, Rao CV. The role of coupled positive feedback in the expression of the SPI1 type three secretion system in Salmonella. PLoS Pathog 2010; 6 (7): e1001025. pmid:20686667
  31. 31. Baxter MA, Fahlen TF, Wilson RL, Jones BD. HilE interacts with HilD and negatively regulates hilA transcription and expression of the Salmonella enterica serovar Typhimurium invasive phenotype. Infect Immun 2003; 71 (3): 1295–1305. pmid:12595445
  32. 32. Chubiz JEC, Golubeva YA, Lin D, Miller LD, Slauch JM. FliZ regulates expression of the Salmonella pathogenicity island 1 invasion locus by controlling HilD protein activity in Salmonella enterica serovar typhimurium. J Bacteriol 2010; 192 (23): 6261–6270. pmid:20889744
  33. 33. López-Garrido J, Casadesús J. Regulation of Salmonella enterica pathogenicity island 1 by DNA adenine methylation. Genetics 2010; 184 (3): 637–649. pmid:20008574
  34. 34. Lee AK, Detweiler CS, Falkow S. OmpR regulates the two-component system SsrA-ssrB in Salmonella pathogenicity island 2. J Bacteriol 2000; 182 (3): 771–781. pmid:10633113
  35. 35. De la Cruz M, Pérez-Morales D, Palacios I, Fernández-Mora M, Calva E, Bustamante V. The two-component system CpxR/A represses the expression of Salmonella virulence genes by affecting the stability of the transcriptional regulator HilD. Frontiers in Microbiology 2015; 6 (807).
  36. 36. Lim S, Yun J, Yoon H, Park C, Kim B, Jeon B, et al. Mlc regulation of Salmonella pathogenicity island I gene expression via hilE repression. Nucleic Acids Res 2007; 35 (6): 1822–1832. pmid:17329372
  37. 37. Baxter MA, Jones BD. Two-component regulators control hilA expression by controlling fimZ and hilE expression within Salmonella enterica serovar Typhimurium. Infect Immun 2015; 83 (3): 978–985. pmid:25547794
  38. 38. Mizusaki H, Takaya A, Yamamoto T, Aizawa S. Signal pathway in salt-activated expression of the Salmonella pathogenicity island 1 type III secretion system in Salmonella enterica serovar Typhimurium. J Bacteriol 2008; 190 (13): 4624–4631. pmid:18441068
  39. 39. Lucas RL, Lee CA. Roles of hilC and hilD in regulation of hilA expression in Salmonella enterica serovar Typhimurium. J Bacteriol 2001; 183 (9): 2733–2745. pmid:11292791
  40. 40. Lee CA, Falkow S. The ability of Salmonella to enter mammalian cells is affected by bacterial growth state. Proc Natl Acad Sci U S A 1990; 87 (11): 4304–4308. pmid:2349239
  41. 41. Gong H, Vu GP, Bai Y, Chan E, Wu R, Yang E, et al. A Salmonella small non-coding RNA facilitates bacterial invasion and intracellular replication by modulating the expression of virulence factors. PLoS Pathog 2011; 7 (9): e1002120. pmid:21949647
  42. 42. Sivick KE, Arpaia N, Reiner GL, Lee BL, Russell BR, Barton GM. Toll-like receptor-deficient mice reveal how innate immune signaling influences Salmonella virulence strategies. Cell Host Microbe 2014; 15 (2): 203–213. pmid:24528866
  43. 43. Janakiraman A, Slauch JM. The putative iron transport system SitABCD encoded on SPI1 is required for full virulence of Salmonella typhimurium. Mol Microbiol 2000; 35 (5): 1146–1155. pmid:10712695
  44. 44. Fu HA, Iuchi S, Lin EC. The requirement of ArcA and Fnr for peak expression of the cyd operon in Escherichia coli under microaerobic conditions. Mol Gen Genet 1991; 226 (1–2): 209–213. pmid:1851949
  45. 45. Alexeeva S, de Kort B, Sawers G, Hellingwerf KJ, de Mattos MJ. Effects of limited aeration and of the ArcAB system on intermediary pyruvate catabolism in Escherichia coli. J Bacteriol 2000; 182 (17): 4934–4940. pmid:10940038
  46. 46. Golubeva YA, Sadik AY, Ellermeier JR, Slauch JM. Integrating global regulatory input into the Salmonella pathogenicity island 1 type III secretion system. Genetics 2012; 190 (1): 79–90. pmid:22021388
  47. 47. Hung CC, Garner CD, Slauch JM, Dwyer ZW, Lawhon SD, Frye JG, et al. The intestinal fatty acid propionate inhibits Salmonella invasion through the post-translational control of HilD. Mol Microbiol 2013; 87 (5): 1045–1060. pmid:23289537
  48. 48. Golubeva YA, Ellermeier JR, Cott Chubiz JE, Slauch JM. Intestinal long-chain fatty acids act as a direct signal to modulate expression of the Salmonella pathogenicity island 1 type III secretion system. MBio 2016; 7 (1): e02170–02115. pmid:26884427
  49. 49. Behlau I, Miller SI. A PhoP-repressed gene promotes Salmonella typhimurium invasion of epithelial cells. J Bacteriol 1993; 175 (14): 4475–4484. pmid:8392513
  50. 50. Prouty AM, Gunn JS. Salmonella enterica serovar typhimurium invasion is repressed in the presence of bile. Infect Immun 2000; 68 (12): 6763–6769. pmid:11083793
  51. 51. Eade CR, Hung CC, Bullard B, Gonzalez-Escobedo G, Gunn JS, Altier C. Bile acids function synergistically to repress invasion gene expression in Salmonella by destabilizing the invasion regulator HilD. Infect Immun 2016; 84 (8): 2198–2208. pmid:27185788
  52. 52. Salmon KA, Hung SP, Steffen NR, Krupp R, Baldi P, Hatfield GW, et al. Global gene expression profiling in Escherichia coli K12: effects of oxygen availability and ArcA. J Biol Chem 2005; 280 (15): 15084–15096. pmid:15699038
  53. 53. Serna A, Espinosa E, Camacho EM, Casadesus J. Regulation of bacterial conjugation in microaerobiosis by host-encoded functions ArcAB and sdhABCD. Genetics 2010; 184 (4): 947–958. pmid:20083612
  54. 54. Lim S, Yoon H, Kim M, Han A, Choi J, Choi J, et al. Hfq and ArcA are involved in the stationary phase-dependent activation of Salmonella pathogenicity island 1 (SPI1) under shaking culture conditions. J Microbiol Biotechnol 2013; 23 (12): 1664–1672. pmid:24018968
  55. 55. Van Immerseel F, Eeckhaut V, Boyen F, Pasmans F, Haesebrouck F, Ducatelle R. Mutations influencing expression of the Salmonella enterica serovar Enteritidis pathogenicity island I key regulator hilA. Antonie Van Leeuwenhoek 2008; 94 (3): 455–461. pmid:18563621
  56. 56. Fink RC, Evans MR, Porwollik S, Vazquez-Torres A, Jones-Carson J, Troxell B, et al. FNR is a global regulator of virulence and anaerobic metabolism in Salmonella enterica serovar Typhimurium (ATCC 14028s). J Bacteriol 2007; 189 (6): 2262–2273. pmid:17220229
  57. 57. Cirillo DM, Valdivia RH, Monack DM, Falkow S. Macrophage-dependent induction of the Salmonella pathogenicity island 2 type III secretion system and its role in intracellular survival. Mol Microbiol 1998; 30 (1): 175–188. pmid:9786194
  58. 58. Datsenko KA, Wanner BL. One-step inactivation of chromosomal genes in Escherichia coli K-12 using PCR products. Proc Natl Acad Sci U S A 2000; 97 (12): 6640–6645. pmid:10829079
  59. 59. Cherepanov PP, 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; 158 (1): 9–14. pmid:7789817
  60. 60. Wang RF, Kushner SR. Construction of versatile low-copy-number vectors for cloning, sequencing and gene expression in Escherichia coli. Gene 1991; 100: 195–199. pmid:2055470
  61. 61. Ryan D, Ojha UK, Jaiswal S, Padhi C, Suar M. The small RNA DsrA influences the acid tolerance response and virulence of Salmonella enterica serovar Typhimurium. Front Microbiol 2016; 7: 599. pmid:27199929
  62. 62. Vazquez-Torres A, Jones-Carson J, Mastroeni P, Ischiropoulos H, Fang FC. Antimicrobial actions of the NADPH phagocyte oxidase and inducible nitric oxide synthase in experimental salmonellosis. I. Effects on microbial killing by activated peritoneal macrophages in vitro. J Exp Med 2000; 192 (2): 227–236. pmid:10899909
  63. 63. Valdivia RH, Falkow S. Fluorescence-based isolation of bacterial genes expressed within host cells. Science 1997; 277 (5334): 2007–2011. pmid:9302299
  64. 64. Yang B, Feng L, Wang F, Wang L. Enterohemorrhagic Escherichia coli senses low biotin status in the large intestine for colonization and infection. Nat Commun 2015; 6: 6592. pmid:25791315
  65. 65. Wu G, Cruz-Ramos H, Hill S, Green J, Sawers G, Poole RK. Regulation of cytochrome bd expression in the obligate aerobe Azotobacter vinelandii by CydR (Fnr). Sensitivity to oxygen, reactive oxygen species, and nitric oxide. J Biol Chem 2000; 275 (7): 4679–4686. pmid:10671497
  66. 66. Lucchini S, Rowley G, Goldberg MD, Hurd D, Harrison M, Hinton JC. H-NS mediates the silencing of laterally acquired genes in bacteria. PLoS Pathog 2006; 2 (8): e81. pmid:16933988