Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2009 Sep 28.
Published in final edited form as: FEMS Microbiol Lett. 2009 May 1;295(2):204–210. doi: 10.1111/j.1574-6968.2009.01595.x

Genome-scale transcriptional profiling in Staphylococcus aureus: Bringing order out of chaos

Vijayaraj Nagarajan 1, Mark S Smeltzer 2, Mohamed O Elasri 1
PMCID: PMC2753426  NIHMSID: NIHMS142378  PMID: 19459979

Abstract

We used the Staphylococcus aureus microarray meta-database (SAMMD) to compare the transcriptional profiles defined by different experiments targeting the same phenomenon in S. aureus. We specifically examined differences associated with the accessory gene regulator (agr), the staphylococcal accessory regulator (sarA), and growth within a biofilm. We found that in all three cases, there was a striking lack of overlap between the transcriptional profiles. For instance, while all experiments focusing on biofilm formation identified hundreds of differentially-expressed genes, only one of these was common to all transcriptomes. Several factors could potentially contribute to this variability including the use of different biofilm models, different growth media, different microarray platforms and, perhaps most importantly, different strains of S. aureus. The last appeared to be particularly important in the case of the agr and sarA transcriptomes. While these results emphasize the need to introduce some degree of standardization into genome-scale, microarray-based transcriptional profiling experiments, they also demonstrate the need to consider multiple strains of S. aureus in order to avoid any strain-specific bias in the interpretation of results. Our comparisons also illustrate how identification of strain-dependent differences using SAMMD can lead to the development of specific hypotheses that can then be experimentally addressed. Based on this, we have added new features to SAMMD that allow for direct comparisons between transcriptional profiling experiments.

Keywords: Staphylococcal accessory regulator, accessory gene regulator, Staphylococcus aureus Microarray Meta-Database

Introduction

Staphylococcus aureus is an important human pathogen capable of causing a diverse array of infections. The pathogenic diversity of S. aureus is due to its capacity to produce a wide range of virulence factors, all of which are under the control of complex regulatory circuits. Modulating the production of these virulence factors allows S. aureus to adapt to diverse and changing conditions within the host, and this no doubt plays a major role in its ability to cause such diverse forms of infection. Genome-scale, microarray-based transcriptional profiling is an important tool that allows the comprehensive analysis of changes that occur in response to diverse stimuli [1], and a growing number of investigators have used this tool to help dissect global regulatory pathways and the nature of the adaptive response of S. aureus.

Studies done to date have examined the differential pattern of gene expression in response to various stress conditions including acid tolerance, the stringent response, exposure to various antibiotics, and growth within a biofilm. Studies have also been done examining the impact of mutations in specific regulatory genes including agrA, sarA, saeRS, vraSR and mgrA. All together, genome-scale transcriptional profiling has been used in at least 92 independent experiments. These studies have generated a vast amount of data that is difficult to analyze in detail, particularly with respect to making comparisons between different stimuli and/or strains of S. aureus. To help address this issue, we developed the S. aureus microarray meta-database (SAMMD; http://www.bioinformatics.org/sammd/) [2]. Specifically, the SAMMD database contains the complete lists of all differentially expressed genes from all published genome-scale transcriptional profiling experiments. Where possible, all ORF designations in SAMMD are mapped to the N315 genome irrespective of the original strain used in each study. To illustrate the utility of SAMMD, and to emphasize the importance of strain-dependent differences, we used SAMMD to compare gene expression profiles from different experiments with a specific emphasis on strain-dependent differences with respect to the accessory gene regulator (agr), the staphylococcal accessory regulator (sarA), and growth in biofilm.

Materials and methods

We selected agr, sarA, and biofilm for inclusion in this analysis because several microarray experiments from different strains were done on them. We used the SAMMD database “compare” function to compare and contrast transcriptomes from different strains. We tested the contribution of the strains towards gene regulation by statistical analysis using the software JMP. In all three cases, the null hypothesis that the strains did not have any association with the gene regulation was rejected based on significant p-values (see below).

The regulatory impact of agr

Although a growing number of regulatory loci have been described, it is clear that the accessory gene regulator (agr) plays a central role in the pathogenesis of S. aureus infection. In fact, recent data suggests that a high level of agr expression accounts in large part for the hypervirulence of community-acquired, methicillin-resistant S. aureus (CA-MRSA) [3]. The agr operon itself contains 4 genes that collectively make up a classic quorum-sensing signal transduction system [45]. These 4 genes are transcribed as part of a polycistronic message designated RNAII. Activation of this system results in increased production of the divergently-transcribed product RNAIII. Although it includes the gene for delta toxin (hld), it is clear that RNAIII functions as a regulatory RNA [6]. Until recently, the response regulator of the agr system (AgrA) had no known function other than activation of transcription from the RNAII (P2) and RNAIII (P3) promoters [47], but a recent report demonstrated that AgrA also regulates production of the cytolytic phenol-soluble modulins (PSMs) in an RNAIII-independent manner [8]. Production of RNAIII modulates the production of a wide variety of genes at both the transcriptional and post-transcriptional levels [45]. Although there are exceptions to the rule, increased production of RNAIII generally results in increased production of extracellular virulence factors and decreased production of surface-associated virulence factors [49].

The first study that examined the global impact of mutating agr on gene expression focused on a strain designated RN27, which is derived from the commonly-studied NCTC8325 lineage [10]. This study found that agr modulates the expression of over 100 genes and, in general, confirmed the regulatory paradigm of agr-mediated activation of extracellular virulence factors and repression of surface-associated virulence factors. However, when the same experiment was repeated with the clinical isolate UAMS-1, only 16 genes were found to overlap in the two agr regulons [11]. Genes that were down-regulated by agr in both strains included an amino acid permease and a putative drug transporter (Table 1). Among the genes that were up-regulated by agr in both strains were the agr operon itself, including RNAIII, the gene encoding alpha-toxin (hla), and the genes involved in arginine metabolism. Importantly, several genes that are classically associated with the agr regulon were not found in common between the two strains. This includes spa, hlgB, hlgC, geh and the genes encoding a variety of other toxins.

Table 1.

Comparison of agr transcriptomes from strains RN27 and UAMS-1

Genes down-regulated by agr in both strains
SA1269 - putative drug transporter,
SA1270 - amino acid permease
Genes up-regulated by agr in both strains
SA0184 - conserved hypothetical protein
SA0185 murQ SIS domain protein
SA0187 - transcriptional regulator, rpiR family domain protein
SA1007 - alpha-hemolysin precursor
SA1842 agrB accessory gene regulator B
SA1843 agrC accessory gene regulator C
SA1844 agrA accessory gene regulator protein A
SA2006 - putative map protein
SA2424 - Crp/Fnr family protein
SA2425 arcC carbamate kinase
SA2426 arcD arginine/ornithine antiporter
SA2427 arcB ornithine carbamoyltransferase
SA2428 arcA arginine deiminase
SA2463 lip triacylglycerol lipase precursor

The contribution of the strains towards gene regulation by agr was statistically tested using LR Chi-Squared tests and the null hypothesis that the strains did not have any association with the gene regulation was rejected based on a significant p-value of less than 10−34.

Although the regulatory events controlling expression of many of these genes has not been examined in detail, spa is an exception in that it is often taken as a “classic” surface protein, production of which is dramatically increased in an agr mutant [4]. We cite this specific example to emphasize how SAMMD can be used to identify such discrepancies and thereby point investigators toward relevant experiments that may be useful in resolving them in a coherent and meaningful way. For example, in the currently-accepted S. aureus regulatory paradigm, the impact of agr on expression of spa is at least partially an indirect effect in that agr represses expression of sarT, which induces expression of sarS. Increased production of SarS then induces increased expression of spa [12]. In this context, mutation of agr would result in increased expression of sarT, which would in turn result in increased expression of sarS and consequently spa. However, sarT is absent in many clinical isolates including UAMS-1 [13]. Whether this accounts for or even contributes to the observed discrepancy with respect to agr and spa expression remains unknown. Indeed, by comparison to the 8325-4 strain RN6390, sarS is expressed at greatly increased levels in UAMS-1 despite the presence of sarT in the former and its absence in the latter [1113]. This clearly implies that an additional regulatory factor is involved, one possibility being tcaR, which is known to modulate the expression of sarS and, while conserved in both strains, is nonfunctional in all 8325-derived strains including RN27 [14]. In fact, recent evidence suggests that sarS may play an important role with respect to other strain-dependent regulatory differences associated with the staphylococcal accessory regulator (sarA) [1516].

The regulatory impact of sarA

The staphylococcal accessory regulator (sarA) is a second global regulator that modulates the production of multiple virulence factors. The effector molecule of the sarA regulatory system is a 15 kDa DNA-binding protein (SarA). Because mutation of sarA results in increased production of some virulence factors and decreased production of others, it is widely presumed that SarA can act as both a repressor and an activator [1718]. However, other reports have suggested that SarA functions primarily as a repressor, with the increased production of certain virulence factors observed in sarA mutants being an indirect effect [1920]. Similarly, it is widely assumed that SarA functions primarily at the transcriptional level [21], but recent data suggests that sarA may also have post-transcriptional regulatory effects associated with stabilization of mRNA transcripts [22].

To date, transcriptional profiling experiments have been done with sarA mutants generated in the S. aureus strains RN27, COL, and UAMS-1. Those done with RN27 and UAMS-1 have been published [1011], while the COL experiments are currently limited to inclusion in SAMMD. The growth medium used in these experiments was similar but not identical, with brain heart infusion (BHI) broth being employed with RN27 and tryptic soy broth (TSB) being employed with COL and UAMS-1. Results from all three experiments were validated by real-time quantitative PCR (RTqPCR). Based on analysis of RNA samples prepared from the post-exponential growth phase alone, more than 100 genes were identified in each experiment. However, only four of these were common to all three sarA regulons. Expression of all four of these genes was increased in a sarA mutant. Three of these genes encode extracellular proteases (SspC, SspB, and Aur), while the 4th encodes a threonine dehydratase (Table 2). No common genes were expressed at decreased levels in a sarA mutant. Overall, this is consistent with the hypothesis that SarA functions primarily as a repressor. Pairwise comparison of the sarA transcriptomes shows that the RN27 and COL share the highest overlap (18 genes), followed by COL and UAMS-1 (13 genes) and finally UAMS-1 and RN27 (7 genes). The pairwise comparison reflects the close relatedness of strains RN27 and COL by comparison to UAMS-1 [13]. More importantly, several genes that are classically associated with the sarA regulon are conspicuously absent. For instance, AgrA is found in the RN27 and COL transcriptomes but not in UAMS-1. This is consistent with targeted mutagenesis experiments demonstrating that sarA activates agr transcription in the 8325-4 strain RN6390 [23] but plays a limited role in that regard in UAMS-1 [11]. Other genes that did not appear in the transcriptomes of all three strains were hla, spa, fnbA, sspA, sarT, sarS, sarU, sarV, and the rest of the genes in the agr operon (agrB, agrC and agrD).

Table 2.

Comparison of sarA transcriptomes in strains UAMS-1, RN27, and COL

Genes down-regulated by sarA in all three strains
SA0899 sspC cysteine protease
SA0900 sspB cysteine protease precursor
SA2430 aur zinc metalloproteinase aureolysin
SACOL1477 ilvA1 threonine dehydratase, catabolic
Genes down-regulated by sarA only in strains UAMS-1 and RN27
SA0746 staphylococcal nuclease
Genes down-regulated by sarA only in strains RN27 and COL
SA0107 spa Immunoglobulin G binding protein A precursor
SA0271 - Unknown
SA0923 purM phosphoribosylformylglycinamidine cyclo-ligase
SA2424 - Crp/Fnr family protein
Genes down-regualted by sarA only in strains UAMS and COL
SA0091 plc 1-phosphatidylinositol phosphodiesterase precursor
SA0136 - phosphonates ABC transporter, permease protein CC0363
SA0901 sspA glutamyl endopeptidase precursor
SA0904 - transcriptional regulator, MarR family family
SA1269 - drug transporter, putative
SA1270 - amino acid permease
SA1272 - alanine dehydrogenase
SA1725 - Staphopain, Cysteine Proteinase
SAR1022 sspA V8 Protease
Gene upregulated only in RN27 and COL
SA0174 - 4-phosphopantetheinyl transferase superfamily family
SA1844 agrA accessory gene regulator protein A
SA2319 - L-serine dehydratase, iron-sulfur-dependent, beta subunit
SA2447 - LPXTG-motif cell wall anchor domain protein
SACOL1187 - antibacterial protein (phenol soluble modulin)
SAS016 - conserved hypothetical protein

The contribution of the strains towards gene regulation by sarA was statistically tested using LR Chi-Squared tests (Whole Model Tests) and the null hypothesis that the strains did not have any association with the gene regulation was rejected based on the significant p-value 0.0004.

The preceding discussion with respect to the agr regulon in different strains makes it obvious why certain of these genes were not identified in all three sarA regulons (e.g. the absence of sarT and sarU in the UAMS-1 genome). In other cases, the reasons are harder to define. For instance, it is clear that mutation of sarA results in a reduced capacity to bind fibronectin in all strains studied to date [20], and because in vitro assays indicate that SarA can bind cis elements associated with fnbA, it has been proposed that this is a direct regulatory effect associated with SarA acting as an activator of fnbA transcription [24]. However, other studies have concluded that the impact of SarA on the ability to bind fibronectin is an indirect effect associated with the increased production of proteases in sarA mutants [2025]. In the case of hla, mutation of sarA results in decreased transcription in RN6390 but increased transcription in UAMS-1 [1011-2026], and it was recently proposed that this disparate effect may also revolve around the tcaR mutation in 8325-4 strains like RN6390 and its impact on expression of sarS [27].

Gene expression patterns in S. aureus biofilms

When assessing the impact of specific regulatory genes on transcriptional profile, the primary variable is strain choice. However, it is certainly not the only consideration in gene profiling experiments. For example, three independent experiments have examined the differential pattern of gene expression in S. aureus biofilms [2830], but a comparison of these experiments using SAMMD reveals an almost complete lack of overlap between the regulons defined by these studies. In addition to employing different strains, all three experiments also employed different growth media and different growth stages at which RNA was harvested for analysis. Additional differences existed between the microarray platforms themselves (Agilent vs. Affymetrix vs. TIGR-PFGRC spotted arrays) and, to a limited extent, the degree to which expression of a gene was altered before being identified as “differentially expressed” (1.5 to 2.5-fold). None of the various growth conditions or experimental approaches employed in these experiments can be taken as more definitive than another, and in fact the discrepancies may only reflect differences between the gene expression pattern in newly formed vs. mature biofilms, but such comparisons nevertheless make it obvious that determining whether or not that is the case would be facilitated by some degree of standardization.

At the same time, even absent that standardization, useful information can be obtained using comparisons made with SAMMD. For instance, a comparison of the single stage from each experiment that would arguably represent the most mature form of a biofilm analyzed in that experiment ([28], 7 day biofilm vs. stationary phase planktonic culture; [30], 48 hr biofilms vs. 48 hr planktonic growth; [29], “mature” biofilms vs. planktonic growth) reveals that only one gene is present in all three regulons (Table 3). This gene is arcB (SA2427), which is part of the arginine deiminase operon. However, eight genes were present in two of three regulons, and included among these eight genes were SA2424, SA2425 (arcC), SA2426 (arcD) and SA2428 (arcA). Taken together, this provides strong support for the hypothesis that the arginine deiminase operon warrants further investigation in the specific context of an S. aureus biofilm. While some studies have subsequently concluded that the arc operon plays a limited role in biofilm formation [31], others have confirmed that one of the characteristics that define CA-MRSA is the presence of a second copy of the arc operon within the arginine catabolite mobile element (ACME). In fact, it has been proposed that the presence of ACME enhances the growth and overall “fitness” of CA-MRSA isolates including clones of the USA300 lineage [32]. How this will ultimately play out remains to be determined, but the more important point in the context of this report is to illustrate how comparisons made using SAMMD can be used to formulate questions that need to be asked and define the best context in which to pursue the answer.

Table 3.

Comparison of biofilm transcriptomes in strains MRSA-M2, SA113, and UAMS-1

Genes up-regulated in biofilm formed by strains MRSA-M2 and SA113 while down-regulated in UAMS-1
SA2204 - phosphoglycerate mutase
Genes up-regulated in biofilm formed by all three strains
SA2427 arcB ornithine carbamoyltransferase
Genes up-regulated in biofilm formed by strains SA113 and UAMS-1
SA2007 - alpha-acetolactate decarboxylase
SA2008 alsS acetolactate synthase
SA2086 ureF urease accessory protein UreF
SA2088 ureD urease accessory protein UreD
SA2424 - Crp/Fnr family protein
SA2425 arcC carbamate kinase
SA2426 arcD arginine/ornithine antiporter
SA2428 arcA arginine deiminase
Gene down-regulated in biofilm formed by strains MRSA-M2 and UAMS-1
SA1382 sodA superoxide dismutase, Mn
Gene down-regulated in biofilm formed by strains SA113 and UAMS-1
SA0916 - phosphoribosylaminoimidazole carboxylase, catalytic subunit
Gene up-regulated in MRSA-M2 biofilm while down-regulated in UAMS-1
SA1150 glnA glutamine synthetase, type I
Genes up-regulated in SA113 biofilm while down-regulated in UAMS-1
SA0572 - hydrolase, alpha/beta fold family
SA1531 ald alanine dehydrogenase
SA2203 - drug resistance transporter, EmrB/QacA subfamily

The contribution of the strains towards gene regulation in biofilm was statistically tested using LR Chi-Squared tests (Whole Model Tests) and the null hypothesis that the strains did not have any association with the gene regulation was rejected based on a significant p-value of less than 10−41.

Conclusion

This report highlights the fact that S. aureus strains are profoundly different from each other and serves as a cautionary tale for drawing sweeping conclusions about regulatory pathways in S. aureus. This is a particularly important consideration in light of the fact that current gene regulation paradigms in S. aureus are based almost exclusively on studies done with derivatives of NCTC8325, all of which are known to carry mutations in at least two important regulatory loci [14-33-37]. However, we also illustrate how SAMMD can be used to sort through the growing body of transcriptome data to identify key questions that warrant further investigation and perhaps even point the way toward specific approaches that can be used to address those questions. To facilitate this approach, we have added a “comparison” feature that allows SAMMD users to make direct comparisons between up to three different transcriptional profiling experiments. We have also added a feature that would allow researchers to examine transcriptional data collectively or by specific strains. It is hoped that this will not only allow investigators to identify relevant discrepancies in a manner that will facilitate their resolution but also to identify those features that are consistent across all strains and perhaps even all growth conditions. The latter is important in that the utility of any therapeutic agent would be greatest if it were effective against as many strains and in as many clinical situations as possible.

Acknowledgments

We thank Jeff Skinner, Biostatistics Specialist from Bioinformatics and Computational Biosciences Branch, Office of Cyber Infrastructure and Computational Biology, National Institute of Allergy and Infectious Diseases, National Institutes of Health, Bethesda, MD 20892, USA, for help with testing the significance of association between strains and regulation of genes.

This work was supported by grant no. 1R15AI062727-01A1 from the National Institute of Allergy and Infectious Diseases (NIAID) to MOE and by The Mississippi Functional Genomics Network (NIH/NCRR P20 RR016476). Further support was provided by grant no. R01AI43356 to MSS.

References

  • 1.Schena M, Shalon D, Davis RW, Brown PO. Quantitative monitoring of gene expression patterns with a complementary DNA microarray. Science. 1995;270:467–470. doi: 10.1126/science.270.5235.467. [DOI] [PubMed] [Google Scholar]
  • 2.Nagarajan V, Elasri MO. SAMMD: Staphylococcus aureus microarray meta-database. BMC Genomics. 2007;8:351. doi: 10.1186/1471-2164-8-351. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 3.Wang R, Braughton KR, Kretschmer D, Bach TH, Queck SY, Li M, Kennedy AD, Dorward DW, Klebanoff SJ, Peschel A, DeLeo FR, Otto M. Identification of novel cytolytic peptides as key virulence determinants for community-associated MRSA. Nat Med. 2007;13:1510–1514. doi: 10.1038/nm1656. [DOI] [PubMed] [Google Scholar]
  • 4.Novick RP. Autoinduction and signal transduction in the regulation of staphylococcal virulence. Mol Microbiol. 2003;48:1429–1449. doi: 10.1046/j.1365-2958.2003.03526.x. [DOI] [PubMed] [Google Scholar]
  • 5.George EA, Muir TW. Molecular mechanisms of agr quorum sensing in virulent staphylococci. Chembiochem. 2007;8:847–855. doi: 10.1002/cbic.200700023. [DOI] [PubMed] [Google Scholar]
  • 6.Janzon L, Arvidson S. The role of the delta-lysin gene (hld) in the regulation of virulence genes by the accessory gene regulator (agr) in Staphylococcus aureus. Embo J. 1990;9:1391–1399. doi: 10.1002/j.1460-2075.1990.tb08254.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 7.Koenig RL, Ray JL, Maleki SJ, Smeltzer MS, Hurlburt BK. Staphylococcus aureus AgrA binding to the RNAIII-agr regulatory region. J Bacteriol. 2004;186:7549–7555. doi: 10.1128/JB.186.22.7549-7555.2004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 8.Queck SY, Jameson-Lee M, Villaruz AE, Bach TH, Khan BA, Sturdevant DE, Ricklefs SM, Li M, Otto M. RNAIII-independent target gene control by the agr quorum-sensing system: insight into the evolution of virulence regulation in Staphylococcus aureus. Mol Cell. 2008;32:150–158. doi: 10.1016/j.molcel.2008.08.005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 9.Projan SJ, Novick RP. The Molecular basis of pathogenicity. In: Crossley GAaK., editor. The Staphylococci in human disease. Churchill Livingstone; New York, N.Y: 1997. pp. 55–81. [Google Scholar]
  • 10.Dunman PM, Murphy E, Haney S, Palacios D, Tucker-Kellogg G, Wu S, Brown EL, Zagursky RJ, Shlaes D, Projan SJ. Transcription profiling-based identification of Staphylococcus aureus genes regulated by the agr and/or sarA loci. J Bacteriol. 2001;183:7341–7353. doi: 10.1128/JB.183.24.7341-7353.2001. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 11.Cassat J, Dunman PM, Murphy E, Projan SJ, Beenken KE, Palm KJ, Yang SJ, Rice KC, Bayles KW, Smeltzer MS. Transcriptional profiling of a Staphylococcus aureus clinical isolate and its isogenic agr and sarA mutants reveals global differences in comparison to the laboratory strain RN6390. Microbiology. 2006;152:3075–3090. doi: 10.1099/mic.0.29033-0. [DOI] [PubMed] [Google Scholar]
  • 12.Schmidt KA, Manna AC, Cheung AL. SarT influences sarS expression in Staphylococcus aureus. Infect Immun. 2003;71:5139–5148. doi: 10.1128/IAI.71.9.5139-5148.2003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 13.Cassat JE, Dunman PM, McAleese F, Murphy E, Projan SJ, Smeltzer MS. Comparative genomics of Staphylococcus aureus musculoskeletal isolates. J Bacteriol. 2005;187:576–592. doi: 10.1128/JB.187.2.576-592.2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 14.McCallum N, Bischoff M, Maki H, Wada A, Berger-Bachi B. TcaR, a putative MarR-like regulator of sarS expression. J Bacteriol. 2004;186:2966–2972. doi: 10.1128/JB.186.10.2966-2972.2004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 15.Gustafsson E, Oscarsson J. Maximal transcription of aur (aureolysin) and sspA (serine protease) in Staphylococcus aureus requires staphylococcal accessory regulator R (sarR) activity. FEMS Microbiol Lett. 2008;284:158–164. doi: 10.1111/j.1574-6968.2008.01198.x. [DOI] [PubMed] [Google Scholar]
  • 16.Oscarsson J, Harlos C, Arvidson S. Regulatory role of proteins binding to the spa (protein A) and sarS (staphylococcal accessory regulator) promoter regions in Staphylococcus aureus NTCC 8325–4. Int J Med Microbiol. 2005;295:253–266. doi: 10.1016/j.ijmm.2005.05.003. [DOI] [PubMed] [Google Scholar]
  • 17.Chien Y, Manna AC, Projan SJ, Cheung AL. SarA, a global regulator of virulence determinants in Staphylococcus aureus, binds to a conserved motif essential for sar-dependent gene regulation. J Biol Chem. 1999;274:37169–37176. doi: 10.1074/jbc.274.52.37169. [DOI] [PubMed] [Google Scholar]
  • 18.Cheung AL, Zhang G. Global regulation of virulence determinants in Staphylococcus aureus by the SarA protein family. Front Biosci. 2002;7:d1825–1842. doi: 10.2741/A882. [DOI] [PubMed] [Google Scholar]
  • 19.Arvidson S, Tegmark K. Regulation of virulence determinants in Staphylococcus aureus. Int J Med Microbiol. 2001;291:159–170. doi: 10.1078/1438-4221-00112. [DOI] [PubMed] [Google Scholar]
  • 20.Blevins JS, Beenken KE, Elasri MO, Hurlburt BK, Smeltzer MS. Strain-dependent differences in the regulatory roles of sarA and agr in Staphylococcus aureus. Infect Immun. 2002;70:470–480. doi: 10.1128/IAI.70.2.470-480.2002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 21.Cheung AL, Nishina K, Manna AC. SarA of Staphylococcus aureus binds to the sarA promoter to regulate gene expression. J Bacteriol. 2008;190:2239–2243. doi: 10.1128/JB.01826-07. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 22.Roberts C, Anderson KL, Murphy E, Projan SJ, Mounts W, Hurlburt B, Smeltzer M, Overbeek R, Disz T, Dunman PM. Characterizing the effect of the Staphylococcus aureus virulence factor regulator, SarA, on log-phase mRNA half-lives. J Bacteriol. 2006;188:2593–2603. doi: 10.1128/JB.188.7.2593-2603.2006. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 23.Chien Y, Manna AC, Cheung AL. SarA level is a determinant of agr activation in Staphylococcus aureus. Mol Microbiol. 1998;30:991–1001. doi: 10.1046/j.1365-2958.1998.01126.x. [DOI] [PubMed] [Google Scholar]
  • 24.Wolz C, Pohlmann-Dietze P, Steinhuber A, Chien YT, Manna A, van Wamel W, Cheung A. Agr-independent regulation of fibronectin-binding protein(s) by the regulatory locus sar in Staphylococcus aureus. Mol Microbiol. 2000;36:230–243. doi: 10.1046/j.1365-2958.2000.01853.x. [DOI] [PubMed] [Google Scholar]
  • 25.Karlsson A, Saravia-Otten P, Tegmark K, Morfeldt E, Arvidson S. Decreased amounts of cell wall-associated protein A and fibronectin-binding proteins in Staphylococcus aureus sarA mutants due to up-regulation of extracellular proteases. Infect Immun. 2001;69:4742–4748. doi: 10.1128/IAI.69.8.4742-4748.2001. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 26.Chan PF, Foster SJ. Role of SarA in virulence determinant production and environmental signal transduction in Staphylococcus aureus. J Bacteriol. 1998;180:6232–6241. doi: 10.1128/jb.180.23.6232-6241.1998. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 27.Oscarsson J, Kanth A, Tegmark-Wisell K, Arvidson S. SarA is a repressor of hla (alpha-hemolysin) transcription in Staphylococcus aureus: its apparent role as an activator of hla in the prototype strain NCTC 8325 depends on reduced expression of sarS. J Bacteriol. 2006;188:8526–8533. doi: 10.1128/JB.00866-06. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 28.Beenken KE, Dunman PM, McAleese F, Macapagal D, Murphy E, Projan SJ, Blevins JS, Smeltzer MS. Global gene expression in Staphylococcus aureus biofilms. J Bacteriol. 2004;186:4665–4684. doi: 10.1128/JB.186.14.4665-4684.2004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 29.Brady RA, Leid JG, Camper AK, Costerton JW, Shirtliff ME. Identification of Staphylococcus aureus proteins recognized by the antibody-mediated immune response to a biofilm infection. Infect Immun. 2006;74:3415–3426. doi: 10.1128/IAI.00392-06. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 30.Resch A, Rosenstein R, Nerz C, Gotz F. Differential gene expression profiling of Staphylococcus aureus cultivated under biofilm and planktonic conditions. Appl Environ Microbiol. 2005;71:2663–2676. doi: 10.1128/AEM.71.5.2663-2676.2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 31.Zhu Y, Weiss EC, Otto M, Fey PD, Smeltzer MS, Somerville GA. Staphylococcus aureus biofilm metabolism and the influence of arginine on polysaccharide intercellular adhesin synthesis, biofilm formation, and pathogenesis. Infect Immun. 2007;75:4219–4226. doi: 10.1128/IAI.00509-07. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 32.Diep BA, Stone GG, Basuino L, Graber CJ, Miller A, des Etages SA, Jones A, Palazzolo-Ballance AM, Perdreau-Remington F, Sensabaugh GF, DeLeo FR, Chambers HF. The arginine catabolic mobile element and staphylococcal chromosomal cassette mec linkage: convergence of virulence and resistance in the USA300 clone of methicillin-resistant Staphylococcus aureus. J Infect Dis. 2008;197:1523–1530. doi: 10.1086/587907. [DOI] [PubMed] [Google Scholar]
  • 33.Kullik I, Giachino P, Fuchs T. Deletion of the alternative sigma factor sigmaB in Staphylococcus aureus reveals its function as a global regulator of virulence genes. J Bacteriol. 1998;180:4814–4820. doi: 10.1128/jb.180.18.4814-4820.1998. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 34.Gertz S, Engelmann S, Schmid R, Ohlsen K, Hacker J, Hecker M. Regulation of sigmaB-dependent transcription of sigB and asp23 in two different Staphylococcus aureus strains. Mol Gen Genet. 1999;261:558–566. doi: 10.1007/s004380051001. [DOI] [PubMed] [Google Scholar]
  • 35.Giachino P, Engelmann S, Bischoff M. Sigma(B) activity depends on RsbU in Staphylococcus aureus. J Bacteriol. 2001;183:1843–1852. doi: 10.1128/JB.183.6.1843-1852.2001. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 36.Jefferson KK, Pier DB, Goldmann DA, Pier GB. The teicoplanin-associated locus regulator (TcaR) and the intercellular adhesin locus regulator (IcaR) are transcriptional inhibitors of the ica locus in Staphylococcus aureus. J Bacteriol. 2004;186:2449–2456. doi: 10.1128/JB.186.8.2449-2456.2004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 37.Shaw LN, Aish J, Davenport JE, Brown MC, Lithgow JK, Simmonite K, Crossley H, Travis J, Potempa J, Foster SJ. Investigations into sigmaB-modulated regulatory pathways governing extracellular virulence determinant production in Staphylococcus aureus. J Bacteriol. 2006;188:6070–6080. doi: 10.1128/JB.00551-06. [DOI] [PMC free article] [PubMed] [Google Scholar]

RESOURCES