Skip to main content
Journal of Bacteriology logoLink to Journal of Bacteriology
. 2004 Jul;186(14):4665–4684. doi: 10.1128/JB.186.14.4665-4684.2004

Global Gene Expression in Staphylococcus aureus Biofilms

Karen E Beenken 1, Paul M Dunman 2, Fionnuala McAleese 2, Daphne Macapagal 2, Ellen Murphy 2, Steven J Projan 3, Jon S Blevins 4, Mark S Smeltzer 1,*
PMCID: PMC438561  PMID: 15231800

Abstract

We previously demonstrated that mutation of the staphylococcal accessory regulator (sarA) in a clinical isolate of Staphylococcus aureus (UAMS-1) results in an impaired capacity to form a biofilm in vitro (K. E. Beenken, J. S. Blevins, and M. S. Smeltzer, Infect. Immun. 71:4206-4211, 2003). In this report, we used a murine model of catheter-based biofilm formation to demonstrate that a UAMS-1 sarA mutant also has a reduced capacity to form a biofilm in vivo. Surprisingly, mutation of the UAMS-1 ica locus had little impact on biofilm formation in vitro or in vivo. In an effort to identify additional loci that might be relevant to biofilm formation and/or the adaptive response required for persistence of S. aureus within a biofilm, we isolated total cellular RNA from UAMS-1 harvested from a biofilm grown in a flow cell and compared the transcriptional profile of this RNA to RNA isolated from both exponential- and stationary-phase planktonic cultures. Comparisons were done using a custom-made Affymetrix GeneChip representing the genomic complement of six strains of S. aureus (COL, N315, Mu50, NCTC 8325, EMRSA-16 [strain 252], and MSSA-476). The results confirm that the sessile lifestyle associated with persistence within a biofilm is distinct by comparison to the lifestyles of both the exponential and postexponential phases of planktonic culture. Indeed, we identified 48 genes in which expression was induced at least twofold in biofilms over expression under both planktonic conditions. Similarly, we identified 84 genes in which expression was repressed by a factor of at least 2 compared to expression under both planktonic conditions. A primary theme that emerged from the analysis of these genes is that persistence within a biofilm requires an adaptive response that limits the deleterious effects of the reduced pH associated with anaerobic growth conditions.


Staphylococcus aureus is a prominent human pathogen that causes a wide variety of infections. Of particular interest in our laboratory are musculoskeletal infections including those associated with orthopedic implants. The hallmark characteristic of these infections is formation of a biofilm, which consists of multiple layers of bacteria encased within an exopolysaccharide glycocalyx. The presence of this glycocalyx protects the enclosed bacteria from host defenses and impedes delivery of at least some antibiotics (64). Moreover, bacteria within biofilms adopt a phenotype that confers intrinsic resistance to many antibiotics. For example, the reduced growth rate of biofilm-associated bacteria limits the efficacy of antibiotics that target cell wall biosynthesis, while the reduced oxidative metabolism limits the uptake of aminoglycosides (33, 64, 65). Consequently, biofilm-associated infections are recalcitrant to antimicrobial therapy and often require surgical intervention to debride infected tissues and/or remove colonized implants.

The formation of three-dimensional biofilms is a complex process that can be subdivided into the relatively distinct phases of attachment, accumulation, maturation, and dispersal (10). With respect to staphylococcal biofilms, the primary emphasis so far has been placed on the attachment and accumulation phases, which appear to be mediated by different types of adhesins. More specifically, a group of surface-exposed proteins collectively referred to as MSCRAMMs (microbial surface components recognizing adhesive matrix molecules) (48) appear to be the primary determinants responsible for the initial attachment to both native tissues and biomaterials, while the accumulation phase appears to be dependent on polysaccharide adhesins that promote adhesive interactions between bacterial cells (26). Although a number of candidate polysaccharides have been described, there is an emerging consensus that the primary determinant of the accumulation phase of staphylococcal biofilm formation is the polysaccharide intercellular adhesin (PIA), production of which is dependent upon the genes within the icaADBC operon (28). Composition studies have demonstrated that PIA consists of polymeric N-acetylglucosamine, and for this reason it has also been referred to as PNAG (40).

The ica operon was first identified in Staphylococcus epidermidis (28) and has been studied most extensively in that species. However, it is also present and appears to serve the same function in S. aureus (14). Most S. aureus strains appear to contain the entire ica operon (14, 22, 53), although there are reports to the contrary (3), and it is clear that there are strain-dependent differences with respect to the overall capacity to form a biofilm in vitro (5, 14, 53). The ica operon is subject to phase variation in S. epidermidis (75), and a number of studies have indicated that expression of ica in both S. epidermidis and S. aureus is also subject to environmental regulation. Perhaps most importantly, McKenney et al. (42) demonstrated that PNAG production in S. aureus is enhanced during in vivo growth. Rachid et al. (52) subsequently demonstrated that expression of ica is at least partially controlled by the stress response transcription factor σB. In addition, anaerobic growth was found to induce expression of the ica operon and PIA production in both S. epidermidis and S. aureus (15).

Recently, Conlon et al. (12) reported that icaR, which is located immediately upstream of the ica operon, encodes a repressor that is important for the environmental regulation of ica expression in S. epidermidis. However, studies done with S. aureus have demonstrated that regulation of ica expression and the ability to form a biofilm also involve regulatory elements other than σB and IcaR (66). Included among these additional regulatory loci are the accessory gene regulator (agr) and the staphylococcal accessory regulator (sarA). The agr locus encodes a two-component quorum-sensing system that modulates production of a regulatory RNA molecule (RNAIII) in a density-dependent manner. Induction of RNAIII synthesis results in reduced production of surface proteins (e.g., MSCRAMMs) and a concomitant increase in production of exotoxins (4, 45). Production of δ-toxin, which is encoded within the RNAIII locus, has been negatively correlated with biofilm formation (69, 70). This suggests that strains expressing agr at high levels would have a reduced capacity to form a biofilm, which is consistent both with our results (5) and results from other laboratories (70).

The sarA locus encodes a 14.5-kDa DNA-binding protein (SarA) that is required at least under some growth conditions for maximum expression from the agr and RNAIII promoters (29). This would imply that mutation of sarA would limit production of RNAIII and thereby enhance the ability to form a biofilm. However, recent reports have confirmed that mutation of sarA results in a reduced capacity to form a biofilm (5, 66). SarA also regulates expression of other genes in an agr-independent manner (6, 19, 72, 74), and Valle et al. (66) recently demonstrated that mutation of sarA results in reduced transcription of the ica operon and a reduced capacity to produce PNAG. They also suggested that SarA may promote biofilm formation in an indirect manner by suppressing transcription of a repressor of PNAG synthesis or a protein involved in the turnover of PNAG.

The persistence of bacteria within a biofilm also requires an adaptive response appropriate for the sessile lifestyle. The availability of complete bacterial genome sequences has facilitated the use of microarray technologies to identify genes that are differentially expressed by biofilm-encased bacteria. Using an array representing 99% of the Bacillus subtilis genome, Stanley et al. (63) identified 519 genes that were differentially expressed in biofilms as opposed to planktonic cultures. Similarly, Schembri et al. (59) found that 5 to 10% of the genes in the Escherichia coli genome were differentially expressed in biofilms, depending on which planktonic growth condition was used as a reference. Included among these genes were 30 of the 65 genes previously reported to be under the regulatory control of the general stress response regulator rpoS (36). Schembri et al. (59) subsequently demonstrated that an E. coli rpoS mutant was incapable of forming a biofilm. However, Whiteley et al. (71) found that expression of rpoS was repressed in Pseudomonas aeruginosa biofilms and that a P. aeruginosa rpoS mutant formed a more extensive biofilm than the corresponding wild-type strain. While these results confirm that biofilms represent a unique growth state by comparison to planktonic cultures, they also suggest the existence of species-specific pathways that contribute to biofilm formation and maintenance of the sessile lifestyle. To date, no comprehensive transcriptional analysis of S. aureus biofilms has been reported. However, Prigent-Combaret et al. (51) demonstrated that biofilm-encased E. coli encounter high osmolarity, oxygen limitation, and higher cell density than cells grown under planktonic conditions, and all of these factors are known to influence gene expression in S. aureus (11, 45).

To further investigate these issues, we generated sarA and ica mutations in a clinical isolate of S. aureus (UAMS-1) and examined their relative capacity to form a biofilm both in vitro and in vivo. We also used a custom-made Affymetrix GeneChip representing the combined genomes of six strains of S. aureus (N315, Mu50, COL, NCTC 8325, EMRSA-16 [strain 252], and MSSA-476) to investigate differential gene expression in a mature S. aureus biofilm.

MATERIALS AND METHODS

Bacterial strains.

The experiments described here focus on the S. aureus clinical isolate UAMS-1. This strain was cultured from the bone of a patient suffering from osteomyelitis and was previously shown to form a biofilm both in vitro (5) and in vivo (21). The UAMS-1 sarA mutant was generated by transduction as previously described (7). Φ11-mediated transduction from a derivative of S. aureus SA113 carrying an ica::tet mutation (14) was used to generate a UAMS-1 ica mutant. Transductants were confirmed by Southern blotting using probes corresponding to the sarA and ica loci (6, 7).

Detection of PNAG production.

To assess the production of PNAG in S. aureus clinical isolate UAMS-1 and its sarA and ica mutants, cultures were grown in tryptic soy broth with the appropriate antibiotic. After overnight incubation, the optical density at 560 nm (OD560) was determined, and an equal number of cells (2 to 5 ml of each culture grown overnight) was harvested by centrifugation. Cells were resuspended in 50 μl of 0.5 M EDTA (pH 8.0) and boiled for 5 min. After cellular debris was removed by centrifugation, a 40-μl aliquot of the supernatant was incubated for 30 min with 10 μl of proteinase K (20 mg/ml) at 37°C to reduce nonspecific background levels. After the addition of 10 μl of Tris-buffered saline (20 mM Tris-HCl, 150 mM NaCl [pH 7.4]), 8 μl of each extract was spotted onto a nitrocellulose membrane using a BIO-dot microfiltration apparatus (Bio-Rad Laboratories, Inc., Hercules, Calif.). After drying, the presence of PNAG in the extract was assessed using the WesternBreeze chemiluminescence immunodetection kit (Invitrogen Corp., Carlsbad, Calif.) and anti-PNAG antiserum (kindly provided by Kimberly Jefferson, Channing Laboratory, Harvard Medical School).

Planktonic culture conditions.

Because biofilm formation by strain UAMS-1 in vitro is dependent on supplementation of the medium with 0.5% glucose and 3.0% sodium chloride (5), these supplements were also added to the medium used for planktonic culture. Specifically, 5-ml samples of cultures grown overnight in tryptic soy broth at 37°C with constant aeration were used to inoculate 250 ml of fresh biofilm medium to an OD560 of 0.05. Cultures were incubated with aeration at 37°C. Aliquots were then removed at the mid-exponential (OD560 = 1.0) and stationary (OD560 = 3.5) growth phases. Aliquots were immediately mixed with an equal volume of ice-cold acetone-ethanol (1:1) and stored at −20°C prior to RNA extraction.

Biofilm cultures.

Biofilms were generated in disposable flow cells (Stovall Life Science, Greensboro, N.C.) as previously described (5). Briefly, flow cells were precoated overnight at 4°C with 20% human plasma diluted in carbonate buffer (pH 9.6). The inlet side of the flow cell was then connected to a sterile reservoir filled with biofilm medium. The outlet side was connected to a waste reservoir. Tubing upstream of each individual cell was injected with 0.5 ml of the appropriate culture grown overnight. After the flow of medium was started and bacteria were allowed to enter the flow cell, the flow was stopped and the chamber was incubated inverted at 37°C for 1 h. After the flow cell was set upright, nonadherent bacteria were flushed by adjusting the flow rate to 0.5 ml/min, which is sufficient to replace the volume of the flow cell once every minute. Cells harvested after 1 week by aspiration from the downstream side of the flow cell were immediately mixed with acetone-ethanol as described above.

Assessment of biofilm formation.

Biofilm formation in vitro was assessed using a microtiter plate assay and flow cells as previously described (5). The murine model of catheter-associated biofilm formation described by Kadurugamuwa et al. (31) was used to assess biofilm formation in vivo. Briefly, 20- to 30-g female BALB/c mice (Charles River, Wilmington, Mass.) were anesthetized with ketamine (100 mg/kg of body weight) and xylazine (5 mg/kg), their flanks were shaved, and the skin was cleansed with Betadyne and alcohol. Under aseptic conditions, a 1-cm segment of 14-gauge Teflon intravenous catheter was implanted subcutaneously. The wound was closed with sutures and then cleansed with a Betadyne rinse. Infection was induced approximately 1 h after the implantation procedure by injecting 105 CFU of the test strain into the lumen of the catheter. In some cases, mice were coinfected by injection of a mixture containing 105 CFU of UAMS-1 and 105 CFU of either the sarA or ica mutant. Mice were euthanized 10 days postinfection. The catheters were removed aseptically and washed with phosphate-buffered saline. Catheters were then placed in 10 ml of sterile phosphate-buffered saline and sonicated for 1 min to remove adherent bacteria. The number of bacteria in the sonicate was then determined by plating on tryptic soy agar. To confirm the identity of recovered bacteria and to determine the proportion of UAMS-1 versus the corresponding sarA and ica mutants in coinfection experiments, colonies obtained on tryptic soy agar were transferred to the appropriate selective medium and scored for growth.

RNA isolation and cDNA labeling.

Aliquots of cells harvested from flow cells and stored as described above were pelleted by centrifugation at 7,500 × g for 10 min at 4°C. Each pellet was washed in an equal volume of TES buffer (150 mM NaCl, 78 mM disodium salt EDTA, 100 mM Tris [pH 7.5]) and resuspended to a concentration of 109 CFU per ml in TES buffer containing 100 μg of lysostaphin (Ambicin L; AMBI, Inc., Lawrence, N.Y.) per ml. Samples were incubated at 37°C for 1 h prior to applying the equivalent of 1010 CFU to a Qiagen RNeasy Maxi column. Total bacterial RNA was isolated according to the manufacturer's directions (Qiagen, Inc., Valencia, Calif.). After purification, contaminating DNA was removed with RNase-free DNase I (10 U/40 μg of total bacterial RNA at 37°C for 20 min). RNA was then repurified using RNeasy Mini columns (Qiagen, Inc.). The amount of recovered RNA was determined spectrophotometrically, and the absence of DNA was verified by PCR using primers (Table 1) corresponding to the collagen adhesin gene (cna). Samples were then stored at −80°C.

TABLE 1.

Sequences of primers and TaqMan probes used in this study

Primer or probea Oligonucleotide sequence (5′→3′)
cna-F CAAGCAGTTATTACACCAGACGG
cna-R CACCTTTTACAGTACCTT
arcA-F GTGGTTGACTCATACATCTAGGGC
arcA-R AGACCAGGCGTTGTAGTGACTTA
arcA-P CCCACGTCCACGTACCAGCTCGCT
pyrR-F TTGATGATGTGCTGTATACTGG
pyrR-R CGAATTGGTAACTCACGATGT
pyrR-P CGGTTCGTGCTTCACTTGATGCT
ureA-F CATTTTACACAACGAGAGCAAG
ureA-R ACGTGCTTTACGACGACG
ureA-P CAACTTCCGCCGCCACTACAATCA
gyrB-F AGTAACGGATAACGGACGTGGTA
gyrB-R CCAACACCATGTAAACCACCAGAT
gyrB-P CCGCCACCGCCGAATTTACCACCA
spa-F GTAACGGCTTCATTCAAAGTCT
spa-R TCATAGAAAGCATTTTGTTGTTCT
spa-P AAAGACGACCCAAGCCAAAGCACT
a

Forward (F) and reverse (R) primers and the Taqman probe (P) for the ORFs are shown.

RNA was converted to cDNA, and microarray analysis was performed according to the manufacturer's instructions (Affymetrix expression analysis technical manual, Affymetrix, Inc., Santa Clara, Calif.) for antisense prokaryotic arrays. Briefly, 10 μg of total RNA that had been mixed with random hexamer primers (Invitrogen) was denatured at 70°C for 10 min and allowed to anneal at 25°C for 10 min. cDNA was synthesized using Superscript II reverse transcriptase (Invitrogen) in 1× first-strand synthesis buffer, dithiothreitol, deoxynucleoside triphosphates, and SUPERase-In (Ambion, Inc., Austin, Tex.). The mixture was incubated at 25°C for 10 min, 37°C for 60 min, and 42°C for 60 min. The reaction was stopped by incubating for 10 min at 70°C prior to degrading the RNA with 1 N NaOH for 30 min at 65°C and neutralizing with 1 N HCl. The cDNA was purified using a QIAquick PCR purification kit (Qiagen) and fragmented with DNase I in One-Phor-All buffer (Amersham Biosciences, Piscataway, N.J.). DNase I was inactivated by heating the reaction mixture for 10 min at 98°C. The fragmented cDNA products were labeled with biotin on the 3′ terminus using the Enzo BioArray terminal labeling kit with biotin ddUTP (Affymetrix, Inc.).

DNA microarray hybridization and analysis.

Labeled cDNA (1.5 μg) was hybridized to custom-made S. aureus GeneChips and detected according to the manufacturer's instructions for antisense prokaryotic arrays (Affymetrix, Inc.). The GeneChip used in these experiments included 7,723 qualifiers representing the consensus open reading frame (ORF) sequences identified in the genomes of the S. aureus strains N315, Mu50, COL, NCTC 8325, EMRSA-16 (strain 252), and MSSA- 476, as well as novel GenBank entries and N315 intergenic regions greater than 50 bp, After hybridization and staining, the arrays were scanned using the Agilent GeneArray laser scanner (Agilent Technologies, Palo Alto, Calif.). The data from duplicate experiments was normalized and analyzed using GeneSpring version 5.1 gene expression software (Silicon Genetics, Redwood City, Calif.). Genes were considered to be induced in a biofilm if they were determined to be present by Affymetrix algorithms in the biofilm condition and they were transcribed at a level at least twofold higher than the corresponding planktonic growth condition. Genes were considered downregulated in a biofilm if they were determined to be present in either planktonic condition and had an expression level no more than half of that observed in the corresponding planktonic growth condition. Differential expression in biofilms was judged to be significant on the basis of statistical analysis, namely, the t test with a P value of ≤0.05.

Real-time PCR.

To confirm the results of our microarray data, the relative expression levels of the arcA, pyrR, ureA, and spa genes in each growth condition were also determined by real-time PCR. Briefly, DNase-treated RNA was reverse transcribed using the iScript cDNA synthesis kit as described by the manufacturer (Bio-Rad Laboratories). A portion (1/20th) of each reaction mixture was then used for real-time PCR using an iCycler iQ real-time PCR detection system, gene-specific primers, and TaqMan probes corresponding to each ORF, and the iQ supermix (Bio-Rad). The sequences of the primers and TaqMan probes are shown in Table 1. Relative expression levels were determined by comparison to the level of gyrB expression in the same cDNA preparations.

RESULTS

Mutation of sarA, but not ica, results in a reduced capacity to form a biofilm in vitro.

In a previous report from our laboratory, we demonstrated that mutation of sarA in clinical isolates of S. aureus results in a reduced capacity to form a biofilm (5). Although our experiments did not address the mechanistic basis for this, Valle et al. (66) also observed that mutation of sarA results in a reduced capacity to form a biofilm and concluded that this was due to the impact of SarA on production of PIA (also known as PNAG). To more definitively address the role of PNAG in biofilm formation by our clinical isolates, we generated an S. aureus UAMS-1 ica mutant and assessed its ability to form a biofilm in vitro. Mutation of the ica locus was confirmed by Southern blotting (Fig. 1A), and the inability of the ica mutant to produce PNAG was confirmed by immunoblotting using PNAG-specific antisera (Fig. 1B). We found that mutation of ica, and the resulting inability to produce PNAG, had little impact on biofilm formation (Fig. 2). In contrast, mutation of sarA resulted in a reduced capacity to form a biofilm. The relative capacities of the UAMS-1 sarA and ica mutants to form a biofilm were evident both in our microtiter plate assay (Fig. 2A) and in flow cells (Fig. 2B). Under in vitro growth conditions, mutation of sarA resulted in reduced production of PNAG (Fig. 1B), which is consistent with the results of Valle et al. (66). However, the results observed with our ica mutant make it difficult to conclude that this would explain the biofilm-deficient phenotype of the UAMS-1 sarA mutant.

FIG. 1.

FIG. 1.

Confirmation of S. aureus UAMS-1 sarA and ica mutants. (A) Genomic DNA isolated from UAMS-1 (U1) and its corresponding ica or sarA mutants was digested with HpaI and blotted with probes corresponding to an internal fragment of the ica operon or the sarA locus. (B) Dot blot analysis of PNAG expression in S. aureus strain UAMS-1 and its corresponding ica and sarA mutants. Duplicate samples prepared as described in Materials and Methods were spotted onto membranes and analyzed using antiserum raised against S. aureus PNAG.

FIG. 2.

FIG. 2.

Biofilm formation in S. aureus UAMS-1 sarA and ica mutants in vitro. Biofilm formation under static (A) and flow (B) conditions was assessed as described in Materials and Methods. WT, wild type.

Mutation of sarA, but not ica, results in a reduced capacity to form a biofilm in vivo.

The results discussed above suggest that ica is not required for biofilm formation in at least some strains of S. aureus and that the reduced capacity of a UAMS-1 sarA mutant to form a biofilm in vitro is not a function of the impact of the sarA mutation on expression of the ica operon or production of PNAG. However, in S. aureus, it is well established that expression of the ica locus is tightly regulated and that it is preferentially expressed under in vivo conditions (42). This leaves open the possibility that the results we observed in vitro do not reflect the situation observed in vivo.

To examine this issue directly, we assessed the relative abilities of S. aureus UAMS-1 and its sarA and ica mutants to form a biofilm in vivo using a murine model of catheter-associated biofilm formation (31). The average numbers of bacteria obtained from explanted catheters at 10 days postinfection were 7.1 × 107 CFU per catheter in mice infected with UAMS-1 and 5.9 × 107 CFU per catheter in mice infected with the ica mutant (Fig. 3). In contrast, we recovered only 2.3 × 107 CFU per catheter from mice infected with the UAMS-1 sarA mutant. Although the sarA mutant was capable of colonizing the implanted catheter, the reduced recovery observed with the sarA mutant was statistically significant compared to the recovery for both UAMS-1 (P = 0.001) and its ica mutant (P = 0.022).

FIG. 3.

FIG. 3.

Biofilm formation in S. aureus UAMS-1 sarA and ica mutants in vivo. Bacteria were recovered from implanted catheters after 10 days in vivo. The number of bacteria recovered from the catheters was determined by plate count as described in Materials and Methods. The value that was statistically significantly different (P < 0.05) from the values for UAMS-1 and the ica mutant is indicated by the asterisk.

To further investigate the relative capacities of S. aureus UAMS-1 and its ica and sarA mutants to form a biofilm in vivo, we also performed experiments in which catheters were coinfected with equivalent numbers of both UAMS-1 and its ica mutant or UAMS-1 and its sarA mutant (in both cases, total inoculum of 2 × 105 CFU). In mice coinfected with UAMS-1 and its ica mutant, we recovered 3.4 × 107 CFU per catheter of the wild-type strain and 2.3 × 107 CFU per catheter of the ica mutant (Fig. 4A). These results confirm our in vitro experiments and demonstrate that UAMS-1 and its corresponding ica mutant have an equivalent capacity to form a biofilm not only in vitro but also in vivo. In contrast, when we examined mice coinfected with UAMS-1 and its sarA mutant, the number of wild-type cells recovered was similar to the number found in previous experiments (4.9 × 107 CFU), but the number of the UAMS-1 sarA mutant we recovered had decreased to an average of only 8.0 × 105 CFU per catheter (Fig. 4B). These results also confirm the results of our in vitro experiments. Moreover, the reduced recovery of the sarA mutant in coinfection experiments relative to in vivo experiments in which the sarA mutant was introduced without competition from the wild-type strain also indicates that the wild-type strain has a competitive advantage that further limits the capacity of a sarA mutant to form a biofilm in vivo.

FIG. 4.

FIG. 4.

Biofilm formation in coinfection experiments with S. aureus UAMS-1 and its sarA and ica mutants in vivo. Catheters were coinfected with equal numbers of UAMS-1 and its ica mutant (A) or UAMS-1 and its sarA mutant (B). The total number of bacteria in the inoculum and the total number recovered from each catheter was determined by plate count on nonselective medium. The number of bacteria resistant to tetracycline (A) or kanamycin (B) was subsequently determined by transferring individual colonies to selective media. Resistance to tetracycline and kanamycin is indicative of the ica and sarA mutations, respectively. Abbreviations: TetS and TetR, tetracycline sensitive and resistant, respectively; KanS and KanR, kanamycin sensitive and resistant, respectively.

Transcriptional profiling in S. aureus UAMS-1 planktonic cultures.

Taken together, the results discussed above confirm that a UAMS-1 sarA mutant has a reduced capacity to form a biofilm and that this is not a function of the impact of sarA on expression of the ica operon or production of PNAG. Moreover, the results observed with our in vitro models were consistent with the results observed in our in vivo model. Because sarA is a global regulator of gene expression in S. aureus, this suggests that other elements of the sarA regulon are also important in biofilm formation both in vitro and in vivo. Presumably, these elements could be identified by defining transcriptional changes observed within S. aureus biofilms and correlating these changes with experiments aimed at defining the sarA regulon. Because comprehensive transcriptional profiling of an S. aureus sarA mutant has been reported (19), we focused our efforts on defining the transcriptional changes that occur when UAMS-1 is grown within a biofilm.

Because biofilm formation by S. aureus UAMS-1 requires supplementation of the medium with both glucose and salt (5), we also added these supplements to the medium used for planktonic culture. We first wanted to investigate whether the current paradigm (e.g., preferential expression of surface proteins during the exponential phase, followed by a shift to exoprotein production in the postexponential phase) was altered by supplementation of the medium. As expected, we found that expression of genes encoding protein A, clumping factor B, collagen adhesion, coagulase, and fibronectin-binding protein (spa, clfB, cna, coa, and fnb, respectively) were upregulated in the exponential versus stationary phase (6, 44, 45) (Table 2). In contrast, expression of the gene (clfA) encoding a second fibrinogen-binding protein (ClfA) was upregulated in the postexponential phase. This is consistent with previous reports examining the temporal expression of clfA (72, 73). Expression of the secreted proteins was also elevated in stationary phase as expected. Specific examples include the genes encoding several cysteine proteases, the Clp proteinase, alpha-toxin, and the genes within the accessory gene regulator (agr) operon (sspA, sspB, sspC, clpC, hla, RNAII, and RNAIII, respectively). Taken together, these results confirm that our growth conditions accurately reflect the transition between the exponential and stationary growth phases and that this transition is not dramatically altered by supplementation of the medium in a manner that promotes biofilm formation.

TABLE 2.

Selected genes differentially expressed in exponential-phase versus stationary-phase cells

N315 ORFa Common namea Functiona ERb
N315-SA0144 cap5A Capsular polysaccharide synthesis enzyme 0.312
N315-SA0145 cap5B Capsular polysaccharide synthesis enzyme 0.306
N315-SA0146 cap5C Capsular polysaccharide synthesis enzyme 0.223
N315-SA0147 cap5D Capsular polysaccharide synthesis enzyme 0.244
N315-SA0148 cap5E Capsular polysaccharide synthesis enzyme 0.206
N315-SA0149 cap5F Capsular polysaccharide synthesis enzyme 0.204
N315-SA0150 cap5G Capsular polysaccharide synthesis enzyme 0.187
N315-SA0155 cap5L Capsular polysaccharide synthesis enzyme 0.277
N315-SA0156 cap5M Capsular polysaccharide synthesis enzyme 0.232
N315-SA0157 cap5N Capsular polysaccharide synthesis enzyme 0.267
N315-SA0158 cap5O Capsular polysaccharide synthesis enzyme 0.262
N315-SA0159 cap5P Capsular polysaccharide synthesis enzyme 0.267
N315-SA0899 sspC Cysteine protease 0.400
N315-SA0900 sspB Cysteine protease precursor 0.393
N315-SA1007 α-Hemolysin precursor 0.327
N315-SA1842 agrB-c Accessory gene regulator B 0.449
N315-SA1843 agrC-124c Accessory gene regulator C 0.269
N315-SA1844 agrA Accessory gene regulator A 0.289
N315-SA2336 clpC ATP-dependent Clp proteinase chain 0.284
agrB-3 99.5% protein IDc to AgrB 0.410
agrC-3 99.5% protein ID to AgrC 0.364
agrC-3c 99.5% protein ID to AgrC 0.431
agrD-3 100% protein ID to AgrD 0.389
cap8H 100% protein ID to Cap8H capsular polysaccharide synthesis enzyme 0.176
cap8I 100% protein ID Cap8I capsular polysaccharide synthesis enzyme 0.227
cap8J 100% protein ID Cap8J capsular polysaccharide synthesis enzyme 0.204
cap8K 100% protein ID to Cap8K capsular polysaccharide synthesis enzyme 0.226
86.7% protein ID to ClfA fibrinogen-binding protein A 0.476
sspA 99.4% protein ID to glutamic acid-specific protease prepropeptide truncated alpha-toxin 0.333
0.311
N315-SA0107 spa Immunoglobulin G-binding protein A precursor 5.084
N315-SA2423 clfB Clumping factor B 2.744
clfB 93.7% protein ID to ClfB (clumping factor B) 3.479
cna 98.7% protein ID to MW2612 collagen adhesin precursor 7.974
coa 99.8% protein ID to Coa coagulase 3.675
fnb homolog 87.8% protein ID to Fnb 3.975
spa 5.007
a

Based on the published sequence of S. aureus strain N315. For genes not present in N315, the gene name and description given are from the S. aureus strain COL genome, available from The Institute for Genomic Research website (http://www.tigr.org) or by the putative function.

b

Normalized values based on the expression ratio (ER), defined as the expression level in exponential-phase cells/expression level in stationary-phase cells.

c

ID, identity.

Transcriptional profiling of S. aureus UAMS-1 biofilms.

We identified a total of 580 genes that were expressed in an altered fashion when UAMS-1 was harvested from a mature biofilm. The greatest distinction, at least in terms of overall numbers of differentially expressed genes, was between the biofilm mode of growth and the exponential growth phase of planktonic culture (Table 3); however, a significant number of genes were also differentially expressed in comparison to stationary-phase cultures (Table 4). These findings clearly imply that S. aureus biofilms represent a unique growth condition by comparison to both exponential- and stationary-phase planktonic cultures. Indeed, we identified 48 genes whose expression was enhanced at least twofold in a biofilm in comparison to both exponential- and stationary-phase planktonic cultures (Table 5). These 48 genes included 30 genes located in six independent clusters as determined by their N315 ORF numbers. Included among these linked genes were the arginine deiminase cluster (arc; N315-SA2424-SA2428), a potassium-specific transport system (kdp; N315-SA1879-SA1881), the pyrimidine biosynthesis operon (pyr; N315-SA1041-SA1049), and the urease operon (ure; N315-SA2081-2088). Interestingly, only one gene in the ica locus (icaD) was found to be significantly upregulated in a biofilm, and this was limited to the comparison between biofilms and the stationary phase of planktonic growth (Table 4). This is consistent with a recent report concluding that ica expression is associated with the initial colonization of S. epidermidis in a foreign body infection model but not with its persistence (67).

TABLE 3.

Genes differentially expressed in a biofilm versus exponential growth phase

N315 ORFa Common namea Producta,b ERc
Cell envelope and cellular processes
    N315-SA1960 mtlF PTS system, mannitol-specific IIBC component 5.46
    N315-SA1882 kdpD Sensor protein KdpD 5.13
    N315-SA2311 Similar to NAD(P)H-flavin oxidoreductase 2.71
    N315-SA1156 ABC transporter (ATP-binding protein) homolog 2.68
    N315-SA0724 Similar to cell division inhibitor 0.496
    N315-SA2253 opp-1C Oligopeptide transporter putative membrane permease domain 0.490
    N315-SA0567 Similar to iron(III) ABC transporter permease protein 0.474
    N315-SA2216 Similar to ABC transporter, ATP-binding protein 0.471
    N315-SA0980 Similar to ferrichrome ABC transporter 0.471
    N315-SA0981 Similar to ferrichrome ABC transporter 0.468
    N315-SA0592 tagA Teichoic acid biosynthesis protein 0.462
    N315-SA1935 hmrA Similar to amidase 0.442
    N315-SA1169 γ-Aminobutyrate permease 0.425
    N315-SA0243 tagB Similar to teichoic acid biosynthesis protein B 0.423
    N315-SA0110 sirB Lipoprotein 0.399
    N315-SA2100 Similar to autolysin E 0.394
    N315-SA1458 lytH N-Acetylmuramoyl-l-alanine amidase 0.393
    N315-SA0109 sirC Lipoprotein 0.378
    N315-SA0106 lctP l-Lactate permease homolog 0.376
    N315-SA0682 Similar to ditripeptide ABC transporter 0.370
    N315-SA2053 Glucose uptake protein homolog 0.331
    N315-SA0479 nupC Pyrimidine nucleoside transport protein 0.315
    N315-SA0111 sirA Lipoprotein 0.312
    N315-SA2339 Similar to antibiotic transport-associated protein 0.291
    N315-SA0566 Similar to iron-binding protein 0.287
    N315-SA2233 Similar to integral membrane efflux protein 0.281
    N315-SA0325 glpT Glycerol-3-phosphate transporter 0.281
    N315-SA2112 Similar to sodium-dependent transporter 0.278
    N315-SA1025 mraY Phospho-N-muramic acid-pentapeptide translocase 0.265
    N315-SA0600 Similar to pyrimidine nucleoside transporter 0.255
    N315-SA1978 Similar to ferrichrome ABC transporter (permease) 0.203
    N315-SA0010 azlC Similar to amino acid permease 0.161
    N315-SA2300 Similar to glucarate transporter 0.141
    N315-SA0691 sstD Lipoprotein, similar to ferrichrome ABC transporter 0.126
    N315-SA0374 pbuX Xanthine permease 0.089
    N315-SA0579 Similar to Na+/H+ antiporter 0.080
    N315-SA0411 ndhF NADH dehydrogenase subunit 5 0.065
    N315-SA2302 stpC Similar to ABC transporter 0.046
    N315-SA2303 smpC Similar to membrane-spanning protein 0.041
Information pathways
    N315-SA1883 kdpE KDP operon transcriptional regulatory protein 5.42
    N315-SA2429 ArgR Similar to arginine repressor 3.92
    N315-SA2296 Similar to transcriptional regulator, MerR family 3.72
    N315-SA2418 Similar to two-component response regulator 2.13
    N315-SA0460 pth Peptidyl-tRNA hydrolase 0.490
    N315-SA0652 Similar to transcription regulation protein 0.452
    N315-SA1853 Similar to DNA mismatch repair protein MutS 0.445
    N315-SA1287 asnS Asparaginyl-tRNA synthetase 0.441
    N315-SA0348 Similar to transcription terminator 0.440
    N315-SA2358 Similar to transcriptional regulator (TetR/AcrR family) 0.382
    N315-SA1697 Similar to protein-tyrosine phosphatase 0.364
    N315-SA1120 Similar to transcription regulator GntR family 0.354
    N315-SA0298 pfoR Similar to regulatory protein PfoR 0.333
    N315-SA1550 tyrS Tyrosyl-tRNA synthetase 0.325
    N315-SA2482 pcp Pyrrolidone-carboxylate peptidase 0.297
    N315-SA1583 rot Repressor of toxins (Rot) 0.295
    N315-SA0653 fruR Similar to transcription repressor of fructose operon 0.229
    N315-SA0904 Probable ATL autolysin transcription regulator 0.191
    N315-SA1725 sspB Staphopain, cysteine proteinase 0.074
Intermediary metabolism
    N315-SA0328 Similar to NADH-dependent FMN reductase 7.25
    N315-SA0122 butA Acetoin (diacetyl)reductase 5.04
    N315-SA2297 Similar to GTP-pyrophosphokinase 3.25
    N315-SA1142 glpD Aerobic glycerol-3-phosphate dehydrogenase 2.97
    N315-SA0016 purA Adenylosuccinate synthase 2.37
    N315-SA2397 Similar to pyridoxal-phosphate-dependent aminotransferase 2.04
    N315-SA2001 Similar to oxidoreductase, aldo/keto reductase family 2.01
    N315-SA1201 trpD Anthranilate phosphoribosyltransferase 0.495
    N315-SA1685 mutY Similar to A/G-specific adenine glycosylase 0.481
    N315-SA2111 Similar to phosphoglycolate phosphatase 0.475
    N315-SA1052 gmk Guanylate kinase homolog 0.462
    N315-SA2279 Similar to phosphomannomutase 0.458
    N315-SA0902 HisC homolog 0.448
    N315-SA0177 argJ Arginine biosynthesis bifunctional protein homolog 0.436
    N315-SA1310 ansA Probable l-asparaginase, gene-ansA 0.429
    N315-SA1309 cmk Cytidylate kinase 0.427
    N315-SA1749 Similar to aspartate transaminase protein 0.423
    N315-SA2140 Similar to esterase 0.412
    N315-SA0507 Similar to N-acyl-l-amino acid amidohydrolase 0.407
    N315-SA0568 Similar to l-2-haloalkanoic acid dehalogenase 0.374
    N315-SA2213 bioB Biotin synthase 0.373
    N315-SA0514 Similar to deoxypurine kinase 0.356
    N315-SA1121 Similar to processing proteinase homolog 0.340
    N315-SA1203 trpF Phosphoriborylanthranilate isomerase 0.323
    N315-SA2342 thgA Similar to O-acetyltransferase 0.280
    N315-SA0511 Similar to UDP-glucose 4-epimerase-related protein 0.256
    N315-SA1202 trpC Indole-3-glycerol phosphate synthase 0.246
    N315-SA2395 ldh l-Lactate dehydrogenase 0.213
    N315-SA0180 bmQ Similar to branched-chain amino acid transport system carrier protein 0.192
    N315-SA1200 trpG Anthranilate synthase component II 0.120
    N315-SA1199 trpE Similar to anthranilate synthase component I 0.101
    N315-SA0373 xprT Xanthine phosphoribosyltransferase 0.077
Other functions
    N315-SA0899 sspC Cysteine protease 7.24
    N315-SA0900 sspB Cysteine protease precursor 7.03
    N315-SA0150 cap5G Capsular polysaccharide synthesis enzyme 4.80
    N315-SA2006 Similar to MHC class II analog 4.69
    N315-SA0149 cap5F Capsular polysaccharide synthesis enzyme 4.58
    N315-SA0148 cap5E Capsular polysaccharide synthesis enzyme 4.05
    N315-SA0146 cap5C Capsular polysaccharide synthesis enzyme 3.31
    N315-SA0147 cap5D Capsular polysaccharide synthesis enzyme 3.30
    N315-SA0145 cap5B Capsular polysaccharide synthesis enzyme Cap5B 2.43
    N315-SA0144 cap5A Capsular polysaccharide synthesis enzyme 2.34
    N315-SA0841 Similar to cell surface protein Map-w 2.22
    N315-SA1709 Similar to ferritin 2.18
    N315-SA0754 Similar to lactococcal prophage ps3 protein 05 0.498
    N315-SA1835 int Similar to integrase (pathogenicity island SaPln1), gene = int 0.431
    N315-SA1559 Similar to smooth muscle caldesmon 0.426
    N315-SA0797 nifU-3 Similar to nitrogen fixation protein NifU 0.407
    N315-SA0780 Similar to hemolysin 0.331
    N315-SA0746 nuc Staphylococcal nuclease 0.170
    N315-SA1766 HP (bacteriophage φN315) 0.069
    N315-SA1775 Similar to scaffolding protein (bacteriophage φN315) 0.048
    N315-SA1765 HP (bacteriophage φN315) 0.042
    N315-SA1777 HP (bacteriophage φN315) 0.038
    N315-SA1771 HP (bacteriophage φN315) 0.036
    N315-SA1762 HP (bacteriophage φN315) 0.036
Similar to unknown proteins
    N315-SA0326 CHP (lactoylglutathione lyase and related lyases) 7.94
    N315-SA0327 CHP (flavin-dependent oxidoreductases) 7.50
    N315-SA2479 CHP 5.04
    N315-SA0007 Predicted sugar kinase 4.57
    N315-SA0380 CHP (pathogenicity island SaPln2) 3.07
    N315-SA0381 CHP (pathogenicity island SaPln2) 2.80
    N315-SA1235 CHP 2.04
    N315-SA1890 CHP 2.03
    N315-SA0230 CHP 2.02
    N315-SA0941 CHP 0.497
    N315-SA0467 CHP (predicted ATPase of the PP-loop superfamily) 0.489
    N315-SA1838 CHP (predicted metal-dependent membrane protease) 0.481
    N315-SA2487 rarD Similar to RarD protein 0.462
    N315-SA1696 CHP 0.453
    N315-SA1448 CHP (TPR repeat-containing proteins) 0.452
    N315-SA0329 CHP 0.445
    N315-SA2328 CHP (putative effector of murein hydrolase, LrgB) 0.434
    N315-SA0979 CHP 0.433
    N315-SA1928 HP 0.421
    N315-SA1601 crcB CHP (integral membrane protein) 0.404
    N315-SA2377 CHP 0.398
    N315-SAS081 CHP (ATPase involved in DNA repair) 0.391
    N315-SA0334 CHP (Sec-independent protein secretion pathway) 0.378
    N315-SA2305 CHP 0.372
    N315-SA0840 CHP (phospholipid-binding protein) 0.370
    N315-SA1903 CHP 0.366
    N315-SA0543 CHP (uncharacterized BCR) 0.353
    N315-SA0413 CHP 0.332
    N315-SA0341 Similar to low-temperature requirement A protein 0.329
    N315-SA0345 CHP (methionine synthase I) 0.324
    N315-SA0257 CHP (SAM-dependent methyltransferases) 0.323
    N315-SA2096 CHP 0.320
    N315-SA0773 CHP (predicted membrane protein) 0.312
    N315-SA2212 Similar to 8-amino-7-oxononanoate synthase 0.303
    N315-SA1252 CHP (histone acetyltransferase) 0.298
    N315-SA2452 CHP (domain typically associated with flavoprotein oxygenases) 0.282
    N315-SA0870 CHP (predicted permease) 0.267
    N315-SA0335 CHP (Sec-independent protein secretion pathway) 0.265
    N315-SA0556 CHP 0.203
    N315-SA0753 CHP (lysine efflux permease) 0.179
    N315-SA2219 CHP (uncharacterized membrane protein) 0.133
    N315-SA0739 CHP 0.129
    N315-SAS001 CHP 0.128
    N315-SA0412 CHP 0.078
No similarity
    N315-SAS016 HP 12.74
    N315-SA0883 HP 5.97
    N315-SA1233 HP 2.25
    N315-SA0414 HP 0.496
    N315-SA2126 HP 0.495
    N315-tRNA12 tRNA-Pro 0.491
    N315-SA1943 HP 0.491
    N315-SA1215 HP 0.486
    N315-tRNA11 tRNA-Arg 0.479
    N315-SA0613 HP 0.470
    N315-SA2485 HP 0.457
    N315-SA0088 HP 0.443
    N315-tRNA57 tRNA-Lys 0.416
    N315-tRNA47 tRNA-Leu 0.409
    N315-SA1607 HP 0.402
    N315-SA0955 HP 0.372
    N315-tRNA07 tRNA-Thr 0.355
    N315-SA0105 HP 0.344
    N315-SA2118 HP 0.335
    N315-SA0336 HP 0.326
    N315-SA0363 HP 0.321
    N315-SA2055 HP 0.321
    N315-SA2224 HP 0.309
    N315-SA2249 HP 0.259
    N315-SA0749 HP 0.233
    N315-tRNA06 tRNA-Val 0.219
    N315-SA0748 HP 0.214
    N315-SA0889 HP 0.179
    N315-SA2488 HP 0.157
    N315-SA1778 HP (bacteriophage φN315) 0.089
    N315-SA1768 HP (bacteriophage φN315) 0.050
    N315-SA1770 HP (bacteriophage φN315) 0.050
    N315-SA1774 HP (bacteriophage φN315) 0.046
    N315-SA1776 HP (bacteriophage φN315) 0.044
    N315-SAS060 HP (bacteriophage φN315) 0.041
    N315-SA1769 HP (bacteriophage φN315) 0.040
    N315-SA1773 HP (bacteriophage φN315) 0.032
    N315-SA1772 HP (bacteriophage φN315) 0.031
    N315-SA1767 HP (bacteriophage φN315) 0.027
No N315 ORF
set5 100% protein ID Set5, exotoxin 5, and HsdM-like protein gene 40.43
sspA 99.4% protein ID S. aureus glutamic acid-specific protease 7.19
cap8H 100% protein ID to capsular polysaccharide synthase enzyme Cap8H 5.11
cap8J 100% protein ID to capsular polysaccharide synthesis enzyme Cap8J 4.52
58.1% protein ID malofactic enzyme, Oenococcus oeni bacteria 3.05
68.1% protein ID SA0329 CHP 2.39
92.8% protein ID MW1748 HP 2.25
84.4% protein ID MW0360 HP 2.16
96.7% protein ID to MW2134 HP 0.500
95.6% protein ID SA2230 0.473
58.4% protein ID BH3950 transposase (10), Bacillus halodurans 0.467
49.1% protein ID MW2618 0.463
COL-SA1788 HP 0.434
87.6% protein ID MW0584 0.423
COL-SA2299 HP 0.421
Similar to splE 62.6% protein ID serine protease SplE 0.420
97.7% protein ID to SA1559 0.419
26.9% protein ID MW2498 0.418
COL-SA0866 HP 0.411
Serine protease 55.4% protein ID serine protease SplB 0.409
96% protein ID MW2325 0.387
95.5% protein ID MW0355 HP 0.371
22.2% protein ID SA0283 HP 0.345
89.4% protein ID SA0553 CHP 0.338
78.2% protein ID MW1720 HP 0.315
COL-SA1556 HP 0.311
99.4% protein ID to MW0053 CHP 0.308
45.9% protein ID BH3950 transposase (10), Bacillus halodurans 0.290
97.7% protein ID MW0355 HP 0.289
splB 97.7% protein ID to serine protease SplB 0.280
79.2% protein ID MW1043 HP 0.270
COL-SA2728 HP 0.237
99.5% id to SAV1992 HP 0.209
37.3% protein ID MW1769 HP 0.202
COL-SA1140 sai-1 29-kDa cell surface protein 0.164
85.2% protein ID MW1042 HP 0.141
99.5% protein ID MWP018 0.138
99.8% protein ID MWP016 S. aureus plasmid pMW2 0.096
93.4% protein ID SAV1953 φPVL ORF 20 and 21 homolog 0.079
Mu50-SAV1953 φPVL ORF 20 and 21 homolog 0.071
100% protein ID to MWP017 HP 0.070
100% protein ID MW1894 HP 0.060
78% protein ID MW1892 HP 0.055
83.2% protein ID SA1763 HP 0.032
100% protein ID SAP019 HP, S. aureus N315 plasmid N315B 0.031
a

Based on the published sequence of S. aureus strain N315. For genes not present in N315, the gene name and description given are from the S. aureus strain COL genome, available from The Institute for Genomic Research website (http://www.tigr.org) or by the putative function.

b

Abbreviations: PTS, phosphotransferase; HP, hypothetical protein; CHP, conserved hypothetical protein; SAM, S-adenosylmethionine; ID, identity.

c

Normalized values based on the expression ratio (ER), which is defined as the expression level in exponential-phase cells/expression level in stationary-phase cells.

TABLE 4.

Genes differentially expressed in a biofilm versus stationary growth phase

N315 ORFa Common namea Producta,b ERc
Cell envelope and cellular processes
    N315-SA0655 fruA Fructose-specific permease 14.43
    N315-SA0263 Similar to proton antiporter efflux pump 9.55
    N315-SA2142 semB Similar to multidrug resistance protein 4.96
    N315-SA0293 Similar to formate transporter NirC 4.64
    N315-SA1140 glpF Glycerol uptake facilitator 4.40
    N315-SA2185 narG Respiratory nitrate reductase alpha chain 4.08
    N315-SA2183 narJ Similar to nitrate reductase delta chain 3.87
    N315-SA2184 narH Nitrate reductase beta chain NarH 3.56
    N315-SA2053 Glucose uptake protein homolog 3.55
    N315-SA0166 Similar to nitrate transporter 3.39
    N315-SA0167 Similar to membrane lipoprotein SrpL 3.29
    N315-SA0702 llm Lipophilic protein affecting bacterial lysis rate and methicillin resistance level 2.83
    N315-SA2222 Similar to bicyclomycin resistance protein TcaB 2.75
    N315-SA0411 ndhF NADH dehydrogenase subunit 5 2.41
    N315-SA2176 narK Nitrite extrusion protein 2.16
    N315-SA1960 mtlF PTS system, mannitol-specific IIBC component 0.486
    N315-SA1381 pbp3 Penicillin-binding protein 3 0.482
    N315-SA1219 Similar to phosphate ABC transporter 0.463
    N315-SA2311 Similar to NAD(P)H-flavin oxidoreductase 0.460
    N315-SA0367 Similar to nitroflavin reductase 0.410
    N315-SA1982 Similar to transporter 0.397
    N315-SA0260 Similar to ribose transporter RbsU 0.385
    N315-SA2074 modA Probable molybdate-binding protein 0.384
    N315-SA1848 amt Probable ammonium transporter 0.343
    N315-SA0138 Similar to alkylphosphonate ABC transporter 0.323
    N315-SA2203 EmrB/QacA subfamily Similar to multidrug resistance protein 0.304
    N315-SA0420 Similar to ABC transporter ATP-binding protein 0.286
    N315-SA0422 Similar to lactococcal lipoprotein 0.272
    N315-SA0421 Similar to ABC transporter permease protein 0.256
    N315-SA0589 Similar to ABC transporter ATP-binding protein 0.163
    N315-SA0849 Similar to peptide-binding protein OppA 0.148
Information pathways
    N315-SA0653 fruR Similar to transcription repressor of fructose operon 14.57
    N315-SA0476 Similar to transcription regulator GntR family 4.98
    N315-SA1058 def Similar to polypeptide deformylase 2.35
    N315-SA0460 pth Peptidyl-tRNA hydrolase 2.07
    N315-SA1516 phoP Alkaline phosphatase synthesis transcriptional regulatory protein 0.479
    N315-SA0130 Similar to trehalose operon transcriptional repressor 0.477
    N315-SA1805 Repressor homolog (bacteriophage φN315) 0.390
    N315-SAS042 rpmG 50S ribosomal protein L33 0.362
    N315-SA1394 glyS Glycyl-tRNA synthetase 0.351
    N315-SA1149 glnR Glutamine synthetase repressor 0.319
    N315-SA1360 Xaa-Pro dipeptidase 0.294
Intermediary metabolism
    N315-SA0654 fruB Fructose-1-phosphate kinase 17.08
    N315-SA1959 glmS Glucosamine-fructose-6-phosphate aminotransferase 9.27
    N315-SA0143 adhE Alcohol-acetaldehyde dehydrogenase 5.79
    N315-SA2186 nasF Uroporphyrin III C-methyl transferase 4.56
    N315-SA2187 nasE Assimilatory nitrite reductase 4.15
    N315-SA1929 pyrG CTP synthase 3.97
    N315-SA2188 nirB Nitrite reductase 3.38
    N315-SA0973 kdtB Phosphopantetheine adenyltransferase homolog 2.29
    N315-SA0572 Similar to esterase or lipase 0.493
    N315-SA0528 Similar to hexulose-6-phosphate synthase 0.484
    N315-SA1231 dal Similar to alanine racemase 0.471
    N315-SA2120 Similar to amino acid amidohydrolase 0.464
    N315-SA0008 hutH Histidine ammonia-lyase 0.461
    N315-SA1584 Lysophospholipase homolog 0.456
    N315-SA1230 hipO Hippurate hydrolase 0.435
    N315-SAS044 dmpI 4-Oxalocrotonate tautomerase 0.432
    N315-SA1225 lysC Aspartokinase II 0.429
    N315-SA1229 dapD Tetrahydrodipicolinate acetyltransferase 0.423
    N315-SA0258 rbsK Probable ribokinase 0.422
    N315-SA0820 glpQ Glycerophosphoryl diester phosphodiesterase 0.419
    N315-SA0181 entB Similar to isochorismatase 0.416
    N315-SA0512 ilvE Branched-chain amino acid aminotroansferase homolog 0.415
    N315-SA2204 gpm Phosphoglycerate mutase 0.408
    N315-SA1227 dapA Dihydrodipicolinate synthase 0.406
    N315-SA2155 mqo Similar to malate:quinone oxidoreductase 0.404
    N315-SA1724 purB Adenylosuccinate lyase 0.394
    N315-SA0304 nanA N-Acetylneuraminate lyase subunit 0.384
    N315-SA1150 glnA Glutamine-ammonia ligase 0.362
    N315-SA1531 ald Alanine dehydrogenase 0.357
    N315-SA1228 dapB Dihydrodipicolinate reductase 0.354
    N315-SA2125 hutG Similar to formiminoglutamase 0.353
    N315-SA0098 Similar to aminoacylase 0.346
    N315-SA2127 rpiA Similar to ribose 5-phosphate isomerase 0.341
    N315-SA1226 asd Aspartate semialdehyde dehydrogenase 0.341
    N315-SA1545 serA d-3-Phosphoglycerate dehydrogenase 0.341
    N315-SA0679 hisC Similar to histidinol-phosphate aminotransferase 0.337
    N315-SA0658 Similar to plant metabolite dehydrogenases 0.335
    N315-SA0656 nagA Probable N-acetylglucosamine-6-phosphate deacetylase 0.333
    N315-SA1184 citB Aconitate hydratase 0.329
    N315-SA0915 folD FolD bifunctional protein 0.326
    N315-SA1858 ilvD Dihydroxy-acid dehydratase 0.280
    N315-SA0430 gltB Glutamate synthase large subunit 0.263
    N315-SA1614 menC o-Succinylbenzoic acid synthetase 0.234
    N315-SA0431 gltD NADH-glutamate synthase small subunit 0.199
    N315-SA1553 fhs Formyltetrahydrofolate synthetase 0.168
    N315-SA0016 purA Adenylosuccinate synthase 0.117
    N315-SA0926 purD Phosphoribosylamine-glycine ligase 0.049
    N315-SA0917 purK Phosphoribosylaminoimidazole carboxylase carbon dioxide fixation chain 0.034
    N315-SA0916 purE Similar to phosphoribosylaminoimidazole carboxylase 0.030
    N315-SA0918 purC Phosphoribosylaminoimidazolesuccinocarboxamide synthetase homolog 0.023
Other functions
    N315-SA2460 icaD Intercellular adhesion protein D 34.06
    N315-SA1898 Similar to SceD precursor 26.36
    N315-SA2206 sbi IgG-binding protein 26.16
    N315-SA1000 Similar to fibrinogen-binding protein 11.51
    N315-SA2097 Similar to SsaA Similar to secretory antigen precursor 4.74
    N315-SA2164 Similar to phage infection protein precursor 2.80
    N315-SA1382 sodA Superoxide dismutase SodA 0.497
    N316-SA1606 Plant metabolite dehydrogenase homolog 0.480
    N315-SA0841 Similar to cell surface protein Map-w 0.448
    N315-SA1549 htrA Similar to serine proteinase Do, heat shock protein 0.434
    N315-SA2406 gbsA Glycine betaine aldehyde dehydrogenase 0.399
    N315-SA0659 Similar to CsbB stress response protein 0.390
    N315-SA1312 ebpS Elastin-binding protein 0.363
    N315-SA0755 Similar to general stress protein 170 0.340
    N315-SA1170 katA Catalase 0.309
    N315-SA0091 plc 1-Phosphatidylinositol phosphodiesterase precursor 0.300
    N315-SA2405 betA Choline dehydrogenase 0.252
Similar to unknown proteins
    N315-SA0213 CHP 17.26
    N315-SA2256 CHP 4.76
    N315-SA0341 HP similar to low-temperature requirement A protein 3.95
    N315-SA1176 CHP 3.05
    N315-SA0929 CHP 3.04
    N315-SA1431 CHP 3.00
    N315-SA1340 CHP (lactoylglutathione lyase) 2.53
    N315-SAS027 CHP 2.50
    N315-SA1932 Similar to HP T13D8.31 Arabidopsis thaliana 2.35
    N315-SA1464 yajC CHP (preprotein translocase subunit YajC) 2.30
    N315-SA1540 CHP (GAF domain-containing protein) 2.24
    N315-SA0165 Similar to α-helical coiled-coil protein SrpF 2.16
    N315-SA0114 CHP 2.10
    N315-SA0529 CHP (predicted sugar phosphate isomerase involved in capsule formation, GutQ) 0.486
    N315-SA1019 CHP 0.467
    N315-SA1737 CHP (3-carboxymuconate cylase) 0.466
    N315-SA0801 CHP (IscA) 0.463
    N315-SA1543 CHP (predicted redox protein, regulator of disulfide bond formation) 0.462
    N315-SA1380 CHP (5-formyltetrahydrofolate cyclo-ligase) 0.454
    N315-SA1129 CHP (predicted HD superfamily hydrolase) 0.447
    N315-SA0861 CHP (hemoglobin-like proteins) 0.443
    N315-SA0230 CHP 0.438
    N315-SA1280 CHP 0.430
    N315-SA0957 CHP 0.429
    N315-SA1331 CHP (predicted oxidoreductases) 0.393
    N315-SA1689 CHP 0.384
    N315-SA0513 CHP (predicted phosphatases, Gph) 0.379
    N315-SA2367 CHP (predicted hydrolases or acyltransferases) 0.356
    N315-SA1167 CHP (predicted hydrolases of the HAD superfamily) 0.337
    N315-SA1690 CHP 0.324
    N315-SA0089 Similar to DNA helicase 0.315
    N315-SA0873 CHP 0.315
    N315-SA1544 Similar to soluble hydrogenase 42-kDa subunit 0.314
    N315-SA0741 CHP (predicted acetyltransferase) 0.310
    N315-SA0362 CHP 0.310
    N315-SA1281 CHP 0.310
    N315-SA0649 CHP (predicted DNA-binding proteins with PD1-like DNA-binding motif) 0.286
    N315-SA0407 CHP (chromosome segregation ATPases) 0.254
    N315-SA1611 CHP (dipeptidyl aminopeptidases/acylaminoacyl-peptidases) 0.227
    N315-SA0919 CHP (phosphoribosylformylglycinamidine [FGAM] synthase) 0.027
No similarity
    N315-SA0663 HP 7.97
    N315-SA2281 HP 5.10
    N315-SA0779 HP 3.44
    N315-SA2376 HP 3.22
    N315-SA0885 HP 3.16
    N315-SA1670 HP 2.80
    N315-SA2126 HP 2.62
    N315-SA0336 HP 2.59
    N315-SA0571 HP 2.37
    N315-SA2058 HP 2.17
    N315-SA0397 lpl2 HP (pathogenicity island SaPln2) 0.488
    N315-SAS031 HP 0.480
    N315-SA1168 HP 0.472
    N315-SA0372 HP 0.468
    N315-SA0404 lpl8 HP (pathogenicity island SaPln2) 0.447
    N315-SA1319 HP 0.421
    N315-SA0090 HP 0.366
    N315-SA1546 HP 0.293
    N315-SA0406 HP 0.258
    N315-SA2497 HP 0.208
    N315-SA0408 HP 0.198
    N315-SA2496 HP 0.188
    No N315 ORF 72.4% protein ID to MW1041 25.20
    COL-SA0674 HP 6.12
    COL-SA1165 HP 5.57
98.1% protein ID to MW2274 CHP 5.33
98.1% protein ID to NasE assimilatory nitrite reductase 4.25
57% protein ID to spyM18_1050 HP, S. pyogenes MGAS8232 3.55
88.2% protein ID to MW2323 3.11
100% protein ID to SAP023 S. aureus N315 plasmid pN315B 2.32
100% protein ID to MW2396 0.497
    COL-SA1345 HP 0.491
100% protein ID to SA1320 HP 0.482
100% protein ID to lpl11 HP, S. aureus MW2 0.465
    COL-SA2676 LPXTG LPXTG-motif cell wall anchor domain protein 0.456
91.5% protein ID to lpl2 HP, S. aureus N315 0.447
    COL-SA0293 CHP 0.444
39.1% protein ID to lin05-11 Listeria innocua 0.436
26.6% protein ID to LigW 5-carboxyvanillate decarboxylase, Sphingomonas paucimobilis 0.416
92.7% protein ID to Lpl7 HP, S. aureus N315 0.405
binL 99.5% protein ID to BinL DNA-invertase, S. aureus plasmid pMW2 0.394
94.8% protein ID to BinL DNA invertase, S. aureus plasmid pMW2 0.392
    COL-SA0601 HP 0.372
89.8% protein ID to Lpl10 HP, S. aureus MW2 0.364
25.8% protein ID to MA2121 CHP, Methanosarcina acetivorans C2A 0.363
88% protein ID to MW1374 CHP 0.363
74.8% protein ID to lpl5 HP, S. aureus N315 0.331
    COL-SA1343 HP 0.328
48.6% protein ID to ycnB homolog of multidrug resistance protein B, B. subtilis 0.309
71% protein ID to MW1201 HP 0.282
93.5% protein ID to MW0402 HP 0.276
84% protein ID to lpl2 HP, S. aureus N315 0.261
35.9% protein ID to Cgl0945 putative multicopper oxidases, Corynebacterium glutamicum 0.259
57.8% protein ID to CopB ATPase, Enterococcus hirae 0.214
64.2% protein ID to SA0753 CHP 0.191
33.6% protein ID to RtxC, Bradyrhizobium elkanii 0.158
a

Based on the published sequence of S. aureus strain N315. For genes not present in N315, the gene name and description given are from the S. aureus strain COL genome, available from The Institute for Genomic Research website (www.tigr.org or by the putative function.

b

Abbreviations: PTS, phosphotransferase; IgG, immunoglobulin G; CHP, conserved hypothetical protein; HP, hypothetical protein; HAD, haloacid dehalogenase-family protein; ID, identity.

c

Normalized values based on the expression ratio (ER), which is defined as the expression level in exponential-phase cells/expression level in stationary-phase cells.

TABLE 5.

Genes differentially expressed in a biofilm versus exponential and stationary phase

N315 ORFa Common namea Producta,b ER vs EPc ER vs SPc
Cell envelope and cellular processes
    N315-SA2426 arcD Arginine/ornithine antiporter 59.91 5.49
    N315-SA1881 kdpA Probable potassium-transporting ATPase A chain 51.58 11.30
    N315-SA1880 kdpB Probable potassium-transporting ATPase B chain 30.99 9.53
    N315-SA1042 pyrP Uracil permease 25.54 7.97
    N315-SA1879 kdpC Probable potassium-transporting ATPase C chain 20.82 8.40
    N315-SA0417 Similar to sodium-dependent transporter 7.33 16.49
    N315-SA2081 Similar to urea transporter 6.49 4.65
    N315-SA1688 Similar to teichoic acid translocation ATP-binding protein TagH 0.47 0.42
    N315-SA0233 PTS enzyme, maltose and glucose specific, factor II homolog 0.44 0.15
    N315-SA0848 oppF Oligopeptide transport system ATP-binding protein homolog 0.42 0.16
    N315-SA0847 oppD Oligopeptide transport system ATP-binding protein homolog 0.39 0.15
    N315-SA0845 oppB Oligopeptide transport system permease protein 0.38 0.16
    N315-SA2242 CHP (predicted permease) 0.37 0.35
    N315-SA0846 oppC Similar to oligopeptide transport system permease protein 0.37 0.17
    N315-SA0758 Similar to thioredoxin 0.28 0.34
    N315-SA2261 Similar to efflux pump 0.25 0.45
    N315-SA2132 Similar to ABC transporter (ATP-binding protein) 0.24 0.34
    N315-SA0217 Similar to periplasmic iron-binding protein BitC 0.23 0.24
    N315-SA1699 Similar to transporter 0.20 0.34
    N315-SA1987 opuD Glycine betaine transporter OpuD homolog 0.11 0.21
Information pathways
    N315-SA2424 acrR Similar to transcription regulator Crp/Fnr family protein 48.16 9.87
    N315-SA1041 pyrR Pyrimidine operon repressor chain A 16.05 6.82
    N315-SA2320 Similar to regulatory protein PfoR 9.87 4.66
    N315-SA2502 rnpA RNase P protein component 3.82 5.26
    N315-SA2134 Similar to DNA 3-methyladenine glycosidase 0.48 0.40
    N315-SA0815 Peptidyl-prolyl cis-trans isomerase homolog 0.44 0.49
    N315-SA2278 Similar to mutator protein MutT 0.41 0.41
    N315-SA1626 hsdM Type I restriction enzyme homolog (SaPln3) 0.41 0.48
    N315-SA0097 Similar to transcription regulator AraC/XylS family 0.40 0.49
    N315-SA2144 Similar to transcriptional regulator (TetR/AcrR family) 0.39 0.37
    N315-SA0189 hsdR Probable type I restriction enzyme restriction chain 0.30 0.36
    N315-SA1806 Probable ATP-dependent helicase (bacteriophage φN315) 0.29 0.25
Intermediary metabolism
    N315-SA2427 arcB Ornithine transcarbamoylase 124.22 5.83
    N315-SA2428 arcA Arginine deiminase 114.59 5.45
    N315-SA2425 arcC Carbamate kinase 37.87 5.76
    N315-SA1044 pyrC Dihydroorotase 17.84 6.11
    N315-SA1045 carA Carbamoyl-phosphate synthase small chain 13.13 6.03
    N315-SA1047 pyrF Orotidine-5-phosphate decarboxylase 10.65 4.93
    N315-SA1046 carB Carbamoyl-phosphate synthase large chain 10.50 4.97
    N315-SA2082 ureA Urease gamma subunit 9.68 3.12
    N315-SA2083 ureAB Urease beta subunit 9.29 2.97
    N315-SA1048 pyrE Orotate phosphoribosyltransferase 8.94 4.87
    N315-SA2319 sdhB Similar to beta-subunit of l-serine dehydratase 8.90 4.84
    N315-SA2084 ureC Urease alpha subunit 8.40 3.50
    N315-SA2086 ureF Urease accessory protein 8.34 3.33
    N315-SA2088 ureD Urease accessory protein 7.72 3.67
    N315-SA2085 ureE Urease accessory protein 7.26 2.82
    N315-SA2087 ureG Urease accessory protein 6.78 3.16
    N315-SA2318 sdhA Similar to l-serine dehydratase 6.49 4.24
    N315-SA1043 pyrB Aspartate transcarbamoylase chain A 4.69 6.34
    N315-SA2007 Similar to α-acetolactate decarboxylase 4.40 3.09
    N315-SA0821 argH Argininosuccinate lyase 3.93 14.65
    N315-SA0822 argG Argininosuccinate synthase 3.52 13.48
    N315-SA2008 budB α-Acetolactate synthase 3.29 2.53
    N315-SA1155 cls Cardiolipin synthetase homolog 2.73 2.28
    N315-SA1160 nuc Thermonuclease 2.30 2.47
    N315-SA2258 Similar to diaminopimelate epimerase 2.03 2.60
    N315-SA1940 deoD Purine nucleoside phosphorylase 0.46 0.49
    N315-SA1615 menE O-Succinylbenzoic acid-CoA ligase 0.44 0.21
    N315-SA0925 purH Bifunctional purine biosynthesis protein 0.43 0.04
    N315-SA0241 Similar to 4-diphosphocytidyl-2C-methyl-d-erythritol synthase 0.42 0.46
    N315-SA0963 pyc Pyruvate carboxylase 0.41 0.26
    N315-SA0011 Similar to homoserine-o-acetyltransferase 0.39 0.44
    N315-SA0534 atoB Acetyl-CoA c-acetyltransferase 0.38 0.43
    N315-SA0920 purQ Phosphoribosylformylglycinamidine synthase I 0.37 0.02
    N315-SA0923 purM Phosphoribosylformylglycinamidine cyclo-ligase 0.34 0.03
    N315-SA0924 purN Phosphoribosylglycinamide formyltransferase 0.34 0.03
    N315-SA0242 Similar to xylitol dehydrogenase 0.34 0.44
    N315-SA0921 purL Phosphoribosylformylglycinamidine synthetase 0.33 0.03
    N315-SA0922 purF Phosphoribosylpyrophosphate amidotransferase 0.33 0.03
    N315-SA0344 metE 5-Methyltetrahydropteroyltriglutamate-homocysteine 0.28 0.43
Methyltransferase
    N315-SA0022 Similar to 5′-nucleotidase 0.26 0.29
    N315-SA1814 Similar to succinyl-diaminopimelate desuccinylase 0.23 0.20
    N315-SA0266 CHP (ABC-type multidrug transport system, ATPase component) 0.15 0.23
Other functions
    N315-SA2353 ssaA Similar to secretory antigen precursor 3.64 5.70
    N315-SA0270 ssaA Similar to secretory antigen precursor 0.42 0.39
    N315-SA1629 splC Serine protease 0.39 0.49
    N315-SA0107 spa Immunoglobulin G-binding protein A precursor 0.01 0.04
Similar to unknown proteins
    N315-SA0023 CHP 0.50 0.35
    N315-SA0814 kapB CHP 0.48 0.44
    N315-SA1692 CHP (putative intracellular protease/amidase) 0.46 0.37
    N315-SA0518 CHP (predicted flavoprotein) 0.46 0.43
    N315-SA1612 CHP (NTP pyrophosphohydrolases) 0.34 0.17
    N315-SA1133 CHP 0.32 0.44
    N315-SA2371 CHP 0.30 0.26
    N315-SA0559 CHP (histone acetyltransferase HPA2 and related acetyltransferases) 0.29 0.36
    N315-SA0872 CHP (enterochelin esterase and related enzymes) 0.29 0.44
    N315-SA2131 CHP (ABC-type Na+ efflux pump, permease component) 0.27 0.38
    N315-SA1733 CHP 0.26 0.29
    N315-SA2322 CHP (permeases of the drug/metabolite transporter superfamily) 0.22 0.45
    N315-SA0269 HP 0.13 0.10
    N315-SA0359 CHP (uncharacterized membrane protein) 0.13 0.37
No similarity
    N315-SA1049 HP 7.42 4.41
    N315-SA0575 HP 2.14 2.10
    N315-SA1152 HP 0.46 0.46
    N315-SA0752 HP 0.41 0.49
    N315-SAS025 HP 0.39 0.47
    N315-SA2372 HP 0.38 0.28
    N315-SA0364 HP 0.36 0.41
    N315-SA1332 HP 0.34 0.26
    N315-SA1015 HP 0.34 0.23
    N315-SA2373 HP 0.32 0.20
    N315-SA0740 HP 0.29 0.24
    N315-SA0268 HP 0.17 0.17
    N315-SA0267 HP 0.17 0.19
    N315-SA1726 HP 0.08 0.17
    No N315 ORF
    COL-SA2069 HP 42.86 13.21
87.1% protein ID to Ssp extracellular ECM and plasma-binding protein 33.06 135.82
No hit in GenPept 16.42 28.08
mapN 99.7% protein ID to MapN protein 14.76 35.81
86.9% protein ID to SA1813 (possibly hemolysin) 6.80 12.73
46.1% protein ID to lin0925 putative membrane protein, Listeria innocua 5.73 23.61
50.7% protein ID to lin0924, Listeria innocua 5.52 14.61
63.2% protein ID to MW0768 4.58 2.71
    COL-SA1559 HP 3.68 2.70
91% protein ID to SA0093 HP 0.37 0.28
41.4% protein ID to HsdS probable restriction modification system 0.36 0.35
45.5% protein ID to SA2490 0.33 0.46
    COL-SA1043 Glycosyl transferase, group 1 0.33 0.15
100% ID to SA1814 0.31 0.25
97.9% protein ID to structure of cassette chromosome (SCC)-like element, stra 0.30 0.21
61.3% protein ID to SA0553 CHP 0.28 0.50
    COL-SA0653 CHP 0.23 0.46
24.1% protein ID to BdrC3, Borrelia hermsii 0.23 0.34
26.5% protein ID to RSc1168 CHP, Ralstonia solanacearum 0.22 0.43
29.4% protein ID to PF1843 chromosome segregation protein Smc, Pyrococcus furiosus 0.22 0.26
    COL-SA0654 CHP 0.21 0.48
50% protein ID to NMB0372 HP, Neisseria meningitidis 0.19 0.48
82.5% protein ID to SA0276 0.17 0.26
32.9% protein ID to ParA, B. subtilis 0.14 0.30
    COL-SA0095 spa Immunoglobulin G-binding protein A precursor 0.02 0.08
a

Based on the published sequence of S. aureus strain N315. For genes not present in N315, the gene name and description given are from the S. aureus strain COL genome, available from The Institute for Genomic Research website (www.tigr.org or by the putative function.

b

Abbreviations: PTS, phosphotransferase; CHP, conserved hypothetical protein; CoA, coenzyme A; NTP, nucleoside triphosphate; HP, hypothetical protein; ID, identity; ECM, extracellular matrix.

c

Normalized values based on the expression ratio (ER), which is defined as the expression level in biofilms/expression level in exponential-phase (EP) or stationary-phase (SP) cells.

We also identified 84 genes whose expression was reduced by a least a factor of at least 2 by comparison with both planktonic growth conditions (Table 5). Included were 25 genes in eight possible operons including an oligopeptide transport system (opp; N315-SA0845-SA0848) and the genes responsible for purine biosynthesis (pur; N315-SA0920-SA0925). Most genes in the other six putative operons encode hypothetical or conserved hypothetical proteins with no known function. However, one well-defined gene that was drastically downregulated in biofilms (60 to 139 times higher in the exponential-phase cultures and 12 to 27 times higher in the stationary-phase cultures) was spa, the gene that encodes protein A.

Confirmation of transcriptional profiling by real-time PCR.

To verify the results of our microarray experiments, we used real-time PCR to examine the relative expression levels of selective target genes. These comparisons were done using RNA isolated from two independent cultures representing each of three growth conditions (biofilm and exponential and stationary growth phases). As observed in our profiling experiments, the arcA, pyrR, and ureA transcripts were present in greater quantities in the biofilm samples than in both exponential- and stationary-phase planktonic cultures (Fig. 5). Indeed, while the patterns of gene expression observed with real-time PCR were consistent with our profiling experiments, the results from the real-time PCR experiments suggest that our profiling experiments may underestimate the actual differences. As with our profiling experiments, we also found that spa was significantly downregulated in biofilms on the basis of real-time PCR comparisons (Fig. 5). Collectively, the real-time PCR results provide independent verification of our DNA microarray results.

FIG. 5.

FIG. 5.

Relative expression levels as determined by real-time PCR. Expression levels of the arcA, pyrR, ureA, and spa genes was determined by real-time PCR. Relative expression levels are illustrated as the ratio of the expression level observed in biofilms (B) versus exponential-phase (E) or stationary-phase (S) planktonic cultures. The numbers in parentheses above the bars indicate the relative expression levels determined by transcriptional profiling.

Roles of genes regulated by sarA in biofilm formation.

For the reasons discussed above, we are particularly interested in genes that are differentially expressed in biofilms and are part of the sarA regulon. Therefore, we compiled a list of genes that were reported by Dunman et al. (19) to be regulated by sarA and were either induced or repressed in a biofilm compared to either planktonic condition. This analysis revealed 27 genes that were part of the sarA regulon and were differentially expressed in biofilms (Table 6). Because these genes may be genes that are required for biofilm formation, genes that are induced in biofilms and positively regulated by sarA would be of particular interest; however, we identified only four genes (sdhB, carA, an unidentified ORF with similarity to a major histocompatibility complex [MHC] class II analog, and a hypothetical protein) that fell into this category. At the same time, it may be equally important that specific genes be turned off to facilitate biofilm formation, and we identified eight genes (arc, phoP, pbp3, nuc, ndhG, spa, and two hypothetical proteins) that were repressed in biofilms and negatively regulated by sarA. The remaining genes were divergently regulated by sarA and in biofilms; however, the possibility that the impact of sarA is indirect in these cases cannot be ruled out.

TABLE 6.

SarA-regulated genes differentially expressed in a biofilm

N315 ORFa Common namea Producta,b B/Ec B/Sc SarAd
N315-SAS016 HP 12.74 Up
N315-SA2425 arcC Carbamate kinase 37.87 5.76 Down
N315-SA2424 arcR Transcription regulator Crp/Fnr family protein 48.16 9.87 Down
N315-SA2321 HP 0.33 Up
N315-SA2319 sdhB Similar to beta: subunit of l-serine dehydratase 8.90 4.84 Up
N315-SA2125 hutG Similar to formiminoglutamase 0.35 Up
N315-SA2006 Similar to MHC class II analog 4.69 Up
N315-SA1928 HP 0.42 Up
N315-SA1611 CHP 0.23 Up
N315-SA1553 fhs Formyltetrahydrofolate synthetase 0.17 Down
N315-SA1516 phoP Alkaline phosphatase synthesis 0.48 Up
N315-SA1381 pbp3 Penicillin-binding protein 3 0.48 Down
N315-SA1319 HP 0.42 Up
N315-SA1140 glpF Glycerol uptake facilitator 4.40 Down
N315-SA1120 Similar to transcription regulator GntR family 0.35 Up
N315-SA1045 carA Carbamoyl-phosphate synthase small chain 13.13 6.03 Up
N315-SA0923 purM Phosphoribosylformylglycinamidine cyclo-ligase 0.34 0.03 Down
N315-SA0900 sspB Cysteine protease precursor 7.03 Down
N315-SA0899 sspC Cysteine protease 7.24 Down
N315-SA0754 Similar to lactococcal prophage ps3 protein 05 0.50 Up
N315-SA0746 nuc Staphylococcal nuclease 0.17 Down
N315-SA0412 CHP 0.08 Down
N315-SA0411 ndhF NADH dehydrogenase subunit 5 0.07 2.41 Down
N315-SA0363 HP 0.32 Down
N315-SA0107 spa Immunoglobulin G-binding protein A precursor 0.01 0.04 Down
N315-SA0016 purA Adenylosuccinate synthase 2.37 0.12 Up
fnbB Fibronectin-binding protein homolog 0.46 Up
a

Based on the published sequence of S. aureus strain N315. For genes not present in N315, the gene name and description given are from the S. aureus strain COL genome, available from The Institute for Genomic Research website (www.tigr.org or by the putative function.

b

Abbreviations: HP, hypothetical protein; CHP, conserved hypothetical protein.

c

Normalized values based on expression levels in biofilms versus exponential-phase cells (B/E) or biofilms versus stationary-phase cells (B/S).

d

Regulation by SarA as reported by Dunman et al. (19).

DISCUSSION

Valle et al. (66) was the first to demonstrate that mutation of sarA results in a reduced capacity to form a biofilm. They concluded that this was due, at least in part, to reduced expression of the icaADBC operon. At the same time, Valle et al. (66) also suggested that SarA enhances biofilm formation by suppressing production of a second, unidentified protein that was either a repressor of PNAG synthesis or was involved in the turnover of PNAG. On this basis, they proposed a model in which the impact of sarA on biofilm formation was dependent on two pathways, both of which functioned by moderating the production of PNAG.

In this report, we also demonstrated that mutation of sarA results in reduced production of PNAG. To further investigate the impact of this on biofilm formation in our clinical isolate, we generated an ica mutant and examined its capacity to form a biofilm both in vitro and in vivo. Surprisingly, our ica mutant formed a biofilm comparable to that of the parental strain under both static and flow conditions. As in our previous experiments (5), the capacity of our UAMS-1 sarA mutant to form a biofilm was reduced in comparison to both the wild-type strain and its corresponding ica mutant. This clearly indicates that the impact of sarA on biofilm formation, at least as defined under in vitro growth conditions, involves a pathway that is independent of the icaADBC operon.

The role of ica in vivo has been addressed in S. epidermidis with contradictory results (10, 23, 41, 47, 49, 55, 56, 57, 62). However, few studies have addressed this issue in S. aureus. Cramton et al. (14) demonstrated that ica is present in S. aureus and that it is required for PIA production and biofilm formation in that species. However, the mutagenesis experiments were limited to a single strain of S. aureus (SA113) that was derived from NCTC 8325 by chemical mutagenesis, so these results may not be representative of the situation observed in clinical isolates. More recently, Vandecasteele et al. (67) analyzed expression of biofilm-associated genes, including icaA and icaC, both in vitro and in vivo. While expression of both genes was induced upon initial exposure to foreign bodies, this induction peaked shortly after the introduction of bacteria and was followed by a slow decrease over time. These results suggest that the ica operon is mainly associated with the initial colonization phase of biofilm formation, rather than maturation and persistence. This is consistent with our results and the fact that our analysis was limited to mature biofilms. Additionally, Francois et al. (23) compared an S. aureus strain and its corresponding ica mutant using a tissue cage model and found that the ica mutant retained the capacity to colonize at a level comparable to that of the wild-type strain. They concluded on this basis that biofilms were not an important factor in their model. However, we believe there are two possible alternative explanations. The first is that their model does not accurately reflect the need to form a biofilm. The second is that their ica mutant retained the capacity to form a biofilm under in vivo conditions. Our results would support the latter hypothesis. Specifically, we found that while the UAMS-1 sarA mutant demonstrated a reduced capacity to colonize catheters in vivo, the UAMS-1 ica mutant colonized the same substrates as well as the wild-type strain did. These results also suggest that PNAG is not required for in vivo colonization by S. aureus and that sarA regulates genes required for biofilm formation independent of its ability to modulate ica expression and/or the production of PNAG. We would note, however, that our results are also limited to a single strain, and it is certainly possible that UAMS-1 has an alternative means of promoting intercellular accumulation that attenuates the need for PIA. Whether this is true of other clinical isolates of S. aureus remains to be determined, but previous work in our laboratory has confirmed that all other sarA-mediated phenotypes are conserved among clinical isolates like UAMS-1 (5, 7).

Dunman et al. (19) recently reported the results of microarray-based transcriptional profiling experiments with S. aureus sarA and agr mutants. These studies confirmed that sarA has global regulatory effects that are mediated through both agr-dependent and agr-independent pathways. In an effort to determine which of these might be involved in biofilm formation and/or maintenance of the sessile lifestyle, we performed comprehensive transcriptional profiling with RNA isolated from mature S. aureus biofilms. Although it would be preferable to do these experiments using RNA derived from biofilms grown in vivo, our attempts to isolate a sufficient quantity of high-quality RNA from in vivo samples have thus far been unsuccessful. However, our studies indicating that mutation of sarA results in a reduced capacity to form a biofilm both in vitro and in vivo suggest that biofilms grown in flow cells may also provide a relevant source of RNA for transcriptional profiling. On this basis, we isolated RNA from UAMS-1 grown in flow cells and compared the pattern of gene expression to RNA from the same strain grown in planktonic culture. The results of our comparisons to both exponential- and stationary-phase planktonic cultures clearly indicate that our flow cell biofilm model represents a unique growth environment. Indeed, we identified a total of 580 genes that are differentially expressed in biofilms by comparison to either or both planktonic conditions.

Several of the operons that were induced in biofilms have been found to be important in acid tolerance in other bacterial species. Indeed, maintenance of pH homeostasis within the bacterial cell and buffering of the surrounding microenvironment have been associated with biofilm formation in the oral bacteria Streptococcus salivarius (34). One way bacteria combat acidic environments is to produce alkaline compounds, such as ammonia, that can neutralize the acids. Two ways in which bacteria generate ammonia are through the urease and arginine deiminase (ADI) pathways. Interestingly, we found that multiple genes from both of these pathways were induced in S. aureus biofilms by comparison to both planktonic conditions.

Under anaerobic conditions, some bacteria are also able to generate ATP as an energy source through catabolism of arginine via the ADI pathway, which is widely distributed in bacteria (16), archaea (54), and eucarya (60). The ADI pathway is comprised of three enzymatic reactions, catalyzed by arginine deiminase (arcA), ornithine transcarbamoylase (arcB), and carbamate kinase (arcC). Collectively, these enzymes convert arginine to ornithine, ammonia, and CO2, yielding 1 mol of ATP per mol of arginine consumed. ArcD is an arginine-ornithine transporter that catalyzes the uptake of arginine and concomitant export of ornithine, while arcR encodes an activator of the ADI operon that is a member of the Crp-Fnr family of regulators. We found that all five of these genes were significantly induced in S. aureus biofilms (Table 5). One additional gene that was upregulated 3.9-fold in a biofilm compared to exponential-phase cells is the arginine repressor encoded by argR (Table 3). Under anaerobic conditions in the presence of arginine, ArgR represses anabolic ornithine carbamoyltransferase and induces the ADI pathway. In addition to its role in generating ATP anaerobically, the ADI pathway is one of two major ammonia-generating pathways utilized by oral bacteria to maintain pH homeostasis when growing in a biofilm. Ammonia generated by the deimination of arginine can neutralize acids generated by bacterial glycolysis.

The ADI system and its role in acid resistance have also been correlated with virulence in Streptococcus pyogenes. ADI in this species was originally called streptococcal acid glycoprotein (SAGP) and was characterized as being an inhibitor of stimulated human peripheral blood mononuclear cell proliferation (17, 18). In addition, it is thought that the acid sensitivity of a SAGP-negative mutant is responsible for its reduced ability to enter and survive in epithelial cells (18).

Also included among the genes induced in biofilms (Table 5) were seven genes that comprise the urease operon (ureABCEFGD). Urease (urea amidohydrolase) is a nickel-containing enzyme that catalyzes the hydrolysis of urea to yield two molecules of ammonia and one molecule of CO2. Ureases of most bacteria are composed of three distinct subunits encoded by three contiguous genes, ureA, ureB, and ureC. Urease gene clusters also encode accessory genes, in addition to these structural genes, that are required for the de novo synthesis of active urease. Urease activity is essential for the colonization of the gastric mucosa by Helicobacter pylori and colonization of the urinary tract by both Proteus mirabilis and Staphylococcus saprophyticus (20, 25, 30). In addition, urease is thought to play a central role in the pathogenesis of Ureaplasma urealyticum urinary and respiratory tract infections (27, 35).

Recently, Saïd-Salim et al. (58) found that genes of the urease operon in S. aureus are negatively regulated by the SarA homolog Rot (repressor of toxin). Interestingly, expression of rot was repressed in biofilms, although this was limited to the comparison with exponential-phase planktonic cultures. The trigger for induced transcription of the urease operon in S. aureus has not yet been studied. However, urease synthesis by Klebsiella aerogenes is stimulated under conditions of nitrogen starvation, such as when the bacteria are cultured in minimal medium containing a poor nitrogen source, such as proline, arginine, or histidine (24). In H. pylori, expression of the urease operon is upregulated by a pH-dependent, posttranscriptional regulatory mechanism. More specifically, Akada et al. (1) showed that a shift to an acidic pH resulted in a significant increase in the level of ure operon mRNA, even in the presence of inhibitors of transcription. Until recently, it was believed that urease was associated with the cell surface and that it directly neutralized the microenvironment surrounding the cell (50). However, it is now thought to play a more important role as an intracellular enzyme required for acid resistance (61).

In addition to the production of ammonia, cation transport ATPases, such as the high-affinity K+-specific transport system encoded by the kdp operon, can also contribute to pH homeostasis through the exchange of K+ for H+ (13). In this study, we found that three genes of the kdp operon (kdpABC) were induced in biofilms by comparison to both planktonic growth conditions (Table 5) and two other genes (kdpDE) were induced by comparison to the exponential growth phase (Table 3). In E. coli, the Kdp system is composed of the ion motive P-type ATPase encoded by the kdpFABC operon, and expression of the kdpFABC operon is regulated by an adjacent operon, kdpDE (2). KdpD is a membrane-spanning sensor kinase, and KdpE is a cytosolic transcriptional activator. In response to an appropriate signal(s) (membrane stretch, alteration in turgor pressure, and external and internal potassium levels), KdpD transphosphorylates KdpE, which in turn upregulates transcription of the kdpFABC operon (2, 43). Interestingly, van der Laan et al. (68) recently reported that NH4+ ions strongly stimulate the ATPase activity of the KdpFABC complex in E. coli.

Several operons of the pyrimidine nucleotide biosynthetic (pyr) pathway (pyrRPBC, carAB, and pyrFE) were also induced in biofilms (Table 5). The pathway for the de novo synthesis of pyrimidines consists of six enzymatic steps leading to the formation of UMP. The first step in the pyrimidine biosynthetic pathway is the formation of carbomyl-phosphate (CP) from bicarbonate, glutamine (or ammonia), and ATP by CP synthase, which is encoded by the carAB genes. Interestingly, CP is also required for the biosynthesis of arginine. In B. subtilis, PyrR regulates transcription of the pyr operon by binding in a uridine nucleotide-dependent fashion to pyr mRNA and altering the secondary structure of the downstream mRNA (37, 38). Binding of PyrR to the downstream mRNA stabilizes a binding loop and prevents the formation of the antiterminator (39). In the absence of PyrR or when levels of the nucleotides UMP and UTP are low, the antiterminator is a stable secondary structure, and transcription of the downstream genes continues (39). While we do not know if the function of B. subtilis and S. aureus PyrR is conserved, our results showing that all of the genes in the pyr operon were induced in biofilms suggests that the level of UMP in cells growing in a biofilm is severely limited. In addition, upregulation of the pyr operon and subsequent CP production may be required for synthesis of sufficient levels of arginine to be used by the ADI pathway during anaerobic growth.

Taken together, the results of our array experiments suggest that mature biofilms are growing anaerobically and that genes of the acid tolerance response are upregulated in response to an acidic environment. While the regulatory pathway for the acid tolerance response in S. aureus has not been well characterized, there have been studies suggesting that the global regulators SigB and SarA are involved. Specifically, mutation of sigB results in strains with impaired abilities to respond to acid and to induce a stationary-phase acid tolerance response (8, 32). In addition, in the absence of RsbU, expression from the SigB-dependent sarA promoter was significantly reduced at pH 5.5 (46).

Our analysis revealed 27 genes that were differentially expressed in biofilms and were part of the sarA regulon as defined by Dunman et al. (19) (Table 6). Given the different mechanistic possibilities of how SarA modulates biofilm production, it is difficult to determine which of these genes might be the most relevant candidates for further consideration. For example, because SarA mutants have a reduced capacity to form a biofilm, SarA would presumably be required for production of an activator of biofilm formation or a required effector molecule, in which case transcription of the relevant gene(s) would be upregulated by SarA and in a biofilm. However, it is also possible that SarA represses production of an effector that is deleterious to biofilm formation, in which case the SarA target would be downregulated within a biofilm. One possibility in that regard is that the increased production of proteases observed in sarA mutants (7) results in degradation of a required surface protein. However, Valle et al. (66) concluded on the basis of both mutagenesis experiments and experiments employing specific inhibitors that protease production was not responsible for the decreased capacity of a sarA mutant to form a biofilm. Nevertheless, it remains possible that a similar scenario in which SarA represses transcription of some factor that has a negative impact on biofilm formation is involved. Moreover, both of these scenarios are based on the assumption that the SarA-dependent regulation of the relevant effector(s) is direct, and it is also possible that sarA modulates biofilm formation in an indirect fashion, perhaps via one of the increasing number of SarA homologs (4, 9, 45). It is also possible that sarA is required for production of an adhesin or some other effector protein that is required only for initial stages of biofilm formation. In this case, expression of the relevant gene may be transient in a fashion that would not have been apparent in our profiling experiments focusing on mature biofilms.

Finally, it should be noted that biofilms are not homogenous populations of cells and because the experiments we performed did not address this issue, it is certainly possible that we failed to detect genes that are within the sarA regulon but are differentially expressed only within certain regions of the biofilm. At the same time, this would imply that the genes we did identify are either differentially expressed throughout the biofilm or that our results actually underestimate the degree of differential gene expression observed within more limited regions. It should also be noted that the profiling experiments of Dunman et al. (19) were done using a first-generation GeneChip that was limited to only 85% of the genes identified in a single strain (COL) of S. aureus, and it is certainly possible that this limited the identification of relevant genes that are conserved among clinical isolates like UAMS-1. Moreover, the profiling experiments of Dunman et al. (19) were done using RNA isolated from derivatives of the S. aureus strain RN6390, and we have previously demonstrated that the regulatory circuits observed in RN6390 and its corresponding sarA and agr mutants are different from those of other strains, including UAMS-1 (7). To address these issues, we are currently performing transcriptional profiling experiments using RNA isolated from UAMS-1 sarA and agr mutants and the more comprehensive chips used for the biofilm profiling experiments described here.

REFERENCES

  • 1.Akada, J. K., M. Shirai, H. Takeuchi, M. Tsuda, and T. Nakazawa. 2000. Identification of the urease operon in Helicobacter pylori and its control by mRNA decay in response to pH. Mol. Microbiol. 36:1071-1084. [DOI] [PubMed] [Google Scholar]
  • 2.Altendorf, K., P. Voelkner, and W. Puppe. 1994. The sensor kinase KdpD and the response regulator KdpE control expression of the kdpFABC operon in Escherichia coli. Res. Microbiol. 145:374-381. [DOI] [PubMed] [Google Scholar]
  • 3.Arciola, C. R., L. Baldassarri, and L. Montanaro. 2001. Presence of icaA and icaD genes and slime production in a collection of staphylococcal strains from catheter-associated infections. J. Clin. Microbiol. 39:2151-2156. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 4.Arvidson, S., and K. Tegmark. 2001. Regulation of virulence determinants in Staphylococcus aureus. Int. J. Med. Microbiol. 291:159-170. [DOI] [PubMed] [Google Scholar]
  • 5.Beenken, K. E., J. S. Blevins, and M. S. Smeltzer. 2003. Mutation of sarA in Staphylococcus aureus limits biofilm formation. Infect. Immun. 71:4206-4211. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 6.Blevins, J. S., A. F. Gillaspy, T. M. Rechtin, B. K. Hurlburt, and M. S. Smeltzer. 1999. The staphylococcal accessory regulator (sar) represses transcription of the Staphylococcus aureus collagen adhesin gene (cna) in an agr-independent manner. Mol. Microbiol. 33:317-326. [DOI] [PubMed] [Google Scholar]
  • 7.Blevins, J. S., K. E. Beenken, M. O. Elasri, B. K. Hurlburt, and M. S. Smeltzer. 2002. Strain-dependent differences in the regulatory roles of sarA and agr in Staphylococcus aureus. Infect. Immun. 70:470-480. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 8.Chan, P. F., S. J. Foster, E. Ingram, and M. O. Clements. 1998. Staphylococcus aureus alternative sigma factor σB controls the environmental stress response but not starvation survival or pathogenicity in a mouse abscess model. J. Bacteriol. 180:6082-6089. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 9.Cheung, A. L., and G. Zhang. 2002. Global regulation of virulence determinants in Staphylococcus aureus by the SarA protein family. Front. Biosci. 1:1825-1842. [DOI] [PubMed] [Google Scholar]
  • 10.Christensen, G. D., L. Baldassarri, and W. A. Simpson. 1994. Colonization of medical devices by coagulase-negative staphylococci, p. 45-78. In A. L. Bisno and F. A. Waldvogel (ed.)., Infections associated with indwelling medical devices, 2nd ed. ASM Press, Washington, D.C.
  • 11.Clements, M. O., and S. J. Foster. 1999. Stress resistance in Staphylococcus aureus. Trends Microbiol. 7:458-462. [DOI] [PubMed] [Google Scholar]
  • 12.Conlon, K. M., H. Humphreys, and J. P. O'Gara. 2002. icaR encodes a transcriptional repressor involved in environmental regulation of ica operon expression and biofilm formation in Staphylococcus epidermidis. J. Bacteriol. 184:4400-4408. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 13.Cotter, P. D., and H. Colin. 2003. Surviving the acid test: responses of gram-positive bacteria to low pH. Microbiol. Mol. Biol. Rev. 67:429-453. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 14.Cramton, S. E., C. Gerke, N. F. Schnell, W. W. Nichols, and F. Götz. 1999. The intercellular adhesion (ica) locus is present in Staphylococcus aureus and is required for biofilm formation. Infect. Immun. 67:5427-5433. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 15.Cramton, S. E., M. Ulrich, F. Götz, and G. Döring. 2001. Anaerobic conditions induce expression of polysaccharide intercellular adhesin in Staphylococcus aureus and Staphylococcus epidermidis. Infect. Immun. 69:4079-4085. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 16.Cunin, R., N. Glansdorff, A. Pierard, and V. Stalon. 1986. Biosynthesis and metabolism of arginine in bacteria. Microbiol. Rev. 50:314-352. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 17.Degnan, B. A., M. A. Kehoe, and J. A. Goodacre. 1997. Analysis of human T cell responses to group A streptococci using fractionated Streptococcus pyogenes proteins. FEMS Immunol. Med. Microbiol. 17:161-170. [DOI] [PubMed] [Google Scholar]
  • 18.Degnan, B. A., M. C. Fontaine, A. H. Doebereiner, J. J. Lee, P. Mastroeni, G. Dougan, J. A. Goodacre, and M. A. Kehoe. 2000. Characterization of an isogenic mutant of Streptococcus pyogenes Manfredo lacking the ability to make streptococcal acid glycoprotein. Infect. Immun. 68:2441-2448. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 19.Dunman, P. M., E. Murphy, S. Haney, D. Palacios, G. Tucker-Kellogg, S. Wu, E. L. Brown, R. J. Zagursky, D. Shlaes, and S. J. Projan. 2001. Transcription profiling-based identification of Staphylococcus aureus genes regulated by the agr and/or sarA loci. J. Bacteriol. 183:7341-7353. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 20.Eaton, K. A., J. V. Gilbert, E. A. Joyce, A. E. Wanken, T. Thevenot, P. Baker, A. Plaut, and A. Wright. 2002. In vivo complementation of ureB restores the ability of Helicobacter pylori to colonize. Infect. Immun. 70:771-778. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 21.Evans, R. P., C. L. Nelson, W. R. Bowen, M. G. Kleve, and S. G. Hickmon. 1998. Visualization of bacterial glycocalyx with a scanning electron microscope. Clin. Orthop. Relat. Res. 347:243-249. [PubMed] [Google Scholar]
  • 22.Fowler, V. G., Jr., P. D. Fey, L. B. Reller, A. L. Chamis, G. R. Corey, and M. E. Rupp. 2001. The intercellular adhesin locus ica is present in clinical isolates of Staphylococcus aureus from bacteremic patients with infected and uninfected prosthetic joints. Med. Microbiol. Immunol. 189:127-131. [DOI] [PubMed] [Google Scholar]
  • 23.Francois, P., P. H. Tu Quoc, C. Bisognano, W. L. Kelly, D. P. Lew, J. Schrenzel, S. E. Cramton, F. Götz, and P. Vaudaux. 2003. Lack of biofilm contribution to bacterial colonisation in an experimental model of foreign body infection by Staphylococcus aureus and Staphylococcus epidermidis. FEMS Immunol. Med. Microbiol. 35:135-140. [DOI] [PubMed] [Google Scholar]
  • 24.Friedrich, B., and B. Magasanik. 1977. Urease of Klebsiella aerogenes: control of its synthesis by glutamine synthetase. J. Bacteriol. 131:446-452. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 25.Gatermann, S., and R. Marre. 1989. Cloning and expression of Staphylococcus saprophyticus urease gene sequences in Staphylococcus carnosus and contribution of the enzyme to virulence. Infect. Immun. 57:2998-3002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 26.Götz, F. 2002. Staphylococcus and biofilms. Mol. Microbiol. 43:1367-1378. [DOI] [PubMed] [Google Scholar]
  • 27.Hedelin, H., J. E. Brorson, L. Grenabo, and S. Pettersson. 1984. Ureplasma urealyticum and upper urinary tract stones. Br. J. Urol. 56:244-249. [DOI] [PubMed] [Google Scholar]
  • 28.Heilmann, C., O. Schweitzer, C. Gerke, N. Vanittanakom, D. Mack, and F. Götz. 1996. Molecular basis of intercellular adhesion in the biofilm-forming Staphylococcus epidermidis. Mol. Microbiol. 20:1083-1091. [DOI] [PubMed] [Google Scholar]
  • 29.Heinrichs, J. H., M. G. Bayer, and A. L. Cheung. 1996. Characterization of the sar locus and its interaction with agr in Staphylococcus aureus. J. Bacteriol. 178:418-423. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 30.Jones, B. D., C. V. Lockatell, D. E. Johnson, J. W. Warren, and H. L. Mobley. 1990. Construction of urease-negative mutant of Proteus mirabilis: analysis of virulence in a mouse model of ascending urinary tract infection. Infect. Immun. 58:1120-1123. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 31.Kadurugamuwa, J. L., L. Sin, E. Albert, J. Yu, K. Francis, M. DeBoer, M. Rubin, C. Bellinger-Kawahara, T. R. Parr, Jr., and P. R. Contag. 2003. Direct continuous method for monitoring biofilm infection in a mouse model. Infect. Immun. 71:882-890. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 32.Kullik, I., P. Giachino, and T. Fuchs. 1998. Deletion of the alternative sigma factor σB in Staphylococcus aureus reveals its function as a global regulator of virulence genes. J. Bacteriol. 180:4814-4820. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 33.Lewis, K. 2001. Riddle of biofilm resistance. Antimicrob. Agents Chemother. 45:999-1007. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 34.Li, Y. H., Y. Y. Chen, and R. A. Burne. 2000. Regulation of urease gene expression by Streptococcus salivarius growing in biofilms. Environ. Microbiol. 2:169-177. [DOI] [PubMed] [Google Scholar]
  • 35.Ligon, J. V., and G. E. Kenny. 1991. Virulence of ureaplasmal urease for mice. Infect. Immun. 59:1170-1171. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 36.Lowen, P., B. Hu, J. Strutinsky, and R. Sparling. 1998. Regulation in the rpoS regulon of Escherichia coli. Can. J. Microbiol. 44:707-717. [DOI] [PubMed] [Google Scholar]
  • 37.Lu, Y., and R. L. Switzer. 1996. Transcriptional attenuation of the Bacillus subtilis pyr operon by the PyrR regulatory protein and uridine nucleotides in vitro. J. Bacteriol. 178:7206-7211. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 38.Lu, Y., R. J. Turner, and R. L. Switzer. 1995. Roles of the three transcriptional attenuators of the Bacillus subtilis pyrimidine biosynthetic operon in the regulation of its expression. J. Bacteriol. 177:1315-1325. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 39.Lu, Y., R. J. Turner, and R. L. Switzer. 1996. Function of RNA secondary structures in transcriptional attenuation of the Bacillus subtilis pyr operon. Proc. Natl. Acad. Sci. USA 93:14462-14467. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 40.Maira-Litran, T., A. Kropec, C. Abeygunawardana, J. Joyce, G. Mark III, D. A. Goldmann, and G. B. Pier. 2002. Immunochemical properties of the staphylococcal poly-N-acetylglucosamine surface polysaccharide. Infect. Immun. 70:4433-4440. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 41.McDermid, K. P., D. W. Morck, M. E. Olson, N. D. Boyd, A. E. Khoury, M. K. Dasgupta, and J. W. Costerton. 1993. A porcine model of Staphylococcus epidermidis catheter-associated infection. J. Infect. Dis. 168:897-903. [DOI] [PubMed] [Google Scholar]
  • 42.McKenney, D., K. L. Pouliot, Y. Wang, V. Murthy, M. Ulrich, G. Doring, J. C. Lee, D. A. Goldmann, and G. B. Pier. 1999. Broadly protective vaccine for Staphylococcus aureus based on an in vivo-expressed antigen. Science 284:1523-1527. [DOI] [PubMed] [Google Scholar]
  • 43.Nakashima, K., A. Sugiura, K. Kanamaru, and T. Mizuno. 1993. Signal transduction between the two regulatory components involved in the regulation of the kdpABC operon in Escherichia coli: phosphorylation-dependent functioning of the positive regulator, KdpE. Mol. Microbiol. 7:109-116. [DOI] [PubMed] [Google Scholar]
  • 44.Ni Edhin, D., S. Perkins, P. Francois, P. Vaudaux, M. Hook, and T. J. Foster. 1998. Clumping factor B (ClfB), a new surface-located fibrinogen-binding adhesin of Staphylococcus aureus. Mol. Microbiol. 30:245-257. [DOI] [PubMed] [Google Scholar]
  • 45.Novick, R. P. 2003. Autoinduction and signal transduction in the regulation of staphylococcal virulence. Mol. Microbiol. 48:1429-1449. [DOI] [PubMed] [Google Scholar]
  • 46.Palma, M., and A. L. Cheung. σB activity in Staphylococcus aureus is controlled by RsbU and an additional factor(s) during bacterial growth. Infect. Immun. 69:7858-7865. [DOI] [PMC free article] [PubMed]
  • 47.Patrick, C. C., S. V. Hetherington, P. K. Roberson, S. Henwick, and M. M. Sloas. 1995. Comparative virulence of Staphylococcus epidermidis isolates in a murine catheter model. Ped. Res. 37:70-74. [DOI] [PubMed] [Google Scholar]
  • 48.Patti, J. M., B. L. Allen, M. J. McGavin, and M. Hook. 1994. MSCRAMM-mediated adherence of microorganisms to host tissues. Annu. Rev. Microbiol. 48:585-617. [DOI] [PubMed] [Google Scholar]
  • 49.Perdreau-Remington, F., M. A. Sande, G. Peters, and H. F. Chambers. 1998. The abilities of Staphylococcus epidermidis wild-type strain and its slime-negative mutant to induce endocarditis in rabbits are comparable. Infect. Immun. 66:2778-2781. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 50.Phadnis, S. H., M. H. Parlow, M. Levy, D. Ilver, C. M. Caulkins, J. B. Connors, and B. E. Dunn. 1996. Surface localization of Helicobacter pylori urease and heat shock protein homolog requires bacterial autolysis. Infect. Immun. 64:905-912. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 51.Prigent-Combaret, C., O. Vidal, C. Dorel, and P. Lejeune. 1999. Abiotic surface sensing and biofilm-dependent regulation of gene expression in Escherichia coli. J. Bacteriol. 181:5993-6002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 52.Rachid, S., K. Ohlsen, U. Wallner, J. Hacker, M. Hecker, and W. Ziebuhr. 2000. Alternative transcription factor σB is involved in regulation of biofilm expression in a Staphylococcus aureus mucosal isolate. J. Bacteriol. 182:6824-6826. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 53.Rohde, H., J. K. Knobloch, M. A. Horstkotte, and D. Mack. 2001. Correlation of biofilm expression types of Staphylococcus epidermidis with polysaccharide intercellular adhesin synthesis: evidence for involvement of icaADBC genotype-independent factors. Med. Microbiol. Immun. 190:105-112. [DOI] [PubMed] [Google Scholar]
  • 54.Ruepp, A., and J. Soppa. 1996. Fermentative arginine degradation in Halobacterium salinarium (formerly Halobacterium halobium): genes, gene products, and transcripts of the arcRACB gene cluster. J. Bacteriol. 178:4942-4947. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 55.Rupp, M. E., J. S. Ulphani, P. D. Fey, and D. Mack. 1999. Characterization of Staphylococcus epidermidis polysaccharide intercellular adhesin/hemagglutinin in the pathogenesis of intravascular catheter-associated infection in a rat model. Infect. Immun. 67:2656-2659. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 56.Rupp, M. E., J. S. Ulphani, P. D. Fey, K. Bartscht, and D. Mack. 1999. Characterization of the importance of polysaccharide intercellular adhesin/hemagglutinin of Staphylococcus epidermidis in the pathogenesis of biomaterial-based infection in a mouse foreign body infection model. Infect. Immun. 67:2627-2632. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 57.Rupp, M. E., P. D. Fey, C. Heilmann, and F. Götz. Characterization of the importance of Staphylococcus epidermidis autolysin and polysaccharide intercellular adhesin in the pathogenesis of intravascular catheter-associated infection in a rat model. J. Infect. Dis. 183:1038-1042. [DOI] [PubMed]
  • 58.Saïd-Salim, B., P. M. Dunman, F. M. McAleese, D. Macapagal, E. Murphy, P. J. McNamara, S. Arvidson, T. J. Foster, S. J. Projan, and B. N. Kreiswirth. 2003. Global regulation of Staphylococcus aureus genes by Rot. J. Bacteriol. 185:610-619. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 59.Schembri, M. A., K. Kjærgaard, and P. Klemm. 2003. Global gene expression in Escherichia coli biofilms. Mol. Microbiol. 48:253-267. [DOI] [PubMed] [Google Scholar]
  • 60.Schofield, P. J., M. R. Edwards, J. Matthews, and J. R. Wilson. 1992. The pathway of arginine catabolism in Giardia intestinalis. Mol. Biochem. Parasitol. 51:29-36. [DOI] [PubMed] [Google Scholar]
  • 61.Scott, D. R., D. Weeks, C. Hong, S. Postius, K. Melchers, and G. Sachs. 1998. The role of internal urease in acid resistance of Helicobacter pylori. Gastroenterology 114:58-70. [DOI] [PubMed] [Google Scholar]
  • 62.Shiro, H., E. Muller, N. Gutierrez, S. Boisot, M. Grout, T. D. Tosteson, D. Goldmann, and G. B. Pier. 1994. Transposon mutants of Staphylococcus epidermidis deficient in elaboration of capsular polysaccharide/adhesin and slime are virulent in a rabbit model of endocarditis. J. Infect. Dis. 169:1042-1049. [DOI] [PubMed] [Google Scholar]
  • 63.Stanley, N. R., R. A. Britton, A. D. Grossman, and B. A. Lazazzera. 2003. Identification of catabolite repression as a physiological regulator of biofilm formation by Bacillus subtilis by use of DNA microarrays. J. Bacteriol. 185:1951-1957. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 64.Stewart, P. S. 2002. Mechanisms of antibiotic resistance in bacterial biofilms. Int. J. Med. Microbiol. 292:107-113. [DOI] [PubMed] [Google Scholar]
  • 65.Stewart, P. S., and J. W. Costerton. 2001. Antibiotic resistance of bacteria in biofilms. Lancet 358:135-138. [DOI] [PubMed] [Google Scholar]
  • 66.Valle, J., A. Toledo-Arana, C. Berasain, J. Ghigo, B. Amorena, J. R. Penades, and I. Lasa. 2003. SarA and not σB is essential for biofilm development by Staphylococcus aureus. Mol. Microbiol. 48:1075-1087. [DOI] [PubMed] [Google Scholar]
  • 67.Vandecasteele, S. J., W. E. Peetermans, R. Merckx, B. J. Rijnders, and J. Van Eldere. 2003. Reliability of the ica, aap, and atlE genes in the discrimination between invasive, colonizing and contaminant Staphylococcus epidermidis isolates in the diagnosis of catheter-related infections. Clin. Microbiol. Infect. 9:114-119. [DOI] [PubMed] [Google Scholar]
  • 68.Van der Laan, M., M. Gassel, and K. Altendorf. 2002. Characterization of amino acid substitutions in KdpA, the K+-binding and -translocating subunit of the KdpFABC complex to Escherichia coli. J. Bacteriol. 184:5491-5494. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 69.Vuong, C., C. Gerke, G. A. Somerville, E. R. Fischer, and M. Otto. 2003. Quorum-sensing control of biofilm factors in Staphylococcus epidermidis. J. Infect. Dis. 188:706-718. [DOI] [PubMed] [Google Scholar]
  • 70.Vuong, C., H. L. Saenz, F. Götz, and M. Otto. 2000. Impact of the agr quorum-sensing system on adherence to polystyrene in Staphylococcus aureus. J. Infect. Dis. 182:1688-1693. [DOI] [PubMed] [Google Scholar]
  • 71.Whiteley, M., M. G. Bangera, R. E. Bumgarner, M. R. Parsek, G. M. Teitzel, S. Lory, and E. P. Greenberg. 2001. Gene expression in Pseudomonas aeruginosa biofilms. Nature 413:860-864. [DOI] [PubMed] [Google Scholar]
  • 72.Wolz, C., D. McDevitt, T. J. Foster, and A. L. Cheung. 1996. Influence of agr on fibrinogen binding in Staphylococcus aureus Newman. Infect. Immun. 64:3142-3147. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 73.Wolz, C., C. Goerke, R. Landmann, W. Zimmerli, and U. Fluckiger. 2002. Transcription of clumping factor A in attached and unattached Staphylococcus aureus in vitro and during device-related infection. Infect. Immun. 70:2758-2762. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 74.Wolz, C., P. Pöhlmann-Dietze, A. Steinhuber, Y. T. Chien, A. Manna, W. van Wamel, and A. Cheung. 2000. Agr-independent regulation of fibronectin-binding proteins by the regulatory locus sar in Staphylococcus aureus. Mol. Microbiol. 36:230-243. [DOI] [PubMed] [Google Scholar]
  • 75.Ziebuhr, W., V. Krimmer, S. Rachid, I. Lössner, F. Götz, and J. Hacker. 1999. A novel mechanism of phase variation of virulence in Staphylococcus epidermidis: evidence for control of the polysaccharide intercellular adhesin synthesis by alternating insertion and excision of the insertion sequence element IS256. Mol. Microbiol. 32:345-356. [DOI] [PubMed] [Google Scholar]

Articles from Journal of Bacteriology are provided here courtesy of American Society for Microbiology (ASM)

RESOURCES