Skip to main content
Infection and Immunity logoLink to Infection and Immunity
. 2009 Nov 30;78(3):887–897. doi: 10.1128/IAI.00882-09

Role of RpoS in Virulence of Pathogens

Tao Dong 1, Herb E Schellhorn 1,*
PMCID: PMC2825926  PMID: 19948835

Abstract

Understanding mechanisms of bacterial pathogenesis is critical for infectious disease control and treatment. Infection is a sophisticated process that requires the participation of global regulators to coordinate expression of not only genes coding for virulence factors but also those involved in other physiological processes, such as stress response and metabolic flux, to adapt to host environments. RpoS is a key response regulator to stress conditions in Escherichia coli and many other proteobacteria. In contrast to its conserved well-understood role in stress response, effects of RpoS on pathogenesis are highly variable and dependent on species. RpoS contributes to virulence through either enhancing survival against host defense systems or directly regulating expression of virulence factors in some pathogens, while RpoS is dispensable, or even inhibitory, to virulence in others. In this review, we focus on the distinct and niche-dependent role of RpoS in virulence by surveying recent findings in many pathogens.


RpoS is an alternative sigma factor of RNA polymerase primarily found in Beta- and Gammaproteobacteria (31, 59). RNA core polymerase requires a sigma factor for promoter recognition and transcription initiation. In addition to housekeeping sigma factors that control transcription of essential genes, bacteria also possess alternative sigma factors that recognize the promoters of a specific set of genes. There are seven known sigma factors in the Gram-negative model bacterium Escherichia coli (67) and 18 in the Gram-positive bacterium Bacillus subtilis (52). The contribution of alternative sigma factors to virulence can be direct through regulated expression of virulence genes or indirect by enhancing survival against host defense and other stress conditions (70).

Pathogenic bacteria experience many stresses during transmission and infection. For example, the enterohemorrhagic E. coli (EHEC) O157:H7 strain may face nutrient limitation and heat exposure in natural environments and acid stress and host defense after entry into human hosts. The ability to quickly adapt to changing environments is therefore critical for bacterial pathogens to successfully transmit and infect hosts. One of the most important adaptation factors in E. coli is RpoS (31, 59). The RpoS regulon, comprising 10% of E. coli genes (32, 33, 78, 108, 141), plays a critical role in survival of several stresses, including acid (124), heat (61), oxidative stress (116), starvation (79), and near-UV exposure (116). In E. coli, the levels of RpoS are low in exponential phase (32, 80), due to reduced transcription (80), attenuated translation (80), and, most importantly, rapid proteolysis mediated by RssB, a chaperone protein that binds to RpoS and directs the RssB-RpoS complex to the ClpXP protease (80, 93, 109, 150). The degradation of RpoS is suppressed in stationary phase (11, 150), resulting in increased RpoS levels (80). Expression of RpoS is sensitive to environmental changes and is under the control of many regulatory factors, such as acetate, ppGpp, and cyclic AMP (cAMP) (reviewed in references 31 and 59).

The RpoS-bearing bacteria have a broad host range, including human pathogens (e.g., E. coli and Vibrio cholerae), animal pathogens (e.g., Citrobacter rodentium and Salmonella enterica serovar Typhimurium), insect pathogens (e.g., Serratia entomophila and Xenorhabdus nematophilus), and plant pathogens (e.g., Burkholderia plantarii, Erwinia carotovora, and Ralstonia solanacearum). RpoS is required for resistance to many stresses in these bacteria (Table 1). However, the effect of RpoS on virulence is variable, differing even in closely related species. RpoS is required for virulence in some pathogens, including Salmonella enterica, V. cholerae, B. plantarii, and S. entomophila, but is less important in other species (Table 1). Despite the considerable accumulated information on RpoS control of virulence functions in specific bacteria, there is, as yet, no comprehensive review on this topic. Therefore, this review summarizes the involvement and contribution of RpoS in virulence of RpoS-bearing pathogens. We place special focus on studies that have tested rpoS in host (animal or plant) or cell culture models.

TABLE 1.

Effects of RpoS on virulence of specific pathogens

Organism RpoS-dependent phenotypea
Virulence factor(s) controlled by RpoS Role of RpoS in virulence Model Reference
Starvation Oxidative Acid Heat Osmotic Motility
Escherichia coli
    K-12 + + + + + NAb NA NA 33, 60, 108
    BJ4 NA + NA + NA NA NA Not required for competitive colonization Female Ssc:CF1 mice (streptomycin treated) 75
    CFT073 NA NA NA + + NA NA Not required for colonization in murine urinary tract Mice, transurethral inoculation 25
    K-1 NA NA + + + NA NA Important for BMECc invasion Cell culture, BMEC 140
    O157:H7 + NA NA NA NA NA NA Important for passage in mice and shedding in calves ICR mice, calves 110
Borrelia burgdorferi NA + +/− NA OspC, DbpA Essential Female C3H/HeJ, BALB/s, and SCID mice 18, 36, 64
Burkholderia plantarii + NA NA NA NA NA NA Important for rice seedling blight but not for colonization Rice seedling leaves 126
Burkholderia pseudomallei NA + + NA NA Not required for intracellular survival Cell culture with HEp-2 and RAW264.7 cells 129
Erwinia carotovora subsp. carotovora + + + NA + NA Downregulates extracellular enzymes and Nip rpoS mutants more virulent Celery, tobacco, and potato 5, 89, 95
Legionella pneumophila + NA +/− + Mip, FliA, Icm, ProA With loss of rpoS impaired intracellular replication during early stage of infection in murine primary and human monocyte-derived macrophages Cell culture, human macrophage-like cell U397, HL-60, and monocyte-derived macrophage, THP-1 3, 6, 7, 15, 55, 56, 152
RpoS critical for growth in amoeba host and for pore-forming ability in erythrocytes Murine bone marrow-derived macrophages
Not required for survival and cytotoxicity in macrophage-like cells Acanthamoeba castellanii and A. polyphaga amoebae
Pseudomonas aeruginosa + + NA + + + Exotoxin A, alginate production rpoS mutants more virulent in rat chronic lung infection, but moderate effect of RpoS on virulence in Galleria mellonella Rat chronic lung infection with agar-bead-embedded bacteria placed in rat left lung; Galleria mellonella (wax moth) 127, 130
Ralstonia solanacearum NA + NA EPS 1, EGL, but downregulates PGL Minor effect of RpoS on virulence Tomato 45
Salmonella enterica serovar Typhimurium + + + NA NA NA SpvR, SpvABCD, and chromosome genes Essential (oral lethal dose 1,000-fold higher for rpoS mutants; CFU of wild-type-infected spleen 1,000-fold higher than that of mutants Female BALB/c and C57BL/6 mice 40, 73
Serratia entomophila NA NA NA NA AnfA1 Important for control of antifeeding effect Costelytra zealandica, larval infection 49
Shigella flexneri NA + + NA NA NA NA Not required for invasion and plaque formation Cell culture, Henle 407 92
Vibrio cholerae + + NA NA + + HA/protease dependent on RpoS, but cholera toxin downregulated by RpoS Mucosa escape response RpoS dependent; not required for intraintestinal survival in infant mice Rabbit ileal loops, infant (4-5 days old) CFW mice 101, 149
NA NA NA NA NA NA NA Required for efficient colonization 5-day-old suckling CD1 mice 90
Xenorhabdus nematophilus + + NA NA NA Required for growth in mutualistic hosts; not required for virulence in insects Mutualistic to Steinernema carpocapsae, pathogenic to Manduca sexta 139
Yersinia enterocolitica + + + + + NA Yst (enterotoxin) No effect in virulence and invasion; no difference in LD50d in mice Cell culture, Hep-2 cells, female BALB/c mice 8, 66
a

+, positive effect; −, negative effect; +/−, either positive or negative depending on strain backgrounds or growth conditions.

b

NA, information not available in the reference(s) cited.

c

BMEC, brain microvascular endothelial cells.

d

LD50, 50% lethal dose.

ENTERIC PATHOGENS

E. coli commensal strains are a common component of human intestinal flora, but there are many E. coli pathogens, including enteropathogenic E. coli (EPEC) and enterohemorrhagic E. coli (EHEC), which can cause severe gastrointestinal disease. Though the regulation of RpoS in gene expression is best studied in E. coli, RpoS involvement during enteropathogenesis is unresolved, probably due to the lack of effective animal models (97). Infection with E. coli in mice does not cause intestinal disease as it does in humans (97). However, several virulence traits are known to be controlled by RpoS. For example, the production of curli, important for colonization, is dependent on RpoS (105). RpoS also controls the expression of the ehxCABD operon, encoding enterohemolysin, in E. coli O157:H7 (83). A common characteristic of EPEC and EHEC infection is the formation of attaching and effacing (A/E) lesions, which requires expression of genes on a pathogenicity island, the locus of enterocyte effacement (LEE) (37). The LEE island harbors five polycistronic operons, which encode a type III secretion system (T3SS) and secreted proteins essential for virulence (27). The effect of RpoS on the expression of LEE genes has been studied by several independent groups, and variable results have been reported. Expression of lacZ fusions to promoters of LEE3 and EspA is higher in the wild-type K-12 strain than in isogenic rpoS mutants (13, 128). However, other studies report that RpoS is a negative regulator of LEE genes (34, 68, 77, 133). It is known that expression of LEE genes varies among E. coli species (35, 114), although the basis for this is not yet fully understood. One likely contributing factor is the sequence variation in the pch prophage adjacent genomic regions that affects expression of LEE genes in E. coli O157:H7 subpopulations (148). Expression of LEE genes is also dependent on environmental conditions (2, 71). We recently found that RpoS positively regulates expression of Ler, a LEE-encoded regulator, in stationary phase in LB media (a noninducing condition for LEE expression), but negatively regulates expression of Ler and other LEE genes under LEE induction conditions (34). Interestingly, mutations in Hfq, a small RNA chaperone protein that is required for effective RpoS translation (94), also result in elevated expression of LEE genes through posttranscriptional control (57). However, this effect is RpoS independent (57).

Because of its importance in the bacterial stress responses, RpoS may be required for E. coli to survive passage through the gastrointestinal tract. When rpoS mutants and the wild type of E. coli O157:H7 are fed to mice and calves, the recovery of the wild type in feces is much higher that that of rpoS mutants, probably due to the RpoS-regulated acid resistance response (110). RpoS also plays a role in intestinal colonization of E. coli strain BJ4 in streptomycin-treated mice (75). Colonization in mice by rpoS mutants is as high as that of wild type in separate infection (75). However, rpoS mutants can outcompete the wild type in mouse colon during coinfection, suggesting that rpoS mutants may be able to better utilize a specific limiting nutrient in colon (75).

In nonenteric E. coli pathogens, RpoS also controls expression of virulence traits. In E. coli K-1 strains that can cause neonatal meningitis, RpoS is important for invasion of brain microvascular endothelial cells, although the mechanism has yet to be identified (140). RpoS also positively controls motility and biofilm formation in uropathogenic E. coli (UPEC) strain UTI89 (76). However, mutations in rpoS have little effect on biofilm formation in UPEC strain 536 (12) or on colonization of urethra, bladder, and kidney in UPEC strain CFT073 (25).

Citrobacter rodentium is a natural murine enteropathogen closely related to E. coli, and, similar to EPEC and EHEC strains, it utilizes the LEE-encoded type III secretion system for delivery of virulence factors (97). Infection of mice using C. rodentium provides a promising alternative model to study enteropathogenesis in natural hosts (97, 143). The rpoS mutant of C. rodentium is more sensitive to heat and oxidative stress than the wild type, indicating a conserved RpoS function (30). Colonization and virulence of C. rodentium are attenuated in rpoS mutants during infection in mice (30). However, rpoS mutants outcompete the wild type during coinfection in mouse colon (30). In contrast to the negative regulation of LEE genes by RpoS in E. coli, RpoS has a moderate yet positive effect on expression of LEE genes in C. rodentium (30).

Salmonella serovars can cause severe systemic infection (typhoid fever) or nonsystemic gastroenteritis, depending on the serotypes (103). The systemic infection of S. enterica serovar Typhimurium in mice resembles the severe infection of S. enterica serovar Typhi in humans causing typhoid fever (103). RpoS plays a critical role in Salmonella virulence (40). Specifically, RpoS is important for persistence in lymphoid organs, such as the spleen (24, 73) and liver (24), and for initial stages of infection in murine Peyer's patches (24, 100). RpoS acts primarily through positive regulation of expression of the plasmid-borne spvR and spvABCD genes, which are required for intracellular growth and systemic infection in mice and humans (40, 102). RpoS positively regulates the expression of SpvR, a LysR family regulator, which accounts for the RpoS dependence of spvABCD (1, 21, 58, 73). Interestingly, rpoS mutants are also less virulent than the plasmid-cured wild type in mouse infections, suggesting that RpoS regulates chromosomal virulence determinants as well (40). Identified chromosomal virulence factors in Salmonella include YedI, SodCII, and genes for curli synthesis. The yedI gene is RpoS dependent and is important for persistence during infection in mice by S. Typhimurium (38). The yedI mutants are impaired in competition with the wild type during oral infection and are sensitive to polymyxin B, a cationic antimicrobial peptide (38). Genes coding for curli production, csgD and csgAB, are positively regulated by RpoS in Salmonella (115). There are two sodC alleles, sodCI and sodCII, encoding superoxide dismutase in Salmonella. Most Salmonella serotypes possess sodCII (39). SodCII is controlled by RpoS and is important for virulence, likely by protecting bacteria against superoxide-dependent host defense (39, 117, 123). However, other studies show that only SodCI, but not SodCII, contributes to virulence (74, 138). The small RNA chaperone protein Hfq, an important RpoS regulator, also plays an essential role in virulence in Salmonella through posttranscriptional regulation of many virulence genes (122). This virulence effect, however, is largely independent of RpoS (122).

In addition to regulating virulence functions, RpoS is essential for survival against stresses, such as oxidative stress, starvation, DNA damage, and low pH, which Salmonella likely encounters during intracellular growth in host macrophages (40). In S. Typhimurium, RpoS and RpoS-regulated genes, including katE and spv, are induced after invasion of epithelial cells and macrophages (22). RpoS is also important for survival of S. Typhi in mouse peritoneal macrophages through protection from nitric oxide produced by macrophages (4). Although not required for survival of S. Typhi in the human promonocytic macrophage THP-1, RpoS is required for the effective induction of macrophage apoptosis by S. Typhi during intracellular infection (72).

Interestingly, Salmonella cells infect and grow intracellularly in cultured epithelial and macrophage cells but not fibroblasts and other nonphagocytic cells (88). Using a random mutagenesis strategy, Cano and colleagues have found that mutations in genes phoP/Q, rpoS, spvR, and spvB can allow for growth in fibroblasts (NRK-49F rat kidney cell line) (20). This growth repression in fibroblasts by these genes is likely restricted to specific cell lines, since mutations in phoP/Q result in enhanced growth in the 3T3 mouse fibroblast cell line, but not in HeLa cells (20). The viable but not growing intracellular state in fibroblasts could conceivably aid in bacterial persistence within infected nonphagocytic cells (20).

Because of the virulence deficiency of rpoS mutants, these mutants are potential candidate vaccine agents (24, 26). However, the potential of using rpoS mutants as a vaccine is serotype dependent (23). Protection from infection of wild-type S. enterica serovar Dublin can be achieved with preinoculation of rpoS mutants of S. Dublin but not with a heterologous preparation made from S. Typhimurium (23).

During outbreaks, Salmonella spreads through contaminated food sources, including vegetables. In an alfalfa sprout model, the S. enterica serovar Newport wild-type strain colonizes the plant tissue much better than rpoS mutants by 24 h, although the number of cells reaches a similar level after 48 h (9). Interestingly, studies with the plant pathogens Erwinia carotovora (5) and Pseudomonas putida (91) have shown that RpoS is important for colonization on tobacco, bean, and cucumber.

Vibrio cholerae is another major food-borne human pathogen. During infection, V. cholerae adheres to the epithelial cells in the small intestine and secretes enterotoxins to disrupt ion transport of attached cells, resulting in severe diarrhea (42). RpoS mutants are impaired in survival under starvation, osmotic shock, and oxidative stress in V. cholerae (149). The hemagglutinin (HA)/protease that processes cholera enterotoxins is positively controlled by RpoS (149). Though HA/protease is not required for colonization and virulence in infant rabbits, it may allow V. cholerae to detach from epithelial cells to be released into the environment (44).

RpoS is required for efficient colonization of V. cholerae in suckling CD1 mice (90). However, another study reports that, after coinfection with wild-type V. cholerae in infant mice, the proportion of rpoS mutants remains stable by 20 h, indicating that RpoS is not required for intestinal survival (149). This difference has been attributed to strain variation within V. cholerae, which will require further study (90).

The last phase of Vibrio infection when cells detach from epithelial layers is termed the “mucosa escape response,” and this phase requires the expression of RpoS (101). The expression of genes required for motility and chemotaxis is upregulated by RpoS in the mucosa escape response and in stationary phase (101). Under in vitro virulence-inducing conditions, production of cholera toxin is 10- to 100-fold higher in the rpoS mutants than in the wild type, and virulence genes, including aphA, toxT, and vpsA, are expressed significantly higher in the rpoS mutants (101). Thus, it is likely that, during the last phase of infection, RpoS represses virulence gene expression and stimulates motility to facilitate transmission (101).

Vibrio vulnificus is a human pathogen that can cause wound infections and septicemia. RpoS protects cells from many stress conditions, except for heat shock (65). RpoS positively regulates the production of extracellular enzymes, such as albuminase, caseinase, and elastase, which may be required for survival of bacteria under many environmental conditions and for host adaptation (65). RpoS is also required for full motility (65). Interestingly, the catalase HPI is controlled by RpoS in V. vulnificus, while the gene encoding catalase HPII, which is highly RpoS dependent in E. coli, is not expressed (107).

Vibrio anguillarum is the causative agent of vibriosis in fish (84). A gene encoding the essential virulence factor EmpA metalloprotease is positively regulated by RpoS (28). The virulence of rpoS mutants is severely reduced in zebra fish (87). Similar to V. vulnificus, the V. anguillarum rpoS mutants are also impaired in production of extracellular enzymes, including phospholipase, diastase, lipase, caseinase, hemolysin, catalase, and protease (87).

Yersinia enterocolitica is an invasive enteropathogen that causes gastroenteritis in humans. Adherence and invasion of Y. enterocolitica initiate at the terminal ileum. RpoS positively regulates the expression of Yst enterotoxin (66), but does not control the expression of inv and ail, two virulence genes that are also required for invasion (8). RpoS has little effect in invasion in cell culture and in virulence of Y. enterocolitica in mouse models (8, 66).

Shigella flexneri infection causes severe dysentery in humans (118). After adherence and invasion of the colon mucous epithelial layer, S. flexneri is engulfed in phagocytic vacuoles (118). Following the lysis of these vacuoles, S. flexneri replicates and spreads to adjacent cells (118). As expected, RpoS is critical for resistance to acidic and oxidative stress in S. flexneri (92). When an rpoS mutant allele of E. coli was introduced to S. flexneri by P1 transduction, the resultant mutant exhibited no defect in invasion and formation of plaques on cultured Henle 407 cell monolayers, indicating that RpoS is not required for intercellular proliferation and spreading (92). However, the invasive ability of S. flexneri rpoS mutants has yet to be tested in animal models.

Overall, RpoS and its regulated genes are important for stress resistance and adaptation in enteric pathogens. Although RpoS plays an unequivocal role in the virulence of Salmonella species, the requirement for RpoS in the virulence and/or host adaptation in other species remains elusive. Nevertheless, given the importance of RpoS in adaptation, mutants of RpoS may be impaired in transmission to hosts due to reduced survival under adverse conditions. However, this has yet to be confirmed in animal models.

RESPIRATORY PATHOGENS

Pseudomonas aeruginosa is an opportunistic pathogen that causes chronic lung infection (104). RpoS is highly expressed in P. aeruginosa isolated from sputum samples of cystic fibrosis (CF) patients with chronic lung infection (46). As is the case in E. coli, RpoS is critical for survival of P. aeruginosa under osmotic shock, heat shock, and oxidative stress conditions (130). The effect of RpoS on expression of known virulence factors varies. For example, RpoS positively regulates the production of exotoxin A, which inhibits eukaryotic protein synthesis, and alginate, an important factor in the persistence of P. aeruginosa in CF lung and evasion of phagocytosis (127, 130). The secreted protease activities of elastase and LasA are also reduced in rpoS mutants (130). However, the production of pyocyanin, a virulence secondary metabolite that interferes with host immune defense response (136), is enhanced in rpoS mutants (130). In a rat chronic-infection model that specifically assesses the effect of extracellular secreted virulence proteins, rpoS mutants survive as well as the wild type but cause more damage to lung tissues, which may be attributable to excess pyocyanin production (130). RpoS is required for full motility of Pseudomonas and thus has been suggested to be important for colonization (130).

The RpoS translational regulator Hfq is critical for virulence of P. aeruginosa O1 in the wax moth Gelleria mellonella and in mice (127), while RpoS only has a moderate virulence effect in G. mellonella (127). Production of pyocyanin is negatively controlled by Hfq and RpoS (127). RpoS has little effect on motility of P. aeruginosa (127), which differs from previous results (130), probably due to differences in testing conditions.

The role of RpoS in quorum sensing of P. aeruginosa remains elusive. It has been shown that transcription of rpoS is controlled by quorum-sensing regulators, LasR and RhlR (81), while another study reports that quorum sensing has little effect on expression of RpoS and is in fact repressed by RpoS (142). The basis for these conflicting effects is unknown. Nonetheless, there certainly is overlapping regulation between regulons of quorum sensing and RpoS in P. aeruginosa. For example, the production of cytotoxic lectins is controlled by both RpoS and the quorum-sensing regulator RhlR (144).

Legionella pneumophila is a facultative intracellular pathogen that can cause severe pneumonia, named “Legionnaires' disease” (131). A natural reservoir for L. pneumophila is a wide range of amoebae living in soil and water sources (43). L. pneumophila is transmitted to the human respiratory tract through contaminated water aerosols (131). During phagocytosis, L. pneumophila engulfed in phagosomes initially suppresses virulence traits until entry into stationary phase, when virulence and transmission traits are activated to stimulate transmission to adjacent cells (131).

RpoS plays a critical role in regulation of transmission and virulence of L. pneumophila (3, 7). Transcription of rpoS peaks in exponential phase, while the protein level of RpoS reaches maximum in postexponential phase (7). RpoS is important for survival in osmotic shock but not other stress conditions in exponential phase (55). In stationary phase, though cells become more stress resistant, RpoS is dispensable (55).

In exponential phase, RpoS downregulates the transcription of L. pneumophila virulence genes csrA, letE, fliA, and flaA and represses motility, infectivity, and cytotoxicity (6, 7). However, in postexponential phase, RpoS is critical for the transcription of flagellar genes fliA and flaA (7). The repression of traits for transmission and cytotoxicity by RpoS in exponential phase may be important to allow cell replication to a high level, while in stationary phase, RpoS repression is relieved and the transmission traits are upregulated by RpoS (7).

The pathogenesis of L. pneumophila requires the virulence factor Mip, a peptidyl-prolyl isomerase, for invasion and replication within both amoebae and macrophages (7). The transcription of mip is severely impaired in postexponential phase rpoS mutants (7). Production of phospholipase and lipophospholipase, two virulence factors, is also under positive control of RpoS (15). In addition, RpoS positively regulates expression of ProA, a secreted virulence protease that is cytotoxic to macrophages and is important for virulence in a guinea pig model (15). RpoS also regulates the expression of the ankyrin genes that play a critical role in intracellular growth within amoeba hosts and human macrophages (54). A pleiotropic regulator, LqsR, is RpoS dependent (132). LqsR-regulated genes are involved in virulence, motility, and cell division, and mutations in lqsR result in attenuated growth in macrophages and the protozoan hosts Acanthamoeba castellanii and Dictyostelium discoideum (132). RpoS may also contribute to blocking phagolysosome formation by preventing the accumulation of LAMP-1, a phagolysomal protein (6). RpoS is crucial for the pore-forming activity of L. pneumophila and adaptation to phagosomal intracellular environments during infection (3).

The expression of the icm and dot genes, encoding the Icm/Dot type IV secretion system in L. pneumophila, is required for cytotoxicity and intracellular replication within macrophages and for intracellular growth in the protozoan host Acanthamoeba castellanii (152). RpoS only has a minor effect on the expression of the Icm/Dot genes (63, 152). However, many genes encoding Icm/Dot secreted proteins require RpoS for full expression (63).

The potential involvement of RpoS in invasion of cell cultures likely depends on the characteristics of macrophages (6). The intracellular environment is likely more deleterious to bacteria in primary macrophages than that in macrophage-like cells (6). L. pneumophila requires RpoS for efficient replication in the protozoan hosts A. castellanii (55) and Acanthamoeba polyphaga (3) and in murine bone marrow-derived macrophages (6) and human monocyte-derived macrophages (3). However, RpoS is not required for replication in cultured human macrophage-like HL-60 and THP-1-derived cells (55). In murine bone marrow-derived macrophages, most rpoS mutants, except for a small subpopulation (∼5%), cannot replicate within infected vacuoles during initial infection in the first 48 h (6). However, rpoS mutants can grow to wild-type levels after 72 h (6).

Burkholderia pseudomallei, a member of the Betaproteobacteria, is the causative agent of melioidosis. B. pseudomallei can invade host cells and induce the formation of a multinucleated giant cell (MNGC) by cell fusion (137). B. pseudomallei requires RpoS for resistance to stresses including starvation, oxidative stress, and acidic conditions, but not to osmotic shock and heat exposure (129). RpoS is not involved in invasion of cultured human epithelial cells HEp-2 and murine macrophage RAW264.7 (129). However, another study reports that RpoS is important in invasion of RAW264.7 cells but not required for intracellular replication after invasion (137). The reason for this difference is not known. Survival of rpoS mutants in gamma interferon (IFN-γ)-activated macrophages is severely impaired in comparison to that of wild-type cells (137). RpoS mutants cannot induce MNGC formation that is important for B. pseudomallei to spread to neighboring cells (137).

To summarize, in the respiratory pathogens L. pneumophila and B. pseudomallei, RpoS-regulated genes are important for survival within the intracellular environment, although this appears to be also dependent on cell lines. In P. aeruginosa, the virulence effect of RpoS is not conclusive. RpoS positively regulates expression of extracellular enzymes but negatively affects production of the virulence factor pyocyanin. Whether RpoS controls colonization and virulence needs to be further tested in animal models.

LYME DISEASE SPIROCHETE

Borrelia burgdorferi, the Lyme-disease-causing bacterium, is readily transmitted between arthropod and mammalian hosts. In contrast to proteobacteria, RpoS in B. burgdorferi is not important for resistance under most stress conditions, except for hyperosmolarity (36) and low pH (18). RpoS is induced in stationary phase, low pH, and during temperature shift from 23°C to 37°C (18, 146). The induction of RpoS is controlled by RpoN and an associated activator, Rrp2 (16, 64, 125, 147). Two-dimensional gel analysis reveals that RpoS controls the expression of a group of proteins in stationary phase (36). RpoS is essential for virulence of B. burgdorferi in mouse models (18). RpoS positively regulates expression of the ospC gene (50, 64), encoding an outer surface lipoprotein critical for virulence (120). The expression of RpoS regulon in vivo is modulated by mammalian host signals, since transcriptome analysis shows that many genes regulated by RpoS are only expressed in vivo within dialysis chambers (19).

INSECT PATHOGENS

Serratia entomophila is a soilborne pathogen that causes amber disease and general septicemia lethal to the grass grub, Costelytra zealandica (49). S. entomophila appears to have only one catalase, whose expression is RpoS independent (49). Both the wild type and rpoS mutants are sensitive to acid conditions (49). RpoS positively regulates the expression of anfA1, which codes for an important virulence factor during the development of larval infection (49).

Xenorhabdus nematophilus, a member of the Gammaproteobacteria, is mutualistic to Steinernema carpocapsae nematodes but pathogenic to many insects (e.g., Manduca sexta). RpoS is important for survival upon exposure to H2O2 but not to osmotic stress (139). In addition, rpoS mutants survive longer than the wild type in long-term batch cultures. The rpoS gene is required for colonization in the mutualistic host, S. carpocapsae nematodes, but not for virulence in insects (139).

PLANT PATHOGENS

The plant pathogen Erwinia carotovora requires RpoS for survival under stresses including starvation, acidic pH, and exposure to H2O2 (95). RpoS mutants are more virulent during infection in celery and tobacco but not potato (5, 95). The expression of a virulence factor, Nip (necrosis-inducing protein), is enhanced in rpoS mutants (89, 95). RpoS also negatively regulates the production of extracellular enzymes, pectate lyase, polygalacturonase, and cellulase, which are important for degradation of plant cell wall during infection (89, 95). This negative regulation is probably mediated through the RpoS-dependent gene rsmA, encoding a repressor for extracellular enzymes (95). A competition study shows that rpoS mutants cannot outcompete the wild type in vitro or in planta in tobacco (5).

Burkholderia plantarii is a plant pathogen that can cause rice seedling blight, and its rpoS mutants show a severe defect in pathogenesis (126). Since RpoS mutants colonize rice leaves as well as wild-type cells, this virulence defect is likely due to control of virulence traits by RpoS (126). However, these RpoS-regulated virulence traits have not been identified.

Ralstonia (previously Pseudomonas) solanacearum is a soilborne phytopathogen that can cause lethal vascular wilt disease in plants, with a wide host range (47). Survival of rpoS mutants is impaired in acid and starvation but not in heat, oxidative, or high-osmolarity conditions (45). The production of extracellular polysaccharide and activity of endogluconase, two known virulence factors (69, 113), are attenuated in rpoS mutants, while the polygalacturonase (PGL) activity is elevated in rpoS mutants (45). Tomatoes infected with rpoS mutants show delayed wilting of leaves compared with plants infected with wild-type P. solanacearum, indicating an attenuation in virulence (45). RpoS is also important for the production of the quorum-sensing autoinducer acylhomoserine lactone (45).

RpoS AS A NICHE-ADAPTATION REGULATOR

Given that RpoS is found in bacteria that occupy very different environments, a natural question arising is “Are there common requirements for gene expression among bacteria that invade different hosts?” Common factors during host adaptation include slow growth as the pathogen adapts to new nutrient sources and possible exposure to host defense mechanisms, including oxidative and acid stress components. These factors are controlled by RpoS regardless of the nature of the host. However, there are also many specific functions that may only be required on an episodic basis for host adaptation and colonization that are not required in other environments. Thus, adhesion factors, extracellular enzymes including lipases, proteases, and sugar-metabolizing functions also may be dependent on RpoS when bacteria are experiencing suboptimal growth conditions. While RpoS is conserved across several genera of the Proteobacteria, the species-specific nature of the regulon can vary considerably. RpoS-controlled genes, although important for full virulence in specific cases, are invariably nonessential genes and thus likely do not have to be expressed under conditions in which RpoS itself is at basal levels, as is the case during exponential growth in nutrient-rich environments. As a result, the regulatory environment in which RpoS-controlled genes exist is fairly permissive. It is possible that horizontally transferred genes, which may enhance host adaptation, might easily integrate into the suboptimally expressed regulon controlled by RpoS. Phylogenetic studies examining the evolutionary relationship of RpoS to the broad class of genes that it controls will be necessary to resolve this question.

CONCLUDING REMARKS

Bacterial pathogenesis is a multifaceted process that requires concerted expression of not only specific virulence factors but also genes encoding other cellular functions, including metabolism and adaptation. As a transcription regulator, RpoS can mediate virulence either directly by controlling expression of virulence factors or indirectly by stimulating the general adaptation response to enhance survival of pathogens in hostile host environments. Since expression of RpoS is tightly controlled by environmental signals, including those specific to infection (e.g., intracellular infection of Salmonella and Legionella), RpoS may be viewed as a transient regulator that allows expression of specific genes to quickly respond to environmental stimuli. The RpoS regulons identified in different bacteria also vary substantially (33, 34, 63, 119, 141). It is possible that, from an evolutionary point of view, RpoS has evolved to modulate temporal expression of specific genes whose expression is only transiently required, such as those genes for host adaptation or genes for adaptation to episodic environmental stresses (e.g., high osmolarity and oxidative stress).

The interaction between a pathogen and its host is complex, having discrete infection stages, including entry, attachment, colonization, and dispersal. Regulators may have roles in one or more of these steps, and each of these must be considered in a complete evaluation of RpoS as a potential virulence factor. In addition, choice of animal models to study regulatory factors may markedly affect the results observed.

Given that RpoS expression and activity are regulated at multiple levels by a number of other regulatory factors (60), expression of genes under the control of RpoS should be considered not as an isolated event but rather a result of complex regulatory interaction between RpoS and other regulators, including H-NS (10), ppGpp (10, 48), and Hfq (94). For instance, H-NS, a nucleoid-associated DNA-binding protein (41, 98), controls the expression of a large number of genes in Salmonella (86, 99) and E. coli (51, 62, 106). By direct binding to AT-rich regions, H-NS represses expression of virulence genes, including the plasmid-borne spv virulence region and all five chromosomal pathogenicity islands in Salmonella (41, 99), the LEE pathogenicity island in enteropathogenic E. coli (17, 53), and all major virulence regions in uropathogenic E. coli (96). Interestingly, H-NS-controlled genes can be derepressed and transcribed by RpoS-associated RNA polymerase (10, 41, 98, 105, 121). Therefore, the episodic functions of RpoS may be required to allow transcription of genes repressed by H-NS or H-NS homologs, such as Ler (134) and StpA (85). In Salmonella, a third of StpA-repressed genes are under positive control of RpoS for expression (85). In addition to the functional antirepression relationship, H-NS and StpA also negatively control RpoS translation and stability (10, 85, 145, 151).

Recent insights from genomic expression profiling studies has expanded our understanding of RpoS from a particular stress regulator to a second vegetative sigma factor that has a much broader physiological function (29, 33, 111, 141). Nevertheless, we still do not know the function of a large number of RpoS-regulated genes, even in the well-studied model organism E. coli. In many pathogens known to require RpoS for full virulence, the exact mechanism has not been identified. Characterization of the RpoS regulon in these pathogens may provide valuable insights. The RpoS regulon varies substantially between even closely related pathogens, which may reflect modulation by other regulatory factors. One example of a known factor is the Crl protein, which regulates RpoS activity by direct interaction in both E. coli and Salmonella (14, 82, 112, 135). How Crl and similar regulatory factors may interact in other RpoS-expressing pathogens has not been examined. Therefore, both genomic and functional approaches are required to advance our understanding of the role that RpoS plays in bacterial pathogenesis and related cellular functions.

Acknowledgments

This study was supported by grants from the Natural Sciences and Engineering Council of Canada (NSERC) and the Canadian Institutes of Health Research (CIHR) to H.E.S. T.D. is a recipient of an Ontario Graduate Scholarship.

We thank C. Joyce and S. M. Chiang for reviewing the manuscript. We are also grateful to the anonymous reviewers for helpful suggestions.

Biography

Inline graphicTao Dong was born and raised in a small city, Heze, in Shandong, China. After obtaining his bachelor's degree in biochemistry at Shandong University, he moved to Hamilton, Canada, to pursue his Ph.D. in biology at McMaster University. His research focuses on the molecular mechanisms of bacterial pathogenesis and adaptation. He is especially interested in understanding how bacteria adapt to stress conditions that they experience in natural environments and during host infection. He is currently supported by an Ontario Graduate Scholarship for his study. He enjoys playing badminton and squash for recreation.

Inline graphicHerb Schellhorn was born in Toronto, Canada, and was educated in schools in southern Ontario. After completing his undergraduate and M.S. degrees at the University of Guelph, Ontario, he moved to North Carolina, where he conducted his Ph.D. research under H. Hassan at North Carolina State University on catalase regulation in Escherichia coli. After postdoctoral study with K. Brooks Low at Yale University, he accepted a position at McMaster University. He is currently a full professor in the Department of Biology. The main focus of his laboratory is the study of bacterial gene regulation during adaptation to the host environment and to environmental stress. He has also recently developed an interest in mammalian vitamin C metabolism and its role in aging and disease. His recreational interests include squash and skiing.

Editor: J. B. Kaper

Footnotes

Published ahead of print on 30 November 2009.

REFERENCES

  • 1.Abe, A., H. Matsui, H. Danbara, K. Tanaka, H. Takahashi, and K. Kawahara. 1994. Regulation of spvR gene expression of Salmonella virulence plasmid pKDSC50 in Salmonella choleraesuis serovar Choleraesuis. Mol. Microbiol. 12:779-787. [DOI] [PubMed] [Google Scholar]
  • 2.Abe, H., I. Tatsuno, T. Tobe, A. Okutani, and C. Sasakawa. 2002. Bicarbonate ion stimulates the expression of locus of enterocyte effacement-encoded genes in enterohemorrhagic Escherichia coli O157:H7. Infect. Immun. 70:3500-3509. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 3.Abu-Zant, A., R. Asare, J. E. Graham, and K. Y. Abu. 2006. Role for RpoS but not RelA of Legionella pneumophila in modulation of phagosome biogenesis and adaptation to the phagosomal microenvironment. Infect. Immun. 74:3021-3026. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 4.Alam, M. S., M. H. Zaki, J. Yoshitake, T. Akuta, T. Ezaki, and T. Akaike. 2006. Involvement of Salmonella enterica serovar Typhi RpoS in resistance to NO-mediated host defense against serovar Typhi infection. Microb. Pathog. 40:116-125. [DOI] [PubMed] [Google Scholar]
  • 5.Andersson, R. A., V. Koiv, C. Norman-Setterblad, and M. Pirhonen. 1999. Role of RpoS in virulence and stress tolerance of the plant pathogen Erwinia carotovora subsp. carotovora. Microbiology 145:3547-3556. [DOI] [PubMed] [Google Scholar]
  • 6.Bachman, M. A., and M. S. Swanson. 2001. RpoS co-operates with other factors to induce Legionella pneumophila virulence in the stationary phase. Mol. Microbiol. 40:1201-1214. [DOI] [PubMed] [Google Scholar]
  • 7.Bachman, M. A., and M. S. Swanson. 2004. Genetic evidence that Legionella pneumophila RpoS modulates expression of the transmission phenotype in both the exponential phase and the stationary phase. Infect. Immun. 72:2468-2476. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 8.Badger, J. L., and V. L. Miller. 1995. Role of RpoS in survival of Yersinia enterocolitica to a variety of environmental stresses. J. Bacteriol. 177:5370-5373. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 9.Barak, J. D., L. Gorski, P. Naraghi-Arani, and A. O. Charkowski. 2005. Salmonella enterica virulence genes are required for bacterial attachment to plant tissue. Appl. Environ. Microbiol. 71:5685-5691. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 10.Barth, M., C. Marschall, A. Muffler, D. Fischer, and R. Hengge-Aronis. 1995. Role for the histone-like protein H-NS in growth phase-dependent and osmotic regulation of sigma S and many sigma S-dependent genes in Escherichia coli. J. Bacteriol. 177:3455-3464. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 11.Becker, G., E. Klauck, and R. Hengge-Aronis. 1999. Regulation of RpoS proteolysis in Escherichia coli: the response regulator RssB is a recognition factor that interacts with the turnover element in RpoS. Proc. Natl. Acad. Sci.U. S. A. 96:6439-6444. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 12.Beloin, C., K. Michaelis, K. Lindner, P. Landini, J. Hacker, J. M. Ghigo, and U. Dobrindt. 2006. The transcriptional antiterminator RfaH represses biofilm formation in Escherichia coli. J. Bacteriol. 188:1316-1331. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 13.Beltrametti, F., A. U. Kresse, and C. A. Guzman. 1999. Transcriptional regulation of the esp genes of enterohemorrhagic Escherichia coli. J. Bacteriol. 181:3409-3418. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 14.Bougdour, A., C. Lelong, and J. Geiselmann. 2004. Crl, a low temperature-induced protein in Escherichia coli that binds directly to the stationary phase sigma subunit of RNA polymerase. J. Biol. Chem. 279:19540-19550. [DOI] [PubMed] [Google Scholar]
  • 15.Broich, M., K. Rydzewski, T. L. McNealy, R. Marre, and A. Flieger. 2006. The global regulatory proteins LetA and RpoS control phospholipase A, lysophospholipase A, acyltransferase, and other hydrolytic activities of Legionella pneumophila JR32. J. Bacteriol. 188:1218-1226. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 16.Burtnick, M. N., J. S. Downey, P. J. Brett, J. A. Boylan, J. G. Frye, T. R. Hoover, and F. C. Gherardini. 2007. Insights into the complex regulation of rpoS in Borrelia burgdorferi. Mol. Microbiol. 65:277-293. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 17.Bustamante, V. H., F. J. Santana, E. Calva, and J. L. Puente. 2001. Transcriptional regulation of type III secretion genes in enteropathogenic Escherichia coli: Ler antagonizes H-NS-dependent repression. Mol. Microbiol. 39:664-678. [DOI] [PubMed] [Google Scholar]
  • 18.Caimano, M. J., C. H. Eggers, K. R. Hazlett, and J. D. Radolf. 2004. RpoS is not central to the general stress response in Borrelia burgdorferi but does control expression of one or more essential virulence determinants. Infect. Immun. 72:6433-6445. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 19.Caimano, M. J., R. Iyer, C. H. Eggers, C. Gonzalez, E. A. Morton, M. A. Gilbert, I. Schwartz, and J. D. Radolf. 2007. Analysis of the RpoS regulon in Borrelia burgdorferi in response to mammalian host signals provides insight into RpoS function during the enzootic cycle. Mol. Microbiol. 65:1193-1217. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 20.Cano, D. A., M. Martinez-Moya, M. G. Pucciarelli, E. A. Groisman, J. Casadesus, and F. Garcia-del Portillo. 2001. Salmonella enterica serovar Typhimurium response involved in attenuation of pathogen intracellular proliferation. Infect. Immun. 69:6463-6474. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 21.Chen, C. Y., N. A. Buchmeier, S. Libby, F. C. Fang, M. Krause, and D. G. Guiney. 1995. Central regulatory role for the RpoS sigma factor in expression of Salmonella dublin plasmid virulence genes. J. Bacteriol. 177:5303-5309. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 22.Chen, C. Y., L. Eckmann, S. J. Libby, F. C. Fang, S. Okamoto, M. F. Kagnoff, J. Fierer, and D. G. Guiney. 1996. Expression of Salmonella typhimurium rpoS and rpoS-dependent genes in the intracellular environment of eukaryotic cells. Infect. Immun. 64:4739-4743. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 23.Coynault, C., and F. Norel. 1999. Comparison of the abilities of Salmonella typhimurium rpoS, aroA and rpoS aroA strains to elicit humoral immune responses in BALB/c mice and to cause lethal infection in athymic BALB/c mice. Microb. Pathog. 26:299-305. [DOI] [PubMed] [Google Scholar]
  • 24.Coynault, C., V. Robbe-Saule, and F. Norel. 1996. Virulence and vaccine potential of Salmonella typhimurium mutants deficient in the expression of the RpoS (sigma S) regulon. Mol. Microbiol. 22:149-160. [DOI] [PubMed] [Google Scholar]
  • 25.Culham, D. E., A. Lu, M. Jishage, K. A. Krogfelt, A. Ishihama, and J. M. Wood. 2001. The osmotic stress response and virulence in pyelonephritis isolates of Escherichia coli: contributions of RpoS, ProP, ProU and other systems. Microbiology 147:1657-1670. [DOI] [PubMed] [Google Scholar]
  • 26.Curtiss, R., III, S. Y. Wanda, B. M. Gunn, X. Zhang, S. A. Tinge, V. Ananthnarayan, H. Mo, S. Wang, and W. Kong. 2009. Salmonella enterica serovar Typhimurium strains with regulated delayed attenuation in vivo. Infect. Immun. 77:1071-1082. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 27.Deng, W., J. L. Puente, S. Gruenheid, Y. Li, B. A. Vallance, A. Vazquez, J. Barba, J. A. Ibarra, P. O'Donnell, P. Metalnikov, K. Ashman, S. Lee, D. Goode, T. Pawson, and B. B. Finlay. 2004. Dissecting virulence: systematic and functional analyses of a pathogenicity island. Proc. Natl. Acad. Sci.U. S. A. 101:3597-3602. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 28.Denkin, S. M., and D. R. Nelson. 2004. Regulation of Vibrio anguillarum empA metalloprotease expression and its role in virulence. Appl. Environ. Microbiol. 70:4193-4204. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 29.Dong, T., S. M. Chiang, C. Joyce, R. Yu, and H. E. Schellhorn. 2009. Polymorphism and selection of rpoS in pathogenic Escherichia coli. BMC Microbiol. 9:118. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 30.Dong, T., B. K. Coombes, and H. E. Schellhorn. 2009. Role of RpoS in the virulence of Citrobacter rodentium. Infect. Immun. 77:501-507. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 31.Dong, T., C. Joyce, and H. E. Schellhorn. 2008. The role of RpoS in bacterial adaptation, p. 313-337. In W. M. El-Sharoud (ed.), Bacterial physiology—a molecular approach. Springer, Berlin, Germany.
  • 32.Dong, T., M. G. Kirchhof, and H. E. Schellhorn. 2008. RpoS regulation of gene expression during exponential growth of Escherichia coli K12. Mol. Genet. Genomics 279:267-277. [DOI] [PubMed] [Google Scholar]
  • 33.Dong, T., and H. E. Schellhorn. 2009. Control of RpoS in global gene expression of Escherichia coli in minimal media. Mol. Genet. Genomics 281:19-33. [DOI] [PubMed] [Google Scholar]
  • 34.Dong, T., and H. E. Schellhorn. 2009. Global effect of RpoS on gene expression in pathogenic Escherichia coli O157:H7 strain EDL933. BMC Genomics 10:349. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 35.Dowd, S. E., and H. Ishizaki. 2006. Microarray based comparison of two Escherichia coli O157:H7 lineages. BMC Microbiol. 6:30. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 36.Elias, A. F., J. L. Bono, J. A. Carroll, P. Stewart, K. Tilly, and P. Rosa. 2000. Altered stationary-phase response in a Borrelia burgdorferi rpoS mutant. J. Bacteriol. 182:2909-2918. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 37.Elliott, S. J., L. A. Wainwright, T. K. McDaniel, K. G. Jarvis, Y. K. Deng, L. C. Lai, B. P. McNamara, M. S. Donnenberg, and J. B. Kaper. 1998. The complete sequence of the locus of enterocyte effacement (LEE) from enteropathogenic Escherichia coli E2348/69. Mol. Microbiol. 28:1-4. [DOI] [PubMed] [Google Scholar]
  • 38.Erickson, K. D., and C. S. Detweiler. 2006. The Rcs phosphorelay system is specific to enteric pathogens/commensals and activates ydeI, a gene important for persistent Salmonella infection of mice. Mol. Microbiol. 62:883-894. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 39.Fang, F. C., M. A. DeGroote, J. W. Foster, A. J. Baumler, U. Ochsner, T. Testerman, S. Bearson, J. C. Giard, Y. Xu, G. Campbell, and T. Laessig. 1999. Virulent Salmonella typhimurium has two periplasmic Cu, Zn-superoxide dismutases. Proc. Natl. Acad. Sci.U. S. A. 96:7502-7507. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 40.Fang, F. C., S. J. Libby, N. A. Buchmeier, P. C. Loewen, J. Switala, J. Harwood, and D. G. Guiney. 1992. The alternative sigma factor katF (rpoS) regulates Salmonella virulence. Proc. Natl. Acad. Sci.U. S. A. 89:11978-11982. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 41.Fang, F. C., and S. Rimsky. 2008. New insights into transcriptional regulation by H-NS. Curr. Opin. Microbiol. 11:113-120. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 42.Faruque, S. M., M. J. Albert, and J. J. Mekalanos. 1998. Epidemiology, genetics, and ecology of toxigenic Vibrio cholerae. Microbiol. Mol. Biol. Rev. 62:1301-1314. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 43.Fields, B. S. 1996. The molecular ecology of Legionellae. Trends Microbiol. 4:286-290. [DOI] [PubMed] [Google Scholar]
  • 44.Finkelstein, R. A., M. Boesman-Finkelstein, Y. Chang, and C. C. Hase. 1992. Vibrio cholerae hemagglutinin/protease, colonial variation, virulence, and detachment. Infect. Immun. 60:472-478. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 45.Flavier, A. B., M. A. Schell, and T. P. Denny. 1998. An RpoS (sigmaS) homologue regulates acylhomoserine lactone-dependent autoinduction in Ralstonia solanacearum. Mol. Microbiol. 28:475-486. [DOI] [PubMed] [Google Scholar]
  • 46.Foley, I., P. Marsh, E. M. Wellington, A. W. Smith, and M. R. Brown. 1999. General stress response master regulator rpoS is expressed in human infection: a possible role in chronicity. J. Antimicrob. Chemother. 43:164-165. [DOI] [PubMed] [Google Scholar]
  • 47.Genin, S., and C. Boucher. 2004. Lessons learned from the genome analysis of Ralstonia solanacearum. Annu. Rev. Phytopathol. 42:107-134. [DOI] [PubMed] [Google Scholar]
  • 48.Gentry, D. R., V. J. Hernandez, L. H. Nguyen, D. B. Jensen, and M. Cashel. 1993. Synthesis of the stationary-phase sigma factor sigmaS is positively regulated by ppGpp. J. Bacteriol. 175:7982-7989. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 49.Giddens, S. R., A. Tormo, and H. K. Mahanty. 2000. Expression of the antifeeding gene anfA1 in Serratia entomophila requires rpoS. Appl. Environ. Microbiol. 66:1711-1714. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 50.Gilbert, M. A., E. A. Morton, S. F. Bundle, and D. S. Samuels. 2007. Artificial regulation of ospC expression in Borrelia burgdorferi. Mol. Microbiol. 63:1259-1273. [DOI] [PubMed] [Google Scholar]
  • 51.Grainger, D. C., D. Hurd, M. D. Goldberg, and S. J. Busby. 2006. Association of nucleoid proteins with coding and non-coding segments of the Escherichia coli genome. Nucleic Acids Res. 34:4642-4652. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 52.Gruber, T. M., and C. A. Gross. 2003. Multiple sigma subunits and the partitioning of bacterial transcription space. Annu. Rev. Microbiol. 57:441-466. [DOI] [PubMed] [Google Scholar]
  • 53.Haack, K. R., C. L. Robinson, K. J. Miller, J. W. Fowlkes, and J. L. Mellies. 2003. Interaction of Ler at the LEE5 (tir) operon of enteropathogenic Escherichia coli. Infect. Immun. 71:384-392. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 54.Habyarimana, F., S. Al Khodor, A. Kalia, J. E. Graham, C. T. Price, M. T. Garcia, and Y. A. Kwaik. 2008. Role for the Ankyrin eukaryotic-like genes of Legionella pneumophila in parasitism of protozoan hosts and human macrophages. Environ. Microbiol. 10:1460-1474. [DOI] [PubMed] [Google Scholar]
  • 55.Hales, L. M., and H. A. Shuman. 1999. The Legionella pneumophila rpoS gene is required for growth within Acanthamoeba castellanii. J. Bacteriol. 181:4879-4889. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 56.Hammer, B. K., E. S. Tateda, and M. S. Swanson. 2002. A two-component regulator induces the transmission phenotype of stationary-phase Legionella pneumophila. Mol. Microbiol. 44:107-118. [DOI] [PubMed] [Google Scholar]
  • 57.Hansen, A. M., and J. B. Kaper. 2009. Hfq affects the expression of the LEE pathogenicity island in enterohaemorrhagic Escherichia coli. Mol. Microbiol. 73:446-465. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 58.Heiskanen, P., S. Taira, and M. Rhen. 1994. Role of rpoS in the regulation of Salmonella plasmid virulence (spv) genes. FEMS Microbiol. Lett. 123:125-130. [DOI] [PubMed] [Google Scholar]
  • 59.Hengge-Aronis, R. 2000. The general stress response in Escherichia coli, p. 161-178. In G. Storz and R. Hengge-Aronis (ed.), Bacterial stress response. ASM Press, Washington, DC.
  • 60.Hengge-Aronis, R. 2002. Signal transduction and regulatory mechanisms involved in control of the sigma(S) (RpoS) subunit of RNA polymerase. Microbiol. Mol. Biol. Rev. 66:373-395. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 61.Hengge-Aronis, R., W. Klein, R. Lange, M. Rimmele, and W. Boos. 1991. Trehalose synthesis genes are controlled by the putative sigma factor encoded by rpoS and are involved in stationary-phase thermotolerance in Escherichia coli. J. Bacteriol. 173:7918-7924. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 62.Hommais, F., E. Krin, C. Laurent-Winter, O. Soutourina, A. Malpertuy, J. P. Le Caer, A. Danchin, and P. Bertin. 2001. Large-scale monitoring of pleiotropic regulation of gene expression by the prokaryotic nucleoid-associated protein, H-NS. Mol. Microbiol. 40:20-36. [DOI] [PubMed] [Google Scholar]
  • 63.Hovel-Miner, G., S. Pampou, S. P. Faucher, M. Clarke, I. Morozova, P. Morozov, J. J. Russo, H. A. Shuman, and S. Kalachikov. 2009. SigmaS controls multiple pathways associated with intracellular multiplication of Legionella pneumophila. J. Bacteriol. 191:2461-2473. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 64.Hubner, A., X. Yang, D. M. Nolen, T. G. Popova, F. C. Cabello, and M. V. Norgard. 2001. Expression of Borrelia burgdorferi OspC and DbpA is controlled by a RpoN-RpoS regulatory pathway. Proc. Natl. Acad. Sci.U. S. A. 98:12724-12729. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 65.Hulsmann, A., T. M. Rosche, I. S. Kong, H. M. Hassan, D. M. Beam, and J. D. Oliver. 2003. RpoS-dependent stress response and exoenzyme production in Vibrio vulnificus. Appl. Environ. Microbiol. 69:6114-6120. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 66.Iriarte, M., I. Stainier, and G. R. Cornelis. 1995. The rpoS gene from Yersinia enterocolitica and its influence on expression of virulence factors. Infect. Immun. 63:1840-1847. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 67.Ishihama, A. 2000. Functional modulation of Escherichia coli RNA polymerase. Annu. Rev. Microbiol. 54:499-518. [DOI] [PubMed] [Google Scholar]
  • 68.Iyoda, S., and H. Watanabe. 2005. ClpXP protease controls expression of the type III protein secretion system through regulation of RpoS and GrlR levels in enterohemorrhagic Escherichia coli. J. Bacteriol. 187:4086-4094. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 69.Kao, C. C., E. Barlow, and L. Sequeira. 1992. Extracellular polysaccharide is required for wild-type virulence of Pseudomonas solanacearum. J. Bacteriol. 174:1068-1071. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 70.Kazmierczak, M. J., M. Wiedmann, and K. J. Boor. 2005. Alternative sigma factors and their roles in bacterial virulence. Microbiol. Mol. Biol. Rev. 69:527-543. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 71.Kenny, B., A. Abe, M. Stein, and B. B. Finlay. 1997. Enteropathogenic Escherichia coli protein secretion is induced in response to conditions similar to those in the gastrointestinal tract. Infect. Immun. 65:2606-2612. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 72.Khan, A. Q., L. Zhao, K. Hirose, M. Miyake, T. Li, Y. Hashimoto, Y. Kawamura, and T. Ezaki. 1998. Salmonella typhi rpoS mutant is less cytotoxic than the parent strain but survives inside resting THP-1 macrophages. FEMS Microbiol. Lett. 161:201-208. [DOI] [PubMed] [Google Scholar]
  • 73.Kowarz, L., C. Coynault, V. Robbe-Saule, and F. Norel. 1994. The Salmonella typhimurium katF (rpoS) gene: cloning, nucleotide sequence, and regulation of spvR and spvABCD virulence plasmid genes. J. Bacteriol. 176:6852-6860. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 74.Krishnakumar, R., M. Craig, J. A. Imlay, and J. M. Slauch. 2004. Differences in enzymatic properties allow SodCI but not SodCII to contribute to virulence in Salmonella enterica serovar Typhimurium strain 14028. J. Bacteriol. 186:5230-5238. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 75.Krogfelt, K. A., M. Hjulgaard, K. Sorensen, P. S. Cohen, and M. Givskov. 2000. rpoS gene function is a disadvantage for Escherichia coli BJ4 during competitive colonization of the mouse large intestine. Infect. Immun. 68:2518-2524. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 76.Kulesus, R. R., K. Diaz-Perez, E. S. Slechta, D. S. Eto, and M. A. Mulvey. 2008. Impact of the RNA chaperone Hfq on the fitness and virulence potential of uropathogenic Escherichia coli. Infect. Immun. 76:3019-3026. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 77.Laaberki, M. H., N. Janabi, E. Oswald, and F. Repoila. 2006. Concert of regulators to switch on LEE expression in enterohemorrhagic Escherichia coli O157:H7: interplay between Ler, GrlA, HNS and RpoS. Int.J. Med. Microbiol. 296:197-210. [DOI] [PubMed] [Google Scholar]
  • 78.Lacour, S., and P. Landini. 2004. SigmaS-dependent gene expression at the onset of stationary phase in Escherichia coli: function of sigmaS-dependent genes and identification of their promoter sequences. J. Bacteriol. 186:7186-7195. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 79.Lange, R., and R. Hengge-Aronis. 1991. Identification of a central regulator of stationary-phase gene expression in Escherichia coli. Mol. Microbiol. 5:49-59. [DOI] [PubMed] [Google Scholar]
  • 80.Lange, R., and R. Hengge-Aronis. 1994. The cellular concentration of the sigma S subunit of RNA polymerase in Escherichia coli is controlled at the levels of transcription, translation, and protein stability. Genes Dev. 8:1600-1612. [DOI] [PubMed] [Google Scholar]
  • 81.Latifi, A., M. Foglino, K. Tanaka, P. Williams, and A. Lazdunski. 1996. A hierarchical quorum-sensing cascade in Pseudomonas aeruginosa links the transcriptional activators LasR and RhIR (VsmR) to expression of the stationary-phase sigma factor RpoS. Mol. Microbiol. 21:1137-1146. [DOI] [PubMed] [Google Scholar]
  • 82.Lelong, C., K. Aguiluz, S. Luche, L. Kuhn, J. Garin, T. Rabilloud, and J. Geiselmann. 2007. The Crl-RpoS regulon of Escherichia coli. Mol. Cell Proteomics 6:648-659. [DOI] [PubMed] [Google Scholar]
  • 83.Li, H., A. Granat, V. Stewart, and J. R. Gillespie. 2008. RpoS, H-NS, and DsrA influence EHEC hemolysin operon (ehxCABD) transcription in Escherichia coli O157:H7 strain EDL933. FEMS Microbiol. Lett. 285:257-262. [DOI] [PubMed] [Google Scholar]
  • 84.Lopez, C. S., and J. H. Crosa. 2007. Characterization of ferric-anguibactin transport in Vibrio anguillarum. Biometals 20:393-403. [DOI] [PubMed] [Google Scholar]
  • 85.Lucchini, S., P. McDermott, A. Thompson, and J. C. Hinton. 2009. The H-NS-like protein StpA represses the RpoS (sigma38) regulon during exponential growth of Salmonella Typhimurium. Mol. Microbiol. 74:1169-1186. [DOI] [PubMed] [Google Scholar]
  • 86.Lucchini, S., G. Rowley, M. D. Goldberg, D. Hurd, M. Harrison, and J. C. Hinton. 2006. H-NS mediates the silencing of laterally acquired genes in bacteria. PLoS. Pathog. 2:e81. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 87.Ma, L., J. Chen, R. Liu, X. H. Zhang, and Y. A. Jiang. 2009. Mutation of rpoS gene decreased resistance to environmental stresses, synthesis of extracellular products and virulence of Vibrio anguillarum. FEMS Microbiol. Ecol. 70:130-136. [DOI] [PubMed] [Google Scholar]
  • 88.Matsui, H., M. Eguchi, and Y. Kikuchi. 2000. Use of confocal microscopy to detect Salmonella typhimurium within host cells associated with Spv-mediated intracellular proliferation. Microb. Pathog. 29:53-59. [DOI] [PubMed] [Google Scholar]
  • 89.Mattinen, L., M. Tshuikina, A. Mae, and M. Pirhonen. 2004. Identification and characterization of Nip, necrosis-inducing virulence protein of Erwinia carotovora subsp. carotovora. Mol. Plant Microbe Interact. 17:1366-1375. [DOI] [PubMed] [Google Scholar]
  • 90.Merrell, D. S., A. D. Tischler, S. H. Lee, and A. Camilli. 2000. Vibrio cholerae requires rpoS for efficient intestinal colonization. Infect. Immun. 68:6691-6696. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 91.Miller, C. D., Y. C. Kim, and A. J. Anderson. 2001. Competitiveness in root colonization by Pseudomonas putida requires the rpoS gene. Can. J. Microbiol. 47:41-48. [DOI] [PubMed] [Google Scholar]
  • 92.Mogull, S. A., L. J. Runyen-Janecky, M. Hong, and S. M. Payne. 2001. dksA is required for intercellular spread of Shigella flexneri via an RpoS-independent mechanism. Infect. Immun. 69:5742-5751. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 93.Muffler, A., D. Fischer, S. Altuvia, G. Storz, and R. Hengge-Aronis. 1996. The response regulator RssB controls stability of the sigma(S) subunit of RNA polymerase in Escherichia coli. EMBO J. 15:1333-1339. [PMC free article] [PubMed] [Google Scholar]
  • 94.Muffler, A., D. Fischer, and R. Hengge-Aronis. 1996. The RNA-binding protein HF-I, known as a host factor for phage Qbeta RNA replication, is essential for rpoS translation in Escherichia coli. Genes Dev. 10:1143-1151. [DOI] [PubMed] [Google Scholar]
  • 95.Mukherjee, A., Y. Cui, W. Ma, Y. Liu, A. Ishihama, A. Eisenstark, and A. K. Chatterjee. 1998. RpoS (sigma-S) controls expression of rsmA, a global regulator of secondary metabolites, harpin, and extracellular proteins in Erwinia carotovora. J. Bacteriol. 180:3629-3634. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 96.Muller, C. M., U. Dobrindt, G. Nagy, L. Emody, B. E. Uhlin, and J. Hacker. 2006. Role of histone-like proteins H-NS and StpA in expression of virulence determinants of uropathogenic Escherichia coli. J. Bacteriol. 188:5428-5438. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 97.Mundy, R., T. T. MacDonald, G. Dougan, G. Frankel, and S. Wiles. 2005. Citrobacter rodentium of mice and man. Cell. Microbiol. 7:1697-1706. [DOI] [PubMed] [Google Scholar]
  • 98.Navarre, W. W., M. McClelland, S. J. Libby, and F. C. Fang. 2007. Silencing of xenogeneic DNA by H-NS-facilitation of lateral gene transfer in bacteria by a defense system that recognizes foreign DNA. Genes Dev. 21:1456-1471. [DOI] [PubMed] [Google Scholar]
  • 99.Navarre, W. W., S. Porwollik, Y. Wang, M. McClelland, H. Rosen, S. J. Libby, and F. C. Fang. 2006. Selective silencing of foreign DNA with low GC content by the H-NS protein in Salmonella. Science 313:236-238. [DOI] [PubMed] [Google Scholar]
  • 100.Nickerson, C. A., and R. Curtiss III. 1997. Role of sigma factor RpoS in initial stages of Salmonella typhimurium infection. Infect. Immun. 65:1814-1823. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 101.Nielsen, A. T., N. A. Dolganov, G. Otto, M. C. Miller, C. Y. Wu, and G. K. Schoolnik. 2006. RpoS controls the Vibrio cholerae mucosal escape response. PLoS Pathog. 2:e109. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 102.Norel, F., V. Robbe-Saule, M. Y. Popoff, and C. Coynault. 1992. The putative sigma factor KatF (RpoS) is required for the transcription of the Salmonella typhimurium virulence gene spvB in Escherichia coli. FEMS Microbiol. Lett. 78:271-276. [DOI] [PubMed] [Google Scholar]
  • 103.Ohl, M. E., and S. I. Miller. 2001. Salmonella: a model for bacterial pathogenesis. Annu. Rev. Med. 52:259-274. [DOI] [PubMed] [Google Scholar]
  • 104.Oliver, A., R. Canton, P. Campo, F. Baquero, and J. Blazquez. 2000. High frequency of hypermutable Pseudomonas aeruginosa in cystic fibrosis lung infection. Science 288:1251-1254. [DOI] [PubMed] [Google Scholar]
  • 105.Olsen, A., A. Arnqvist, M. Hammar, S. Sukupolvi, and S. Normark. 1993. The RpoS sigma factor relieves H-NS-mediated transcriptional repression of csgA, the subunit gene of fibronectin-binding curli in Escherichia coli. Mol. Microbiol. 7:523-536. [DOI] [PubMed] [Google Scholar]
  • 106.Oshima, T., S. Ishikawa, K. Kurokawa, H. Aiba, and N. Ogasawara. 2006. Escherichia coli histone-like protein H-NS preferentially binds to horizontally acquired DNA in association with RNA polymerase. DNA Res. 13:141-153. [DOI] [PubMed] [Google Scholar]
  • 107.Park, K. J., M. J. Kang, S. H. Kim, H. J. Lee, J. K. Lim, S. H. Choi, S. J. Park, and K. H. Lee. 2004. Isolation and characterization of rpoS from a pathogenic bacterium, Vibrio vulnificus: role of sigmaS in survival of exponential-phase cells under oxidative stress. J. Bacteriol. 186:3304-3312. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 108.Patten, C. L., M. G. Kirchhof, M. R. Schertzberg, R. A. Morton, and H. E. Schellhorn. 2004. Microarray analysis of RpoS-mediated gene expression in Escherichia coli K-12. Mol. Genet. Genomics 272:580-591. [DOI] [PubMed] [Google Scholar]
  • 109.Pratt, L. A., and T. J. Silhavy. 1996. The response regulator SprE controls the stability of RpoS. Proc. Natl. Acad. Sci.U. S. A. 93:2488-2492. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 110.Price, S. B., C. M. Cheng, C. W. Kaspar, J. C. Wright, F. J. DeGraves, T. A. Penfound, M. P. Castanie-Cornet, and J. W. Foster. 2000. Role of rpoS in acid resistance and fecal shedding of Escherichia coli O157:H7. Appl. Environ. Microbiol. 66:632-637. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 111.Rahman, M., M. R. Hasan, T. Oba, and K. Shimizu. 2006. Effect of rpoS gene knockout on the metabolism of Escherichia coli during exponential growth phase and early stationary phase based on gene expressions, enzyme activities and intracellular metabolite concentrations. Biotechnol. Bioeng. 94:585-595. [DOI] [PubMed] [Google Scholar]
  • 112.Robbe-Saule, V., M. D. Lopes, A. Kolb, and F. Norel. 2007. Physiological effects of Crl in Salmonella are modulated by sigmaS level and promoter specificity. J. Bacteriol. 189:2976-2987. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 113.Roberts, D. P., T. P. Denny, and M. A. Schell. 1988. Cloning of the egl gene of Pseudomonas solanacearum and analysis of its role in phytopathogenicity. J. Bacteriol. 170:1445-1451. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 114.Roe, A. J., H. Yull, S. W. Naylor, M. J. Woodward, D. G. Smith, and D. L. Gally. 2003. Heterogeneous surface expression of EspA translocon filaments by Escherichia coli O157:H7 is controlled at the posttranscriptional level. Infect. Immun. 71:5900-5909. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 115.Romling, U., Z. Bian, M. Hammar, W. D. Sierralta, and S. Normark. 1998. Curli fibers are highly conserved between Salmonella typhimurium and Escherichia coli with respect to operon structure and regulation. J. Bacteriol. 180:722-731. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 116.Sammartano, L. J., R. W. Tuveson, and R. Davenport. 1986. Control of sensitivity to inactivation by H2O2 and broad-spectrum near-UV radiation by the Escherichia coli katF (rpoS) locus. J. Bacteriol. 168:13-21. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 117.Sansone, A., P. R. Watson, T. S. Wallis, P. R. Langford, and J. S. Kroll. 2002. The role of two periplasmic copper- and zinc-cofactored superoxide dismutases in the virulence of Salmonella choleraesuis. Microbiology 148:719-726. [DOI] [PubMed] [Google Scholar]
  • 118.Schroeder, G. N., and H. Hilbi. 2008. Molecular pathogenesis of Shigella spp.: controlling host cell signaling, invasion, and death by type III secretion. Clin. Microbiol. Rev. 21:134-156. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 119.Schuster, M., A. C. Hawkins, C. S. Harwood, and E. P. Greenberg. 2004. The Pseudomonas aeruginosa RpoS regulon and its relationship to quorum sensing. Mol. Microbiol. 51:973-985. [DOI] [PubMed] [Google Scholar]
  • 120.Schwan, T. G., J. Piesman, W. T. Golde, M. C. Dolan, and P. A. Rosa. 1995. Induction of an outer surface protein on Borrelia burgdorferi during tick feeding. Proc. Natl. Acad. Sci.U. S. A. 92:2909-2913. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 121.Shin, M., M. Song, J. H. Rhee, Y. Hong, Y. J. Kim, Y. J. Seok, K. S. Ha, S. H. Jung, and H. E. Choy. 2005. DNA looping-mediated repression by histone-like protein H-NS: specific requirement of Esigma70 as a cofactor for looping. Genes Dev. 19:2388-2398. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 122.Sittka, A., V. Pfeiffer, K. Tedin, and J. Vogel. 2007. The RNA chaperone Hfq is essential for the virulence of Salmonella typhimurium. Mol. Microbiol. 63:193-217. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 123.Sly, L. M., D. G. Guiney, and N. E. Reiner. 2002. Salmonella enterica serovar Typhimurium periplasmic superoxide dismutases SodCI and SodCII are required for protection against the phagocyte oxidative burst. Infect. Immun. 70:5312-5315. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 124.Small, P., D. Blankenhorn, D. Welty, E. Zinser, and J. L. Slonczewski. 1994. Acid and base resistance in Escherichia coli and Shigella flexneri: role of rpoS and growth pH. J. Bacteriol. 176:1729-1737. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 125.Smith, A. H., J. S. Blevins, G. N. Bachlani, X. F. Yang, and M. V. Norgard. 2007. Evidence that RpoS (sigmaS) in Borrelia burgdorferi is controlled directly by RpoN (sigma54/sigmaN). J. Bacteriol. 189:2139-2144. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 126.Solis, R., I. Bertani, G. Degrassi, G. Devescovi, and V. Venturi. 2006. Involvement of quorum sensing and RpoS in rice seedling blight caused by Burkholderia plantarii. FEMS Microbiol. Lett. 259:106-112. [DOI] [PubMed] [Google Scholar]
  • 127.Sonnleitner, E., S. Hagens, F. Rosenau, S. Wilhelm, A. Habel, K. E. Jager, and U. Blasi. 2003. Reduced virulence of a hfq mutant of Pseudomonas aeruginosa O1. Microb. Pathog. 35:217-228. [DOI] [PubMed] [Google Scholar]
  • 128.Sperandio, V., J. L. Mellies, W. Nguyen, S. Shin, and J. B. Kaper. 1999. Quorum sensing controls expression of the type III secretion gene transcription and protein secretion in enterohemorrhagic and enteropathogenic Escherichia coli. Proc. Natl. Acad. Sci.U. S. A. 96:15196-15201. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 129.Subsin, B., M. S. Thomas, G. Katzenmeier, J. G. Shaw, S. Tungpradabkul, and M. Kunakorn. 2003. Role of the stationary growth phase sigma factor RpoS of Burkholderia pseudomallei in response to physiological stress conditions. J. Bacteriol. 185:7008-7014. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 130.Suh, S. J., L. Silo-Suh, D. E. Woods, D. J. Hassett, S. E. West, and D. E. Ohman. 1999. Effect of rpoS mutation on the stress response and expression of virulence factors in Pseudomonas aeruginosa. J. Bacteriol. 181:3890-3897. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 131.Swanson, M. S., and B. K. Hammer. 2000. Legionella pneumophila pathogenesis: a fateful journey from amoebae to macrophages. Annu. Rev. Microbiol. 54:567-613. [DOI] [PubMed] [Google Scholar]
  • 132.Tiaden, A., T. Spirig, S. S. Weber, H. Bruggemann, R. Bosshard, C. Buchrieser, and H. Hilbi. 2007. The Legionella pneumophila response regulator LqsR promotes host cell interactions as an element of the virulence regulatory network controlled by RpoS and LetA. Cell. Microbiol. 9:2903-2920. [DOI] [PubMed] [Google Scholar]
  • 133.Tomoyasu, T., A. Takaya, Y. Handa, K. Karata, and T. Yamamoto. 2005. ClpXP controls the expression of LEE genes in enterohaemorrhagic Escherichia coli. FEMS Microbiol. Lett. 253:59-66. [DOI] [PubMed] [Google Scholar]
  • 134.Torres, A. G., G. N. Lopez-Sanchez, L. Milflores-Flores, S. D. Patel, M. Rojas-Lopez, C. F. Martinez de la Pena, M. M. Arenas-Hernandez, and Y. Martinez-Laguna. 2007. Ler and H-NS, regulators controlling expression of the long polar fimbriae of Escherichia coli O157:H7. J. Bacteriol. 189:5916-5928. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 135.Typas, A., C. Barembruch, A. Possling, and R. Hengge. 2007. Stationary phase reorganisation of the Escherichia coli transcription machinery by Crl protein, a fine-tuner of sigmas activity and levels. EMBO J. 26:1569-1578. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 136.Ulmer, A. J., J. Pryjma, Z. Tarnok, M. Ernst, and H. D. Flad. 1990. Inhibitory and stimulatory effects of Pseudomonas aeruginosa pyocyanine on human T and B lymphocytes and human monocytes. Infect. Immun. 58:808-815. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 137.Utaisincharoen, P., S. Arjcharoen, K. Limposuwan, S. Tungpradabkul, and S. Sirisinha. 2006. Burkholderia pseudomallei RpoS regulates multinucleated giant cell formation and inducible nitric oxide synthase expression in mouse macrophage cell line (RAW 264.7). Microb. Pathog. 40:184-189. [DOI] [PubMed] [Google Scholar]
  • 138.Uzzau, S., L. Bossi, and N. Figueroa-Bossi. 2002. Differential accumulation of Salmonella[Cu, Zn] superoxide dismutases SodCI and SodCII in intracellular bacteria: correlation with their relative contribution to pathogenicity. Mol. Microbiol. 46:147-156. [DOI] [PubMed] [Google Scholar]
  • 139.Vivas, E. I., and H. Goodrich-Blair. 2001. Xenorhabdus nematophilus as a model for host-bacterium interactions: rpoS is necessary for mutualism with nematodes. J. Bacteriol. 183:4687-4693. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 140.Wang, Y., and K. S. Kim. 2000. Effect of rpoS mutations on stress-resistance and invasion of brain microvascular endothelial cells in Escherichia coli K1. FEMS Microbiol. Lett. 182:241-247. [DOI] [PubMed] [Google Scholar]
  • 141.Weber, H., T. Polen, J. Heuveling, V. F. Wendisch, and R. Hengge. 2005. Genome-wide analysis of the general stress response network in Escherichia coli: sigmaS-dependent genes, promoters, and sigma factor selectivity. J. Bacteriol. 187:1591-1603. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 142.Whiteley, M., M. R. Parsek, and E. P. Greenberg. 2000. Regulation of quorum sensing by RpoS in Pseudomonas aeruginosa. J. Bacteriol. 182:4356-4360. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 143.Wiles, S., W. P. Hanage, G. Frankel, and B. Robertson. 2006. Modelling infectious disease—time to think outside the box? Nat. Rev. Microbiol. 4:307-312. [DOI] [PubMed] [Google Scholar]
  • 144.Winzer, K., C. Falconer, N. C. Garber, S. P. Diggle, M. Camara, and P. Williams. 2000. The Pseudomonas aeruginosa lectins PA-IL and PA-IIL are controlled by quorum sensing and by RpoS. J. Bacteriol. 182:6401-6411. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 145.Yamashino, T., C. Ueguchi, and T. Mizuno. 1995. Quantitative control of the stationary phase-specific sigma factor, sigma S, in Escherichia coli: involvement of the nucleoid protein H-NS. EMBO J. 14:594-602. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 146.Yang, X., M. S. Goldberg, T. G. Popova, G. B. Schoeler, S. K. Wikel, K. E. Hagman, and M. V. Norgard. 2000. Interdependence of environmental factors influencing reciprocal patterns of gene expression in virulent Borrelia burgdorferi. Mol. Microbiol. 37:1470-1479. [DOI] [PubMed] [Google Scholar]
  • 147.Yang, X. F., S. M. Alani, and M. V. Norgard. 2003. The response regulator Rrp2 is essential for the expression of major membrane lipoproteins in Borrelia burgdorferi. Proc. Natl. Acad. Sci.U. S. A. 100:11001-11006. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 148.Yang, Z., J. Kim, C. Zhang, M. Zhang, J. Nietfeldt, C. M. Southward, M. G. Surette, S. D. Kachman, and A. K. Benson. 2009. Genomic instability in regions adjacent to a highly conserved pch prophage in Escherichia coli O157:H7 generates diversity in expression patterns of the LEE pathogenicity island. J. Bacteriol. 191:3553-3568. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 149.Yildiz, F. H., and G. K. Schoolnik. 1998. Role of rpoS in stress survival and virulence of Vibrio cholerae. J. Bacteriol. 180:773-784. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 150.Zhou, Y., and S. Gottesman. 1998. Regulation of proteolysis of the stationary-phase sigma factor RpoS. J. Bacteriol. 180:1154-1158. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 151.Zhou, Y., and S. Gottesman. 2006. Modes of regulation of RpoS by H-NS. J. Bacteriol. 188:7022-7025. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 152.Zusman, T., O. Gal-Mor, and G. Segal. 2002. Characterization of a Legionella pneumophila relA insertion mutant and roles of RelA and RpoS in virulence gene expression. J. Bacteriol. 184:67-75. [DOI] [PMC free article] [PubMed] [Google Scholar]

Articles from Infection and Immunity are provided here courtesy of American Society for Microbiology (ASM)

RESOURCES