Introduction

 

Listeria monocytogenes (LM) is a Gram-positive zoonotic pathogen that infects humans through ingestion of contaminated food (Pérez-Trallero et al. 2014; Fagerlund et al. 2016), which can cause abortion in pregnant women, meningoencephalitis in infants, gastroenteritis and other fatal illnesses (Lotfollahi et al. 2017). Likewise, livestock population is mainly infected after fed with LM-contaminated feed, leading to abortion and a high mortality rate of 20–30% (Lomonaco et al. 2015; Boqvist et al. 2018; McDougal and Sauer 2018; Yao et al. 2018). What’s more, as an important food-borne pathogen, LM can survive in a variety of complex environmental stress conditions, such as low temperature, high temperature, high concentration of salt and low pH, causing serious harm to food safety and animal husbandry (Colagiorgi et al. 2017; Boqvist et al. 2018; Chaturongakul et al. 2008; Hurley et al. 2019).

As a facultative intracellular pathogen, the process of LM infection requires the participation of many virulence factors and genes, such as internalin A (inlA), internalin B (inlB), adhesion protein (LAP), phospholipase C (plcA and plcB), listeriolysin O (LLO) and actin polymerizing factor (actA) (Chen et al. 2017; Drolia et al. 2018). However, the expression of these virulence factors is regulated by PrfA, SigB and other regulators, enabling LM to survive and reproduce in complex intracellular environments (O'Byrne and Karatzas 2008; Heras et al. 2011).

So far, many studies have found that AraC family proteins play important roles in regulating virulence, multiple drug resistance and metabolic reactions in bacteria (Gallegos et al. 1997; Bailey et al. 2010). lmo2672, as a member of AraC family protein, is encoded by a novel differential gene lmo2672 between virulent and avirulent strains of LM (Roche et al. 2008; Hadjilouka et al. 2016), However, the biological function of lmo2672 is still unclear. The purpose of this study is to explore the role of lmo2672 on pathogenicity by construction of lmo2672 deletion strain and its comparison with that of the wild-type strain, which would provide insights into the biological function of lmo2672 in LM virulent regulation.

 

Materials and Methods

 

Strain and plasmid

 

LM EGD-e strain (serotype 1/2a) was kindly donated by Dr. W. Goebel, University of Woodsburg, Germany. A temperature-sensitive plasmid pKSV7 was a gift from Professor Zhu Guoqiang of Yangzhou University. Escherichia coli (E. coli) DH5α (TaKaRa, Japan) and BL21 (DE3) (TaKaRa, Japan) were grown in Luria-Bertani (LB) broth (Difco, USA). LM strain was cultured in brain-heart infusion (BHI) broth (Difco, USA) or tryptic soy agar (TSA) plate (Difco, USA) at 37ºC

 

Primer design and synthesis

 

Eleven pairs of primers were designed by Primer Premier 5.0 (Premier Inc., Canada) according to the genome sequence of LM EGD-e in GenBank (accession number: AL591824.1) (Table 1). The primers were synthesized by Huada Biotechnology Co., Ltd. (Beijing, China).

 

Generation of LM-lmo2672 deletion strain

 

A lmo2672 deletion strain was constructed using homologous recombination technique. Briefly, the DNA fragments containing upstream and downstream regions flanking the lmo2672 gene in LM EGD-e strain were amplified using PCR with primers F1/F2 and F3/F4, respectively. The PCR products were recovered and the deletion fragment of the lmo2672 gene was generated using overlap extension PCR (SOE-PCR). The deletion fragment was then cloned into the plasmid pKSV7 to generate pKSV7-lmo2672. The recombinant shuttle plasmid (pKSV7-lmo2672) was introduced into a competent strain LM EGD-e by electroporation (2.5 kV, 100 Ω and 25 µF). Then, the transformed bacteria were selected on BHI agar (Difco, USA) with chloramphenicol (12.5 µg/mL) (Amresco, USA) at 42ºC for several passages. The recombinant strain LM-lmo2672 was further screened by PCR with the primer pair F7/F8, and PCR product was sequenced for molecular identification.

 

Assay of hemolytic activity

 

In brief, overnight cultures of wild-type strain and lmo2672 mutant were diluted 1:10 in fresh medium, respectively, and 100 µL of the diluted bacterial suspension was cultivated on tryptic soy agar (TSA) plate (Difco, USA) with 5% sheep red blood cells (SRBCs) and incubated at 37ºC for 24 h. The size of transparent hemolytic ring around the colony was recorded. The hemolytic activity was determined as previously reported with slight modifications (Alonzo et al. 2009). The optical density at 600 nm (OD600) was measured and normalized for each strain (OD600 = 0.5), followed by the addition of 1 mL PBS (pH=5.6) containing SRBCs, then 30 min of incubation at 37ºC. The mixtures were then centrifuged to pellet unlysed cell and the hemoglobin absorbance in the supernatants was measured at 543 nm.

 

Macrophage adhesion, invasiveness and intracellular survival assays

 

Briefly, mouse macrophages cells RAW264.7 were seeded in the 6-well plates (Gibco, USA) at 37ºC in 5% CO2. Wild-type strain and lmo2672 mutant were collected at the mid-log phase, respectively. Then washed three times with PBS (pH=7.2) and re-suspended in DMEM (Gibco, USA). The cells were subjected to infection with the wild-type strain and lmo2672 mutant at a multiplicity of infection (MOI) of 10:1. After infection, the cells were washed three times with PBS, and lysed with 500 µL 0.1% Triton X-100 (Amresco Inc., USA) for 10 min. Then the bacteria in the lysate were counted. The adhesion rate, invasion rate and survival rate were calculated as previously reported (Zhang et al. 2013). All the infections were performed three replicates.

 

Mouse virulence assay

 

The overnight cultures of wild-type strain and LM-lmo2672 were plated on BHI agar to calculate the colony number. Three groups of Six-week-old BALB/c mice, 10 mice per group, were inoculated intraperitoneally with 200 µL different dilutions of bacteria (105, 106, 107, 108, and 109 CFU/mouse, respectively), one group was inoculated with 200 µL sterile PBS (pH=7.2). Then mice were assessed 10 days after inoculation and recorded for the number of deaths daily and the LD50 was calculated using the Karber method. Besides, according to the death of mice in the group of 107 CFU injection dose, the Kaplan-Meier survival curve was drawn. Meanwhile, the mice were divided into three groups randomly and intraperitoneally injected with 2×105 CFU of LM EGD-e, LM-lmo2672 and sterile PBS (pH=7.2), respectively. Then, the spleens and livers of mice were homogenized, and bacteria loads were determined through enumeration of CFU. On the fifth day after infection, the organs were fixed in formalin and tissue sections were prepared for histopathological analysis.

 

Transcriptional analysis of virulence genes

 

Briefly, LM-lmo2672 and LM EGD-e were incubated to the logarithmic growth phase. At this point, bacteria were harvested by centrifugation, and total RNA were extracted for using Trizol agent (Invitrogen, USA). Reverse transcription into cDNA was performed using the AMV reverse transcription kit (TaKaRa, Japan). The qRT-PCR was performed using the reversely transcripted cDNA using SYBR Premix Ex TaqTM kit (TaKaRa, Japan) according to the manufacturer’s instructions on LightCycler 480 instrument (Roche, Swiss). qRT-PCR analysis was conducted to determine the effects of lmo2672 on the transcription of genes associated with virulence, sigB, prfA, hly, actA and inlB genes were determined using the 2-△△CT method with the housekeeping gene rpoB as internal reference.

 

Electrophoretic mobility shift assay (EMSA)

 

Eectrophoretic mobility shift assay (EMSA) was carried out to further confirm the interaction between lmo2672 and the upstream DNA sequence of prfA gene (containing promoter P1 and P2). Briefly, the full length lmo2672 gene was cloned into pET32a vector, and then the recombinant plasmid (pET32a- lmo2672) was transformed into DE3 for the expression under the induction of 0.5 mM IPTG for 8 h at 37ºC. The recombinant protein lmo2672 was purified using Ni-NTA affinity chromatography (Invitrogen, USA). The promoter fragment of prfA gene was generated by PCR and purified using a Gel Extraction Kit (Omega, USA). The lmo2672 protein was mixed with 100 ng DNA in 20 µL of the gel-shift binding buffer (100 mM NaCl, 100 mM Tris-HCl, 1 mM DTT, 10% glycerol). Meanwhile, Bovine serum albumin (BSA) was used as negative control. After incubation at 25ºC for 30 min, the sample was analyzed by 8% non-denaturing polyacrylamide gel electrophoresis, then the polacrylamide gel was stained with ethidium bromide (EB) and visualizaed under ultraviolet light.

 

Statistical analysis

 

One-way analysis of variance was performed using GraphPad Prism software, version 5.0 (GraphPad Software Inc., USA). The analyses were expressed as the mean ± standard error of the mean. The statistical comparison between different variables was done using P < 0.05 and P < 0.01 as level of significance.

 

Results

 

Using the flanking primer pair F5/F6, a 1502 bp fragment was generated from LM-lmo2672. Compared with the wild-type strain LM EGD-e, 807 bp was lost from LM-lmo2672 (Supplementary Fig. 1A). After 20 passages, a fragment of expected size was amplified by PCR (Supplementary Fig. 1B). Sequence analysis showed that the lmo2672 gene was successfully deleted (Supplementary Fig. 2).

The hemolytic ring (Fig. 1A, Fig. 1B) and erythrocytolysis (Fig. 1C, Fig. 1D) of LM-lmo2672 were smaller or significantly weaker than these of wild-type strain (P < 0.05), indicating that the deletion of lmo2672 reduces the hemolytic ability of LM, which is consistent with the results of transcription level.

Compared with wild-type strain LM EGD-e, there was no significant impacted adhesion rate oflmo2672 mutant (P > 0.05) (Fig. 2A), but the invasion rate was significantly lower (P < 0.01) (Fig. 2B). Within 12 h after infection, the proliferation capability to RAW 264.7 cells in lmo2672 mutant strain was significantly lower than that in wild-type strain LM EGD-e (P < 0.01) (Fig. 2C), suggesting that lmo2672 gene deficiency can reduce the ability of LM infection.

Compared with wild-type strain LM EGD-e, the LD50 of lmo2672 mutant increased by 6.3 times (P < 0.05) (Fig. 3A). The average survival time of mice infected with lmo2672 mutant was significantly increased (P < 0.05) (Fig. 3B). The bacterial load of LM-lmo2672 was significantly less than that of wild-type strain (P < 0.05) in liver (Fig. 3C), while, the bacterial load was significantly less than that of wild-type strain (P < 0.05) in spleen (Fig. 3D), suggesting that the gene deletion of lmo2672 significantly reduced survival and reproduction ability of wild-type strain in liver and spleen.

Compared with normal mice (injected with PBS) (Fig. 4E, Fig. 4F), histopathological changes in liver of the mice injected with wild-type strain and lmo2672 mutant were characterized by vacuoles in hepatocytes, degeneration and necrosis, and hemorrhagic necrosis, while the structure of splenic cord disappeared with hemorrhage and lymphocyte infiltration. Compared with wild-type strain LM EGD-e (Fig. 4A and Fig. 4B), however, the pathological changes of liver and spleen in LM-lmo2672 infected mice were significantly reduced (Fig. 4C and Fig. 4D).

The qRT-PCR analysis revealed that lmo2672 deletion down-regulated the expression of five genes associated with virulence. Compared with wild-type strain, the transcription levels of virulence genes prfA, hly and actA in lmo2672 mutant except inlB and sigB were significantly lower (P < 0.05) (Fig. 5).

The recombinant protein lmo2672 could bind to upstream DNA sequence of prfA, (Fig. 6, Supplementary Fig. 3), which implied that there was an interaction between lmo2672 and the upstream DNA sequence of prfA gene.

 

Discussion

 

AraC family transcription regulators (AFTRs) have been shown to be an important class of regulators in bacteria. They are generally composed of 200 and 300 amino acids arranged in two characteristic domains: one is the conserved DNA binding domain at the C-terminal (CTD), the other is the variable N-terminal domain (NTD) (Gallegos et al. 1997; Yang et al. 2011), which are necessary for in vivo transcriptional activation (Porter and Dorman 2002). Furthermore, most members of the AraC family have been demonstrated to be involved in virulence regulation in many bacteria such as Vibrio cholerae (Champion et al. 1997; Krukonis and DiRita 2003), Citrobacter rodentium (Kelly et al. 2006; Hart et al. 2008; Santiago et al. 2016) and Shigella flexneri (Koppolu et al. 2013; Martino et al. 2016). Garrity-Ryan et al. confirmed that the virulence of Yersinia pseudotuberculosis could be significantly reduced by inhibiting or knocking out the transcriptional regulatory genes of the AraC family (Gallegos et al. 1997; Martin and Rosner 2001; Garrity-Rya et al. 2010). However, the role of AraC family member lmo2672 in pathogenicity of LM is still unclear. Here, the impacts of lmo2672 gene deficiency on LM pathogenicity were investigated using LM-lmo2672 decifiency strain, erythrocytolysis, macrophage infection and mouse infection assays. The results showed that the lmo2672 deletion strain has significantly reduced erythrocytolysis, adhesion, invasion and intracellular proliferation of macrophage RAW264.7, and significantly lowered virulence to mice, suggesting that lmo2672 is involved in pathogenicity of LM.

LM can survive in complex environment inside hosts, which is due to its complex gene regulatory network (O'Byrne and Karatzas 2008; Heras et al. 2011; Guariglia-Oropeza et al. 2014). It has been shown that the AraC family transcription factors regulate the expression of bacterial genes (Bailey et al. 2010). Recently, Rowe et al. found that AraC-type regulator Rbf influences the Staphylococcus epidermidis biofilm formation by regulating the expression of the icaADBC gene, and subsequently influences its drug resistance and virulence (Rowe et al. 2016). Fang et al. found that BfvR directly or indirectly regulate the expression of psaABC and psaEF to inhibit the pathogenicity of Y. pestis in mice (Fang et al. 2018).

 

 

Fig. 1: Determination of hemolytic activity of wild-type strain LM EGD-e and LM-lmo2672

(A) and (B): LM EGD-e and LM-lmo2672 cultured on sheep blood agar for 24 h, respectively. (C): Hemolytic activity of mutant and LM-lmo2672. (D): OD543 nm value of the supernatant was measured. Error bars represent the mean ± standard error (SE) of triplicate experiments. Stars indicate P-values < 0.05 (*)

 

 

 

Fig. 2: Determination of adhesion, invasion and replication of wild-type strain LM EGD-e and LM-lmo2672 in RAW 264.7 cells

The macrophage RAW 264.7 is infected with bacterial suspension to a MOI of 10 bacteria per cell

(A): Adhesion rates of wild-type strain LM EGD-e and lmo2672 mutant; (B): Invasion rates of wild-type strain LM EGD-e and lmo2672 mutant; (C): The number of bacteria in live in different time. Error bars represent the mean ± standard error (SE) of triplicate experiments. Stars indicate P-values < 0.01(**)

In this study, the qRT-PCR found that lmo2672 mutant repressed the expression of three deletion of virulence genes (prfA, actA, hly). PrfA is the member of the Crp/Fnr transcription regulator families in LM (Hall et al. 2016), which can directly or indirectly regulate the expression of many virulence genes in LM (Heras et al. 2011). ActA (actin aggregation factor) is involved in the cell adhesion and invasion of LM (Bruhn et al.

 

 

Fig. 3: Effects of lmo2672 gene deletion of LM on virulence in mice

(A): The LD50 of wild-type strain LM EGD-e and LM-lmo2672; (B): Survival curves of mice infected by intraperitoneal injection of wild-type strain LM EGD-e and LM-lmo2672; (C): Bacteria count in liver; (D): Bacteria count in spleen. Error bars represent the mean ± standard error (SE) of triplicate experiments. Stars indicate P-values < 0.05 (*)

 

 

 

Fig. 4: Histopathological analysis of liver and spleen of mice infected by wild-type strain LM EGD-e and LM-lmo2672 (HE staining, × 400)

(A) and (C): The liver of the mice injected with LM EGD-e and LM-lmo2672. Hepatocytes showed vacuol-like appearance, degeneration necrosis and hemorrhagic necrosis (indicated by arrows); (B) and (D): The spleen of the mice injected with LM EGD-e and LM-lmo2672. The structure of splenic cord disappeared with hemorrhage and lymphocyte infiltration (indicated by arrows); (E) and (F): Liver and spleen tissue’s slices of mice inoculated with PBS (control)

2007; Hadjilouka et al. 2018), which promotes the aggregation of actin to drive LM through the cytoplasm into adjacent cells (Milohanic et al. 2003; Hamon et al. 2012; Brenner et al. 2018). Hly encodes virulence factor LLO, enabling LM to escape from the phage of host cells and replicate inside cells (Chen et al. 2017). This study showed the significant decrease of hly gene transcription level is consistent with the phenotype of reduced hemolytic activity. Besides, the transcription levels of prfA and actA in LM-lmo2672 are significantly reduced, which is consistent with the results of LM macrophage cell and mouse infection assays. In addition, the interaction between lmo2672 and the upstream sequence of prfA gene was also confirmed by EMSA.

Conclusion

 

We for the first time demonstrated that lmo2672 gene deficiency could reduce the virulence of LM, indicating that Table 1: List of primer sequences used in this study

 

Target  gene

Primer name

Nucleotide sequence (5’→3’)

Product size (bp)

lmo2672-L

F1

GGGGTACCCTGTCATTTTTTCTCCTCCT

638

 

F2

CAAAGCATTTACGTTTTAAAGAGACCCCCTTTTC

lmo2672-R

F3

GAAAAGGGGGTCTCTTTAAAACGTAAATGCTTTG

698

 

F4

AACTGCAG GCGAATCAAGTCTTTATCTC

lmo2672-I 

F5

ATAACGTCGCAAGGTGCATG

2309/1502

 

F6

GGTCCATACAGAAAACCACGA

lmo2672-E

F7

ATGATTAATGAATTTGTTTGTA

840

 

F8

TTAGAGTTTTTCGACAGTCT

rpoB

rpoB-F

TGCCATTTATGCCAGAC

188

 

rpoB-R

TTCTTCCACTGTGCTCC

prfA

prfA-F

TTAGCGAGAACGGGACCAT

392

 

prfA-R

TGCGATACCGCTTGAATAG

actA

actA-F

GGCGAAAGAGTCAGTTGC

492

 

actA-R

GTTGGAGGCGGTGGAAAT

hly

hly-F

TGTAAACTTCGGCGCAATC

462

 

hly-R

TAAGCAATGGGAACTCCTG

sigB

sigB-F

TCATCGGTGTCACGGAAGAA

310

 

sigB-R

TGACGTTGGATTCTAGACAC

inlB

inlB-F

TATGCAGCATGGCTTGTAACC

200

 

inlB-R

GTTCTTGCAGAGATGGCACG

prfA-EMSA

prfA-1

AGTATATCTCCGAGCAACCTCG

243

 

prfA-2

GTCTCATCCCCCAATCGTT

 

lmo2672 was involved in the pathogenicity of LM, which provided an insight into the biological function of lmo2672 in LM virulent regulation.

 

 

Fig. 5: Determination of transcriptional levels of virulence related genes in LM EGD-e and LM-lmo2672

Error bars represent the mean ± standard error (SE) of triplicate experiments. Stars indicate P-values < 0.05 (*) and < 0.01(**)

 

 

 

Fig. 6: Analysis of the interaction between lmo2672 and the upstream DNA sequence of prfA gene using EMSA

(A): Analysis of expressed lmo2672 by SDS-PAGE

Lane 1, Purified lmo2672 protein; lane 2, Expression of pET32a vector in E. coli BL21(DE3); lanes 3–4, Expression of recombinant lmo2672 in E. coli BL21(DE3)

(B) The interaction between lmo2672 and the upstream DNA sequence of prfA gene

Lane 1-5, The upstream DNA sequence of prfA gene incubated with an increasing protein lmo2672;

(C) The interaction between BSA and the upstream DNA sequence of prfA gene

Lane 6-10: an increasing BSA (unrelated protein) binds to the upstream DNA sequence of prfA gene as negative control; Shifted bands are indicated by arrow heads

 

Acknowledgements

 

We thank Dr. W. Goebel and Professor Zhu Guoqiang who provided the samples for this study. This work was supported by the national key research and development program (No. 2016YFD0500900), National Natural Science Foundation of China (31360596, 30960274), Young and middle-aged leading science and technology innovation talents plan of Xinjiang Corps (No. 2016BC001), International Science & Technology Cooperation Program of China (No. 2014DFR31310) and Outstanding young and middle-aged talents Training Project of State Key Laboratory for Sheep Genetic Improvement and Healthy Production (SKLSGIHP2017A03).

 

References

 

Alonzo F, GC Port, M Cao, NE Freitag (2009). The post translocation chaperone PrsA2 contributes to multiple facets of Listeria monocytogenes pathogenesis. Infect Immun 77:26122623

Bailey AM, A Ivens, R Kingsley, JL Cottell, J Wain, LJ Piddock (2010). RamA, a member of the AraC/XylS family, influences both virulence and efflux in Salmonella enterica serovar Typhimurium. J Bacteriol 192:16071616

Boqvist S, K Söderqvist, I Vågsholm (2018). Food safety challenges and One Health within Europe. Acta Vet Scand 60:1–13

Brenner M, L Lobel, I Borovok, N Sigal, AA Herskovits (2018). Controlled branched-chain amino acids auxotrophy in Listeria monocytogenes allows isoleucine to serve as a host signal and virulence effector. PLoS Genet 14; Article e1007283

Bruhn KW, N Craft, JF Miller (2007). Listeria as a vaccine vector. Microb Infect 9:12261235

Champion GA, MN Neely, MA Brennan, VJ DiRita (1997). A branch in the ToxR regulatory cascade of Vibrio cholerae revealed by characterization of toxT mutant strains. Mol Microbiol 23:323331

Chaturongakul S, S Raengpradub, M Wiedmann, KJ Boor (2008). Modulation of stress and virulence in Listeria monocytogenes. Trends Microbiol 16:388396

Chen GY, DA Pensinger, JD Sauer (2017). Listeria monocytogenes cytosolic metabolism promotes replication, survival, and evasion of innate immunity. Cell Microbiol 19:110

Colagiorgi A, I Bruini, PAD Ciccio, E Zanardi, S Ghidini, A Ianieri (2017). Listeria monocytogenes biofilms in the wonderland of food industry. Pathogens 6; Article 41

Drolia R, S Tenguria, AC Durkes, JR Turner, AK Bhunia (2018). Listeria adhesion protein induces intestinal epithelial barrier dysfunction for bacterial translocation. Cell Host Microb 23:470484

Fagerlund A, S Langsrud, BCT Schirmer, T Møretrø, E Heir (2016). Genome analysis of Listeria monocytogenes sequence type 8 strains persisting in Salmon and poultry processing environments and comparison with related strains. PLoS One 11; Article e0151117

Fang H, L Liu, Y Zhang, H Yang, Y Yan, X Ding, Y Han, D Zhou, R Yang (2018). BfvR, an AraC-family regulator, controls biofilm formation and pH6 antigen production in opposite ways in Yersinia pestis Biovar Microtus. Front Cell Infect Microbiol 8; Article 347

Gallegos MT, R Schleif, A Bairoch, K Hofmann, JL Ramos (1997). Arac/XylS family of transcriptional regulators. Microbiol Mol Biol Rev 61:393410

Garrity-Ryan LK, OK Kim, JM Balada-Llasat, VJ Bartlett, AK Verma, ML Fisher, C Castillo, W Songsungthong, SK Tanaka, SB Levy, J Mecsas, MN Alekshun (2010). Small molecule inhibitors of LcrF, a Yersinia pseudotuberculosis transcription factor, attenuate virulence and limit infection in a murine pneumonia model. Infect Immun 78:46834690

Guariglia-Oropeza V, RH Orsi, H Yu, KJ Boor, M Wiedmann, C Guldimann (2014). Regulatory network features in Listeria monocytogenes-changing the way we talk. Front Cell Infect Microbiol 4; Article 14

Hadjilouka A, S Paramithiotis, EH Drosinos (2018). Genetic Analysis of the Listeria Pathogenicity Island 1 of Listeria monocytogenes 1/2a and 4b Isolates. Curr Microbiol 75:857865

Hadjilouka A, C Molfeta, O Panagiotopoulou, S Paramithiotis, M Mataragas, EH Drosinos (2016). Expression of Listeria monocytogenes key virulence genes during growth in liquid medium, onrocket and melon at 4, 10 and 30°C. Food Microbiol 55:715

Hall M, C Grundström, A Begum, MJ Lindberg, UH Sauer, F Almqvist, J Johansson, AE Sauer-Eriksson (2016). Structural basis for glutathione-mediated activation of the virulence regulatory protein PrfA in Listeria. Proc Natl Acad Sci USA 113:1473314738

Hamon MA, D Ribet, F Stavru, P Cossart (2012). Listeriolysin O: The Swiss army knife of Listeria. Trends Microbiol 20:360368

Hart E, J Yang, M Tauschek, M Kelly, MJ Wakefield, G Frankel, EL Hartland, RM Robins-Browne (2008). RegA, an AraC-like protein, is a global transcriptional regulator that controls virulence gene expression in Citrobacter rodentium. Infect Immun 76:52475256

Heras ADL, RJ Cain, MK Bielecka, JA Vázquez-Boland (2011). Regulation of Listeria virulence: PrfA master and commander. Curr Opin Microbiol 14:118127

Hurley D, L Luque-Sastre, CT Parker, S Huynh, AK Eshwar, SV Nguyen, N Andrews, A Moura, EM Fox, K Jordan, A Lehner, R Stephan, S Fanning (2019). Whole-genome sequencing-based characterization of 100 Listeria monocytogenes isolates collected from food processing environments over a four-year period. mSphere 4; Article e00252-19

Kelly M, E Hart, R Mundy, O Marchès, S Wiles, L Badea, S Luck, M Tauschek, G Frankel, RM Robins-Browne, EL Hartland (2006). Essential role of the type III secretion system effector NleB in colonization of mice by Citrobacter rodentium. Infect Immun 74:23282337

Koppolu V, I Osaka, JM Skredenske, B Kettle, PS Hefty, J Li, SM Egan (2013). Small-molecule inhibitor of the Shigella flexneri master virulence regulator VirF. Infect Immun 81:42204231

Krukonis ES, VJ DiRita (2003). From motility to virulence: Sensing and responding to environmental signals in Vibrio cholerae. Curr Opin Microbiol 6:186190

Lomonaco S, D Nucera, V Filipello (2015). The evolution and epidemiology of Listeria monocytogenes in Europe and the United States. Infect Genet Evol 35:172183

Lotfollahi L, A Chaharbalesh, M Ahangarzadeh Rezaee, A Hasani (2017). Prevalence, antimicrobial susceptibility and multiplex PCR-serotyping of Listeria monocytogenes isolated from humans, foods and livestock in Iran. Microb Pathog 107:425429

Martin RG, JL Rosner (2001). The AraC transcriptional activators. Curr Opin Microbiol 4:132137

Martino MLD, C Romilly, EGH Wagner, B Colonna, G Prosseda (2016). One gene and two proteins: A leaderless mRNA supports the translation of a shorter form of the Shigella VirF regulator. mBio 7; Article e01860-16

McDougal CE, JD Sauer (2018). Listeria monocytogenes: The Impact of Cell Death on Infection and Immunity. Pathogens 7; Article 8

Milohanic E, P Glaser, JY Coppée, L Frangeul, Y Vega, JA Vázquez-Boland, F Kunst, P Cossart, C Buchrieser (2003). Transcriptome analysis of Listeria monocytogenes identifies three groups of genes differently regulated by PrfA. Mol Microbiol 47:16131625

O'Byrne CP, KA Karatzas (2008). The role of sigma B (sigma B) in the stress adaptations of Listeria monocytogenes: Overlaps between stress adaptation and virulence. Adv Appl Microbiol 65:115140

Pérez-Trallero E, C Zigorraga, J Artieda, M Alkorta, JM Marimón (2014). Two outbreaks of Listeria monocytogenes infection, Northern Spain. Emerg Infect Dis 20:21552157

Porter ME, CJ Dorman (2002). In vivo DNA-binding and oligomerization properties of the Shigella flexneri AraC-like transcriptional regulator VirF as identified by random and site-specific mutagenesis. J Bacteriol 184:531539

Roche SM, P Velge, D Liu (2008). Virulence determination. In: Handbook of Listeria Monocytogenes, pp:241–270. Liu D (Ed.). Taylor & Francis Group, New York, USA

Rowe SE, C Campbell, C Lowry, ST O'Donnell, ME Olson, JK Lindgren, EM Waters, PD Fey, JP O'Gara (2016). AraC-Type Regulator Rbf Controls the Staphylococcus epidermidis Biofilm Phenotype by Negatively Regulating the icaADBC Repressor SarR. J Bacteriol 198:29142924

Santiago AE, MB Yan, M Tran, N Wright, DH Luzader, MM Kendall, F Ruiz-Perez, JP Nataro (2016). A large family of anti-activators accompanying XylS/AraC family regulatory proteins. Mol Microbiol 101:314–332

Yang J, M Tauschek, RM Robins-Browne (2011). Control of bacterial virulence by AraC-like regulators that respond to chemical signals. Trends Microbiol 19:128135

Yao H, M Kang, Y Wang, Y Feng, S Kong, X Cai, Z Ling, S Chen, X Jiao, Y Yin (2018). An essential role for hfq involved in biofilm formation and virulence in serotype 4b Listeria monocytogenes. Microbiol Res 215:148154

Zhang Z, Q Meng, J Qiao, L Yang, X Cai, G Wang, C Chen, L Zhang (2013). RsbV of Listeria monocytogenes contributes to regulation of environmental stress and virulence. Arch Microbiol 195:113120