Skip to main content
Journal of Bacteriology logoLink to Journal of Bacteriology
. 2006 Jun;188(11):3731–3739. doi: 10.1128/JB.01780-05

A Complex Transcription Network Controls the Early Stages of Biofilm Development by Escherichia coli

Birgit M Prüß 1,*, Christopher Besemann 2, Anne Denton 2, Alan J Wolfe 3
PMCID: PMC1482888  PMID: 16707665

Historically, researchers have studied bacterial signaling as if it functioned as a set of isolated, linear pathways. More recent studies, however, have demonstrated that many signaling pathways interact and that these interacting pathways should be construed as an intricate network. This network integrates diverse signals, both extracellular and intracellular, to ensure that the the correct amount of the appropriate subset of genes is expressed at the proper time. Complete delineation of this complex signal transduction network and use of the network to predict the full range of cellular behaviors are major goals of systems biology.

Despite considerable progress, we remain near the beginning of this process, which thus far has been dominated by the development of enabling technologies and the compilation of gene lists. Although development and compilation will continue to be essential, the next critical step must be to organize the copious data compiled over 5 decades of pregenomics research and the massive amount of postgenomics data generated over the last decade. This minireview, in which we describe a portion of the overall network of Escherichia coli, is an attempt to perform part of this next step.

THE NETWORK

As the model organism for this network, we chose the enterobacterium E. coli. We focused specifically on the common laboratory strain K-12 in order to mine the wealth of information available for it. When appropriate, we included observations made with other E. coli variants (e.g., enterohemorragic E. coli [EHEC] or uropathogenic E. coli) or with the close relative Salmonella enterica. With easy to moderate effort, the network can be adapted to other enterobacterial relatives. However, more distantly related species may lack some of the global regulators discussed here.

As a unifying theme, we chose the early stages of biofilm development. Defined as a sessile community of bacteria encased in a matrix, a biofilm tends to develop on a surface or an interface in a series of ordered steps, designated reversible attachment, irreversible attachment, maturation-1, maturation-2, and dispersion (121). Each step requires reprogramming of gene expression that occurs in response to the changing environment (122). The reprogramming associated with the earliest steps of biofilm development can be identified easily by the distinct organelles that decorate the bacterial surface. For example, reversible attachment often involves flagella that permit individual planktonic cells to swim toward an appropriate biotic or abiotic surface. Irreversible attachment involves the loss of these flagella and the elaboration of adhesive organelles (e.g., curli or type 1 fimbriae); the type of organelle depends on the environment. Finally, production of the colanic acid capsule permits construction of the distinctive three-dimensional structure typical of mature biofilms (for a recent review of biofilm formation, see reference 149).

For the surface organelles to appear in proper order, expression of these organelles must be coordinately regulated (137). Indeed, there is evidence for regulatory relationships between flagella and fimbriae (10, 75), between flagella and capsule (80, 124, 158), and between different types of fimbriae (52, 159). The coordinate regulation of these surface organelles, whose expression responds to similar subsets of external signals, second messengers, and regulators, is the main focus of this minireview.

The total network consists of 16 regulators and the several hundred genes that they regulate. Some regulators in this network function globally. For example, CRP (162) and H-NS (13, 53) each regulate hundreds of genes. In contrast, some regulators, including LrhA (73) and HdfR (67), affect transcription of only a small number of genes. Some global regulators are members of a family of two-component signal transduction (2CST) pathways, a predominant system used by bacteria to relay environmental signals in order to elicit changes in cellular functions (for reviews of 2CST pathways, see references 36, 61, 100, and 156). Each 2CST pathway consists of a sensor and a response regulator. The sensor, often an integral cytoplasmic membrane protein, is a histidine kinase that uses ATP as its phosphodonor to autophosphorylate a conserved histidine residue (H1). Some sensors also possess phosphatase activity. In contrast, the response regulator is an aspartyl kinase that uses the phosphorylated sensor as its phosphodonor to autophosphorylate a conserved aspartyl residue (D1). Most, but not all, response regulatory domains are fused to a DNA binding domain and thus function as transcription factors. E. coli possesses 29 histidine kinases and 32 response regulators (80, 86), including EnvZ/OmpR, QseC/QseB (QseCB), and CpxA/CpxR (CpxAR).

Deletions of the genes that encode all these 2CST pathways have been constructed and analyzed by microarray technology. This has been done for both gene expression analysis (97) and phenotypic characterization (163). For example, one analysis showed that the EnvZ/OmpR pathway, initially discovered as a regulator of the outer membrane porins OmpC (79) and OmpF (148), actually regulates a much larger set of genes (97). Another analysis showed that QseCB, identified as a regulator of quorum sensing in EHEC (134), also regulates flagellum biogenesis (27, 135). Other studies showed that the CpxAR system, which was first shown to sense envelope insult (31), also regulates the DNA repair gene ung (93), genes encoding the type IV bundle-forming pili (91), and both curli operons (59).

A more complex variant of the 2CST pathway is the multistep phosphorelay, which includes four domains instead of two domains (6, 86). Like the conventional 2CST pathway, the multistep phosphorelay proceeds from the sensor to a response regulator. A histidine phosphotransferase then transfers the phosphate from this first response regulator to a second response regulator. The second response regulator is often a transcription factor. RcsC/RcsD/RcsB (RcsCDB) is one of the few phosphorelays possessed by E. coli (6, 86). One of the four domains that comprise the complete phosphorelay, RcsC, contains both the sensor and the first response regulator. RcsD includes the histidine phosphotransferase, and RcsB carries the final response regulator. The signal travels from H1 to D1 on RcsC, then to H2 on RcsD, and finally to D2 on RcsB (for a review of RcsCDB signaling, see reference 80). Originally identified as a regulator of the capsule synthesis genes (cps) (43), RcsCDB is now known to regulate up to 5% of the E. coli genome (32, 44).

To construct the network, data were gleaned from the literature and/or from work performed in our laboratories. These data came primarily from functional genomics experiments, such as microarray analysis or analysis of genomic libraries of reporter gene fusions, but they also were obtained from direct interaction studies, such as electrophoretic mobility shift assays or DNase I footprint analyses.

To visualize the network, the open source TouchGraph visualization software was adapted as follows: functionality was added to distinguish multiple types of relationships and to systematically select a center of focus. TouchGraph uses the spring layout concept and allows user interactions through focus and context techniques (51). The network is presented in its entirety in the supplemental material (see Fig. S1 in the supplemental material). The remaining information in the supplemental material focuses on areas where there is intense regulation.

Here we summarize the subset of genes most closely aligned with biogenesis of the surface structures associated with the transition from motile, planktonic individuals to a sessile biofilm community (Fig. 1). First, we focus on three biofilm-associated surface organelles (flagella, curli, and type 1 fimbriae) whose biogenesis is controlled by the network. We then shift our attention to the network itself, concentrating on the three most prominent regulators, FlhD/FlhC (FlhDC), EnvZ/OmpR, and RcsCDB. A fourth biofilm-associated structure, the capsule, is mentioned in the context of RcsCDB. Finally, we discuss three small molecules. Two of these molecules, cyclic AMP (cAMP) and acetyl phosphate (acetyl∼P), affect the network directly. The third, cyclic di-GMP (c-di-GMP), is included because it influences the formation of biofilms (131) in parallel with the network, although how it performs this function remains unknown.

FIG. 1.

FIG. 1.

Global network of transcriptional regulation in E. coli. Positive regulatory effects are indicated by solid lines and arrowheads. Negative regulatory effects are indicated by dotted lines with blunt ends. Microarray data were obtained for EnvZ/OmpR (97), RcsCDB (32, 44, 97), LrhA (73), H-NS (53), CRP (162), CsgD (19), FlhD/FlhC (110, 111), FlhD (110, 112), and Aer (110). Further regulation of flhD expression has been documented, as follows: QseCB (134, 135), CRP (132), H-NS (14), DnaK, DnaJ, and GrpE (127), DnaA (88, 90), HdfR (67), and insertion element insertion (9). The expression of csgD and csgB is further regulated by EnvZ/OmpR (106), CpxAR (59, 106), and H-NS (59). FliA has been described as an alternative sigma factor specific for the flagellar genes (78) and mediates the regulation of aer expression by FlhD/FlhC (B. Prüß and P. Matsumura, unpublished).

THE REGULATED PROCESSES

The three major cellular processes regulated by the network are biogenesis of the surface organelles flagella, curli, and type I fimbriae.

Flagellum biogenesis.

Flagella enable bacteria to reach favorable environments, and they have functions in adhesion, biofilm formation, and colonization (47). The environmental conditions that control the levels of expression of flhDC include temperature (1), osmolarity (128), pH (133), the concentrations of catabolite-repressing carbon sources (161), and a number of small molecules (65, 71, 72, 74, 89, 109, 125, 126), including acetate and propionate (104). The mechanisms by which these factors regulate flhDC expression are largely unknown.

In E. coli K-12, the transcription initiation site for flhDC is located 198 bp upstream of the translation start site for flhD (132) (Fig. 2); in EHEC, it is located (27) only 53 bp upstream of the flhD open reading frame (132). In E. coli K-12, DNA binding sites have been identified for H-NS (132), phosphorylated OmpR (128), LrhA (73), RcsB (33), and CRP (132) (Fig. 2); in EHEC, additional DNA sites have been identified for phosphorylated QseB (27).

FIG. 2.

FIG. 2.

Regulation of flagella and motility in E. coli. Environmental control of the flagellar system is mediated by regulation of the flhDC promoter. The translational start site was determined by Soutourina et al. (132). Footprinting data have been obtained for H-NS (132), phosphorylated OmpR (128), LrhA (73), RcsAB (33), and cAMP-CRP (132). Insertion sites for the IS1 and IS5 elements are indicated (9). The figure is modified from a previous study (8); posttranscriptional control by CsrA (154) and posttranslational control by ClpXP (146, 147) have been added.

Additional experimental evidence indicates that there is transcriptional regulation of flhDC by the chaperones DnaK, DnaJ, and GrpE (127), the nucleoid protein DnaA (88, 90), and the transcription factor HdfR (67) (see Fig. S2 in the supplemental material). Furthermore, insertion of insertion elements increases transcription of flhDC, presumably by uncoupling upstream binding sites for negative regulators from the core promoter (9). flhDC also is regulated posttranscriptionally by the carbon storage regulator CsrA (Fig. 2), which binds to flhDC mRNA and increases transcript stability (154). A regulatory RNA, CsrB, sequesters and represses CsrA. A complex regulatory circuit involving UvrY (also known as SirA) regulates CsrB (142). Posttranslational regulation is mediated by protease ClpX/ClpP in S. enterica (146, 147).

Curli biogenesis.

Curli (also known as thin aggregative fimbriae) are adhesive fibers (115) that promote biofilm formation by facilitating initial cell-surface interactions and subsequent cell-cell interactions (96, 150). The environmental conditions that control curli expression include temperature, oxygen tension, starvation, osmolarity, iron, and pH (39, 106, 117, 141). Because they had not been observed under conditions that mimic the mammalian host environment (i.e., high osmolarity and high temperature) (45, 84, 95), for a long time curli were considered unable to contribute to human infections. A recent report, however, showed that cells can express curli under these conditions, if they are grown under static conditions that facilitate biofilm formation (63). At least eight regulators affect the expression of the curli genes (see Fig. S3 in the supplemental material), which cluster in two divergent operons, csgDEFG and csgBA (45). These regulators include three two-component systems, EnvZ/OmpR (106), RcsCDB (32), and CpxAR (59, 106), and four other regulators, CRP (162), H-NS (7, 53, 95), MlrA (20), and FlhDC (110). Most of these regulators act upon the csgD operon (40, 106), which encodes a transcriptional regulator of csgB. CsgD also regulates yaiC, yagS, pepD (19), and glyA (25). In addition to CsgD, csgB expression requires σS, an effect enhanced by the small protein Crl (18).

The best-investigated regulation of csgD expression involves the interplay between the negative regulator CpxAR and the positive regulator EnvZ/OmpR (59, 106). The phosphorylated forms of both CpxR and OmpR were shown to bind to overlapping DNA sites immediately upstream of the csgD promoter (106). CpxR bound cooperatively to six sites within the csgD promoter (59). Binding of CpxR and OmpR was not competitive, as both regulators could bind simultaneously. Considering that the expression and the phosphorylation state of CpxR both increased upon a shift to high osmolarity, it was postulated that induction of CpxAR mediates csgD repression at high osmolarity, whereas EnvZ/OmpR mediates csgD activation at low osmolarity (59).

Type I fimbria biogenesis.

Type I fimbriae mediate adherence to mannose-containing receptors and promote bacterial attachment to and/or invasion of host cells during urinary tract infections (29, 82). The structural genes (fimA to fimH) are located in a single large operon (81) that is driven by a single promoter located upstream of fimA (123). Expression data indicate that a strong terminator is located immediately after fimA (123). Expression of the fim operon is controlled primarily by an invertible 314-bp switch element that is located upstream of fimA and is flanked by inverted repeats. The inversion, called phase variation, is mediated by two recombinases, FimE and FimB (66). The genes encoding these recombinases are located upstream of the switch element and are transcribed in the same direction as the fim operon. Generally, FimB can promote inversion in both directions. FimE, in contrast, promotes only the switch from phase-ON to phase-OFF (34).

Phase variation is subject to tight environmental control, which is mediated by at least six global regulators (see Fig. S4 in the supplemental material). For example, the leucine response protein LrpA mediates the response to amino acids (e.g., alanine, isoleucine, leucine, and valine). LrpA binds directly to the switch, affecting fimB- and fimE-promoted switching (35). Similarly, IHF affects switching by both recombinases (15), while H-NS affects only the fimB-mediated inversion (92). In microarray experiments, LrpA (56) and H-NS (53) had an overall positive effect on the levels of expression of the fim genes. Interestingly, LrhA had a positive effect on the level of expression of fimE and a negative effect on the level of expression of the fim operon (16). This was likely due to the strong bias for phase switching from the phase-ON to the phase-OFF orientation of FimE. Other microarray studies showed that EnvZ/OmpR had a negative effect on the levels of expression of the fim operon (97).

THE REGULATORS

Within the network, three regulators (FlhDC, EnvZ/OmpR, and RcsCDB) affect expression of the majority of the genes, primarily the genes involved in the biogenesis of flagella, curli, and type I fimbriae.

FlhDC.

Initially described as the master regulator of flagellum biogenesis in E. coli and S. enterica (68-70, 129, 130), FlhDC also regulates nonflagellar genes (110, 111) (see Fig. S5 in the supplemental material). Encoded by the flhDC operon (11), FlhDC sits atop a transcriptional hierarchy of flagellar genes (for reviews of flagellar hierarchy, see references 2, 24, 69, and 108). The FlhDC complex binds the upstream regions of three flagellar operons (fliA, fliL, and flhB) (77) and activates their transcription from σ70-dependent promoters. FlhDC also activates transcription directly from a subset of promoters that depend upon σ28, the product of fliA. The remaining σ28-dependent promoters are under indirect control of FlhDC through its activation of fliA (50, 78, 94).

Twenty-nine nonflagellar FlhDC-dependent operons in E. coli were revealed by microarray analysis (110). Approximately one-half of these operons function in respiration. Transcription of the operons that encode aerobic respiratory pathways was inhibited, while transcription of the operons that encode anaerobic pathways was enhanced. This enhancement, as well as that of the Entner-Doudoroff pathway, was mediated by the oxygen sensor and chemoreceptor Aer (110). In addition, FlhDC enhanced transcription of the two curli operons, csgB and csgD. Finally, it modulated transcription of a number of genes encoding transporters and enzymes involved in amino acid metabolism.

In addition to this experimental evidence, bioinformatic analysis suggests that there are additional FlhDC targets. A consensus sequence for putative FlhDC binding sites was developed and used to identify putative targets (136). The promoter regions of four of these genes (b1904, b2446, wzzfepE, and gltI) showed both binding and regulation by FlhDC. In addition, a FliA consensus sequence was proposed and used to identify several putative FliA targets (99). Two of these targets (ygbK and ppdAB) also were dependent on FlhDC, as determined with promoter-lacZ fusions.

Envz/OmpR.

EnvZ/OmpR, a two-component signal transduction pathway originally shown to regulate expression of the outer membrane porins OmpF and OmpC (79, 148), also controls expression of more than 100 nonporin genes (97) (see Fig. S6 in the supplemental material). EnvZ/OmpR regulates transcription of ompF and ompC inversely; at low osmolarity, it activates ompF, and at high osmolarity it represses ompF while activating ompC (4). To activate transcription, OmpR binds three tandem sites upstream of and proximal to the −35 hexamers of both ompC (C1 to C3) and ompF (F1 to F3) (55, 79, 87, 148). To repress ompF transcription, OmpR binds a fourth distal site (F4) (55, 98). Occupancy of this distal site is believed to facilitate formation of a DNA loop between OmpR bound at F4 and OmpR bound to one or more of the proximal binding sites (F1 to F3). The binding of OmpR to C1 to C3 and to F1 to F4 seems to be independent of the degree of OmpR phosphorylation (48). Rather, the binding appears to be mediated by an osmolarity-induced conformational change (83).

The non-porin-associated functions of OmpR include regulation of the permease encoded by tppB in S. enterica (41) and E. coli (42), the maltose regulator encoded by malT (23), and the murein regulator encoded by bolA (160). A recent microarray study (97) identified 125 OmpR-dependent genes. The phenotypes exhibited by an ompR-envZ mutant include increased resistance to several antibiotics (attributed to the defect in porin synthesis) and increased use of several hexoses as carbon sources (allose, fructose, mannitol, N-acetyl-d-glucosamine, and glucose) (163).

Other cellular processes affected by OmpR include the biogenesis of curli (59), type I fimbriae (97), and flagella (128). DNase I footprinting demonstrated that there is direct binding of OmpR to the flhDC promoter at two discrete regions (128). This arrangement resembles that present at ompF; thus, a repression loop similar to that predicted for ompF might be responsible for repression of flhDC transcription. In contrast to ompF repression, regulation of flhDC depends on the phosphorylation state of OmpR. Phosphorylated OmpR bound the flhDC promoter with 10-fold-higher affinity than unphosphorylated OmpR bound the flhDC promoter (128). Electrophoretic mobility shift assays have demonstrated that there is binding of phosphorylated OmpR to the csgD promoter, which drives expression of one of the two curli operons (59, 106).

RcsCDB.

The RcsCDB phosphorelay, discovered as a regulator of capsule synthesis (43), is responsible for the regulation of up to 5% of the E. coli genome (32, 44, 97) (see Fig. S7 in the supplemental material). Many of the target genes encode parts of surface appendages (e.g., flagella and curli), components of the cellular multistress response (e.g., osmB, osmY, and osmC), or proteins involved in cell division (ftsAZ) (22, 30, 32, 33, 140). RcsB can bind either as a homodimer to the RcsB box (e.g., at ftsAZ and osmC [22, 30, 140]) or as a heterodimer in a complex with the auxiliary protein RcsA (e.g., at cps [60, 139, 152, 153]). RcsA is related to the response regulators, except for the lack of the conserved aspartate site that is required for phosphorylation (139). The RcsAB box resembles the RcsB box. The differences in the consensus sequences are indicative of the presence of RcsA in the heterodimer, and it was hypothesized that the conformation of RcsB might be modulated upon interaction with RcsA, resulting in recognition of different DNA targets (107).

Like EnvZ/OmpR, the RcsCDB phosphorelay regulates the biogenesis of flagella, curli, and type 1 fimbriae. It may also regulate an uncharacterized fimbrial locus (sfm). Regulation of the flagellar system by RcsCDB was shown first in Proteus mirabilis (12) and later in E. coli (33, 143). The 2CST regulator RcsB binds directly to the flhDC promoter to inhibit its transcription. This regulation may also involve RcsA (33), but only when an excess of it is present (C. E. Fredericks and A. J. Wolfe, unpublished). A recent study provided evidence that RcsCDB activates fim expression (Fredericks, and Wolfe, unpublished), while a microarray analysis indicated that RcsCDB negatively regulates both the biogenesis of curli and the expression of fimZ (32). In S. enterica, the 2CST regulator FimZ activates fim, while it represses flhDC (28). In E. coli, however, FimZ probably does not regulate the fim locus but rather regulates the sfm (salmonella-like fimbriae) locus in which it resides (http://genolist.pasteur.fr/Colibri/, http://ca.expasy.org/sprot/). If this is true, then RcsCDB regulates fim and sfm inversely, increasing fim expression while decreasing expression of sfm. Whether FimZ regulates flhDC in E. coli remains unknown.

Taken together, this evidence provides strong support for the hypothesis that the RcsCDB phosphorelay plays an important role in adapting the bacterial cell surface to growth on a solid surface (32) and, thus, a critical role in the development of biofilms (L. Ferrieres and D. Clarke, personal communication).

SMALL MOLECULES

In E. coli, signal transduction pathways either can produce small molecules as second messengers or can be influenced by small molecules. Below, we discuss the impact of three of these molecules, cAMP, acetyl∼P, and c-di-GMP, on our network.

Cyclic AMP.

The product of a signal transduction pathway that consists of the phosphoenolpyruvate:carbohydrate phosphotransferase system and adenylate cyclase, cAMP is a second messenger that reports on the nutritional status of the external environment. When levels of catabolite-repressing carbon sources decrease, cAMP levels increase (105). The cAMP then docks with CRP to activate the transcription of genes required for the metabolism of secondary carbon sources and other cellular processes (for a review, see reference 46), including the biogenesis of flagella and curli.

Acetyl phosphate.

The intermediate of the phosphotransacetylase-acetate kinase pathway (21, 118), acetyl∼P, has a larger ΔG0 of hydrolysis than ATP (76). Thus, acetyl∼P stores more energy than ATP stores and, indeed, donates its phosphoryl group to ADP to generate ATP. This tendency to donate phosphoryl groups also forms the basis for its proposed impact on 2CST pathways (85, 151).

There is much evidence which supports the hypothesis that acetyl∼P can interact with 2CST pathways. In vitro, many response regulators autophosphorylate using acetyl∼P as the phosphoryl donor. Numerous in vivo studies have shown that there is a strong correlation between the status of the acetyl∼P pool and activation of some 2CST targets, implicating acetyl∼P in the activation of a subset of response regulators (for a review, see reference 157). One of these studies demonstrated that acetyl∼P can influence the in vivo expression of almost 100 genes (158), verifying that acetyl∼P correlates with decreased expression of genes involved in flagellum biogenesis (113) and showing that it correlates with increased expression of genes involved in type 1 fimbria assembly (fim), the biosynthesis of capsule (cps), and the response to multiple stresses (e.g., osmB, osmY, and osmC) (158). These results can be explained, in part, by the following observations: (i) acetyl∼P can donate its phosphoryl group to both OmpR (62) and RcsB (F. Bernhard, personal communication), (ii) the Rcs phosphorelay controls the biosynthesis of capsule (138) and many of the stress-associated genes (30, 32), (iii) both OmpR and the Rcs phosphorelay regulate the biogenesis of flagella, curli, and type 1 fimbriae (32, 33, 97, 109, 128), and (iv) acetyl∼P acts upon capsule biosynthesis and flagellum biogenesis via the RcsCDB phosphorelay (Fredericks, and Wolfe, unpublished).

Cyclic di-GMP.

The second messenger, c-di-GMP, also regulates the transition from motile, planktonic cells to a sessile biofilm. Like acetyl∼P, it inhibits flagellum biogenesis while enhancing capsule biosynthesis (for recent reviews, see references 58, 116, and 131). However, in contrast to cAMP and acetyl∼P, which influence this transition by controlling transcription initiation, c-di-GMP tends to act posttranslationally (3, 49, 54, 64, 103).

c-di-GMP is synthesized by diguanylate cyclases (DGCs) and is degraded by phosphodiesterases (PDEs). DGC activity has been associated with the highly conserved GGDEF domain (101, 102, 120, 131, 144), while PDE activity has been associated with the highly conserved EAL domain (17, 26, 145). GGDEF and EAL domains are ubiquitous in bacteria (133). On the basis of sheer abundance, they represent a major family of signaling pathways (37). Many bacterial species possess multiple proteins with GGDEF and/or EAL domains. For example, Pseudomonas aeruginosa has 33 such proteins, Vibrio cholerae possesses 41, and E. coli has 36 (58, 116, 131). This abundance suggests that there is a network of pathways that either integrates multiple signals into a single second messenger or instead permits synthesis of the second messenger in response to diverse signals (38). More likely, pathways work in relative isolation due to localization or the existence of microenvironments (54, 101, 131).

Processes influenced by c-di-GMP also are abundant, but most of them result in phenotypic changes that are related to the transition between motile, planktonic individuals and a sessile biofilm (131). The mechanisms used by c-di-GMP to influence behavior remain obscure, although this molecule likely works by direct interaction with its targets (37). Such is the case with the biosynthesis of cellulose, an extracellular polysaccharide, in Acetobacter xylinum and S. enterica. In these organisms, multiple DGCs and PDEs regulate the intracellular concentration of c-di-GMP, which binds directly to a cellulose synthesis complex that includes BscA, which consists of the newly discovered c-di-GMP-binding domain PilZ attached to a glycosyltransferase (5). The result is an activated complex capable of synthesizing cellulose, which is required for the formation of biofilms and the development of rugose colonies (57, 114, 119, 131, 155). A second example is YcgR, which consists of a PilZ domain attached to a domain whose function is unknown (5). YcgR, along with the EAL domain protein YhjH, has been implicated in the ability of E. coli flagella to rotate (67) The mechanism is not understood, nor are the mechanisms that underly the regulation of other targets of c-di-GMP understood.

CONCLUDING REMARKS

In summary, here we describe the regulation by three global regulators of three cellular processes involved in early biofilm development. FlhDC is a positive regulator of flagella and curli, OmpR is a negative regulator of flagella and type I fimbriae and a positive regulator of curli, and RcsCDB is a negative regulator of flagella and curli and a positive regulator of type 1 fimbriae and capsule (Fig. 3). The differential use of these three global regulators to integrate diverse signals, second messengers, and metabolites likely provides much of the basis for the ability of cells to coordinate surface organelle biogenesis so that they can build a proper biofilm. Additional global regulators (e.g., CRP and H-NS) and more specific regulators (e.g., HdfR and CsgD) could provide an opportunity to further calibrate the process.

FIG. 3.

FIG. 3.

Regulation of biofilm formation by FlhDC, OmpR/EnvZ, and RcsCDB. Observations from the network are projected on a time course of the early stages of biofilm formation. Biofilm formation starts with reversible attachment, proceeds to irreversible attachment, and progresses to three-dimensional construction of the mature biofilm, until the biofilm disassociates. The action points of the three major regulators in our network are indicated by vertical arrows. These points are based on the cellular processes that are affected by the regulators. A time course of expression of the regulators is hypothesized and is indicated by horizontal arrows, paralleling the time course of biofilm formation.

We see this minireview as a semiglobal approach to relate information about the entire network to a specific biological question. While the ultimate goal of systems biology is to decipher the entire regulatory network of the cell, here we focused on one part of that network, the sector that controls major cellular processes involved in early biofilm development. We envision this network as just one system in which multiple environmental signals feed into numerous global regulators to regulate diverse cellular processes involved in a complex behavior. We anticipate that there are other systems.

Supplementary Material

[Supplemental material]

Acknowledgments

We thank David Clarke and Lionel Ferrieres (University of Bath, Bath, United Kingdom) and Frank Bernhard (Johann Wolfgang Goethe Universität, Frankfurt, Germany) for personal communications, Philip Matsumura (University of Illinois at Chicago, Chicago, IL) and Christine Fredericks (Loyola University Chicago, Maywood, IL) for providing unpublished data, Karen Visick and David Keating (Loyola University Chicago, Maywood, IL) for helpful discussions, and Clive Barker (University of Illinois at Chicago, Chicago, IL) for critically reading the manuscript and improving the figures.

B.M.P. was supported by the ND EPSCoR program of North Dakota (through grant EPS-0132289 from the National Science Foundation), the North Dakota Agricultural Experiment Station, and the “Biosecurity, Disease Surveillance, and Food Safety” earmark grant (through USDA APHIS). A.D. and C.B. were supported by grant IDM-0415190 from the National Science Foundation. A.J.W. was supported by grant GM066130 from the National Institute of General Medical Sciences and by grant LU#11200 from the Loyola University Potts Foundation.

Footnotes

Supplemental material for this article may be found at http://jb.asm.org/.

REFERENCES

  • 1.Adler, J., and B. Templeton. 1967. The effect of environmental conditions on the motility of Escherichia coli. J. Gen. Microbiol. 46:175-184. [DOI] [PubMed] [Google Scholar]
  • 2.Aldridge, P., and K. T. Hughes. 2002. Regulation of flagellar assembly. Curr. Opin. Microbiol. 5:160-165. [DOI] [PubMed] [Google Scholar]
  • 3.Aldridge, P., and U. Jenal. 1999. Cell cycle-dependent degradation of a flagellar motor component requires a novel-type response regulator. Mol. Microbiol. 32:379-391. [DOI] [PubMed] [Google Scholar]
  • 4.Alphen, W. V., and B. Lugtenberg. 1977. Influence of osmolarity of the growth medium on the outer membrane protein pattern of Escherichia coli. J. Bacteriol. 131:623-630. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 5.Amikam, D., and M. Y. Galperin. 2006. PilZ domain is part of the bacterial c-di-GMP binding protein. Bioinformatics 22:3-6. [DOI] [PubMed] [Google Scholar]
  • 6.Appleby, J. L., J. S. Parkinson, and R. B. Bourret. 1996. Signal transduction via the multi-step phosphorelay: not necessarily a road less traveled. Cell 86:845-848. [DOI] [PubMed] [Google Scholar]
  • 7.Arnqvist, A., A. Olsen, and S. Normark. 1994. Sigma S-dependent growth-phase induction of the csgBA promoter in Escherichia coli can be achieved in vivo by sigma 70 in the absence of the nucleoid-associated protein H-NS. Mol. Microbiol. 13:1021-1032. [DOI] [PubMed] [Google Scholar]
  • 8.Barker, C. S., and B. M. Prüß. 2005. FlhD/FlhC, a global transcriptional regulator in Escherichia coli, p. 13-30. In B. M. Prüß (ed.), Global regulatory networks in enteric bacteria. Research Signpost, Trivandrum, Kerala, India.
  • 9.Barker, C. S., B. M. Prüß, and P. Matsumura. 2004. Increased motility of Escherichia coli by insertion sequence element integration into the regulatory region of the flhD operon. J. Bacteriol. 186:7529-7537. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 10.Barnich, N., J. Boudeau, L. Claret, and A. rfeuille-Michaud. 2003. Regulatory and functional co-operation of flagella and type 1 pili in adhesive and invasive abilities of AIEC strain LF82 isolated from a patient with Crohn's disease. Mol. Microbiol. 48:781-794. [DOI] [PubMed] [Google Scholar]
  • 11.Bartlett, D. H., B. B. Frantz, and P. Matsumura. 1988. Flagellar transcriptional activators FlbB and FlaI: gene sequences and 5′ consensus sequences of operons under FlbB and FlaI control. J. Bacteriol. 170:1575-1581. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 12.Belas, R., R. Schneider, and M. Melch. 1998. Characterization of Proteus mirabilis precocious swarming mutants: identification of rsbA, encoding a regulator of swarming behavior. J. Bacteriol. 180:6126-6139. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 13.Bertin, P., F. Hommais, E. Krin, O. Soutourina, C. Tendeng, S. Derzelle, and A. Danchin. 2001. H-NS and H-NS-like proteins in Gram-negative bacteria and their multiple role in the regulation of bacterial metabolism. Biochimie 83:235-241. [DOI] [PubMed] [Google Scholar]
  • 14.Bertin, P., E. Terao, E. H. Lee, P. Lejeune, C. Colson, A. Danchin, and E. Collatz. 1994. The H-NS protein is involved in the biogenesis of flagella in Escherichia coli. J. Bacteriol. 176:5537-5540. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 15.Blomfield, I. C., D. H. Kulasekara, and B. I. Eisenstein. 1997. Integration host factor stimulates both FimB- and FimE-mediated site-specific DNA inversion that controls phase variation of type 1 fimbriae expression in Escherichia coli. Mol. Microbiol. 23:705-717. [DOI] [PubMed] [Google Scholar]
  • 16.Blumer, C., Q. H. Tran, P. M. Selzer, A. Kleefeld, D. Weberskirch, M. Heintz, and G. Unden. 2005. Regulation of motility, chemotaxis, virulence and biofilm formation by the LysR-type regulator LrhA in E. coli and related bacteria, p. 75-92. In B. M. Prüß (ed.), Global regulatory networks in enteric bacteria. Research Signppost, Trivandrum, Kerala, India.
  • 17.Bobrov, A. G., O. Kirillina, and R. D. Perry. 2005. The phosphodiesterase activity of the HmsP EAL domain is required for negative regulation of biofilm formation in Yersinia pestis. FEMS Microbiol. Lett. 247:123-130. [DOI] [PubMed] [Google Scholar]
  • 18.Bougdour, A., C. Lelong, and J. Geiselmann. 2004. Crl, a low temperature-induced protein in Escherichia coli that binds directly to the stationary phase sigma subunit of RNA polymerase. J. Biol. Chem. 279:19540-19550. [DOI] [PubMed] [Google Scholar]
  • 19.Brombacher, E., C. Dorel, A. J. B. Zehnder, and P. Landini. 2003. The curli biosynthesis regulator CsgD co-ordinates the expression of both positive and negative determinants for biofilm formation in Escherichia coli. Microbiology 149:2847-2857. [DOI] [PubMed] [Google Scholar]
  • 20.Brown, P. K., C. M. Dozois, C. A. Nickerson, A. Zuppardo, J. Terlonge, and R. Curtiss III. 2001. MlrA, a novel regulator of curli (AgF) and extracellular matrix synthesis by Escherichia coli and Salmonella enterica serovar Typhimurium. Mol. Microbiol. 41:349-363. [DOI] [PubMed] [Google Scholar]
  • 21.Brown, T. D., M. C. Jones-Mortimer, and H. L. Kornberg. 1977. The enzymic interconversion of acetate and acetyl-coenzyme A in Escherichia coli. J. Gen. Microbiol. 102:327-336. [DOI] [PubMed] [Google Scholar]
  • 22.Carballes, F., C. Bertrand, J. P. Bouche, and K. Cam. 1999. Regulation of Escherichia coli cell division genes ftsA and ftsZ by the two-component system rcsC-rcsB. Mol. Microbiol. 34:442-450. [DOI] [PubMed] [Google Scholar]
  • 23.Case, C. C., B. Bukau, S. Granett, M. R. Villarejo, and W. Boos. 1986. Contrasting mechanisms of envZ control of mal and pho regulon genes in Escherichia coli. J. Bacteriol. 166:706-712. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 24.Chilcott, G. S., and K. T. Hughes. 2000. Coupling of flagellar gene expression to flagellar assembly in Salmonella enterica serovar Typhimurium and Escherichia coli. Microbiol. Mol. Biol. Rev. 64:694-708. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 25.Chirwa, N. T., and M. B. Herrington. 2003. CsgD, a regulator of curli and cellulose synthesis, also regulates serine hydroxymethyltransferase synthesis in Escherichia coli K-12. Microbiology 149:525-535. [DOI] [PubMed] [Google Scholar]
  • 26.Christen, M., B. Christen, M. Folcher, A. Schauerte, and U. Jenal. 2005. Identification and characterization of a cyclic di-GMP-specific phosphodiesterase and its allosteric control by GTP. J. Biol. Chem. 280:30829-30837. [DOI] [PubMed] [Google Scholar]
  • 27.Clarke, M. B., and V. Sperandio. 2005. Transcriptional regulation of flhDC by QseBC and sigma 28 (FliA) in enterohaemorrhagic Escherichia coli. Mol. Microbiol. 57:1734-1749. [DOI] [PubMed] [Google Scholar]
  • 28.Clegg, S., and K. T. Hughes. 2002. FimZ is a molecular link between sticking and swimming in Salmonella enterica serovar Typhimurium. J. Bacteriol. 184:1209-1213. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 29.Connell, I., W. Agace, P. Klemm, M. Schembri, S. Marild, and C. Svanborg. 1996. Type 1 fimbrial expression enhances Escherichia coli virulence for the urinary tract. Proc. Natl. Acad. Sci. USA 93:9827-9832. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 30.Davalos-Garcia, M., A. Conter, I. Toesca, C. Gutierrez, and K. Cam. 2001. Regulation of osmC gene expression by the two-component system rcsB-rcsC in Escherichia coli. J. Bacteriol. 183:5870-5876. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 31.Duguay, A. R., and T. J. Silhavy. 2004. Quality control in the bacterial periplasm. Biochim. Biophys. Acta 1694:121-134. [DOI] [PubMed] [Google Scholar]
  • 32.Ferrieres, L., and D. J. Clarke. 2003. The RcsC sensor kinase is required for normal biofilm formation in Escherichia coli K-12 and controls the expression of a regulon in response to growth on a solid surface. Mol. Microbiol. 50:1665-1682. [DOI] [PubMed] [Google Scholar]
  • 33.Francez-Charlot, A., B. Laugel, G. A. Van, N. Dubarry, F. Wiorowski, M. P. Castanie-Cornet, C. Gutierrez, and K. Cam. 2003. RcsCDB His-Asp phosphorelay system negatively regulates the flhDC operon in Escherichia coli. Mol. Microbiol. 49:823-832. [DOI] [PubMed] [Google Scholar]
  • 34.Gally, D. L., J. Leathart, and I. C. Blomfield. 1996. Interaction of FimB and FimE with the fim switch that controls the phase variation of type 1 fimbriae in Escherichia coli K-12. Mol. Microbiol. 21:725-738. [DOI] [PubMed] [Google Scholar]
  • 35.Gally, D. L., T. J. Rucker, and I. C. Blomfield. 1994. The leucine-responsive regulatory protein binds to the fim switch to control phase variation of type 1 fimbrial expression in Escherichia coli K-12. J. Bacteriol. 176:5665-5672. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 36.Galperin, M. Y. 2004. Bacterial signal transduction network in a genomic perspective. Environ. Microbiol. 6:552-567. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 37.Galperin, M. Y. 2005. A census of membrane-bound and intracellular signal transduction proteins in bacteria: bacterial IQ, extroverts and introverts. BMC Microbiol. 5:35. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 38.Garcia, B., C. Latasa, C. Solano, P. F. Garcia-del, C. Gamazo, and I. Lasa. 2004. Role of the GGDEF protein family inSalmonella cellulose biosynthesis and biofilm formation. Mol. Microbiol. 54:264-277. [DOI] [PubMed] [Google Scholar]
  • 39.Gerstel, U., and U. Romling. 2001. Oxygen tension and nutrient starvation are major signals that regulate agfD promoter activity and expression of the multicellular morphotype in Salmonella typhimurium. Environ. Microbiol. 3:638-648. [DOI] [PubMed] [Google Scholar]
  • 40.Gerstel, U., and U. Romling. 2003. The csgD promoter, a control unit for biofilm formation in Salmonella typhimurium. Res. Microbiol. 154:659-667. [DOI] [PubMed] [Google Scholar]
  • 41.Gibson, M. M., E. M. Ellis, K. A. Graeme-Cook, and C. F. Higgins. 1987. OmpR and EnvZ are pleiotropic regulatory proteins: positive regulation of the tripeptide permease (tppB) of Salmonella typhimurium. Mol. Gen. Genet. 207:120-129. [DOI] [PubMed] [Google Scholar]
  • 42.Goh, E. B., D. F. Siino, and M. M. Igo. 2004. The Escherichia coli tppB (ydgR) gene represents a new class of OmpR-regulated genes. J. Bacteriol. 186:4019-4024. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 43.Gottesman, S., P. Trisler, and A. Torres-Cabassa. 1985. Regulation of capsular polysaccharide synthesis in Escherichia coli K-12: characterization of three regulatory genes. J. Bacteriol. 162:1111-1119. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 44.Hagiwara, D., M. Sugiura, T. Oshima, H. Mori, H. Aiba, T. Yamashino, and T. Mizuno. 2003. Genome-wide analyses revealing a signaling network of the RcsC-YojN-RcsB phosphorelay system in Escherichia coli. J. Bacteriol. 185:5735-5746. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 45.Hammar, M., A. Arnqvist, Z. Bian, A. Olsen, and S. Normark. 1995. Expression of two csg operons is required for production of fibronectin- and Congo red-binding curli polymers in Escherichia coli K-12. Mol. Microbiol. 18:661-670. [DOI] [PubMed] [Google Scholar]
  • 46.Harman, J. G. 2001. Allosteric regulation of the cAMP receptor protein. Biochim. Biophys. Acta 1547:1-17. [DOI] [PubMed] [Google Scholar]
  • 47.Harshey, R. M., I. Kawagishi, J. Maddock, and L. J. Kenney. 2003. Function, diversity, and evolution of signal transduction in prokaryotes. Dev. Cell 4:459-465. [DOI] [PubMed] [Google Scholar]
  • 48.Head, C. G., A. Tardy, and L. J. Kenney. 1998. Relative binding affinities of OmpR and OmpR-phosphate at the ompF and ompC regulatory sites. J. Mol. Biol. 281:857-870. [DOI] [PubMed] [Google Scholar]
  • 49.Hecht, G. B., and A. Newton. 1995. Identification of a novel response regulator required for the swarmer-to-stalked-cell transition in Caulobacter crescentus. J. Bacteriol. 177:6223-6229. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 50.Helmann, J. D., L. M. Marquez, and M. J. Chamberlin. 1988. Cloning, sequencing, and disruption of the Bacillus subtilis sigma 28 gene. J. Bacteriol. 170:1568-1574. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 51.Herman, I., G. Melancon, and M. S. Marshall. 2000. Graph visualization and navigation in information visualization: a survey. IEEE Trans. Vis. Comput. Graphics 6:24-43.
  • 52.Holden, N. J., B. E. Uhlin, and D. L. Gally. 2001. PapB paralogues and their effect on the phase variation of type 1 fimbriae in Escherichia coli. Mol. Microbiol. 42:319-330. [DOI] [PubMed] [Google Scholar]
  • 53.Hommais, F., E. Krin, C. Laurent-Winter, O. Soutourina, A. Malpertuy, J. P. Le Caer, A. Danchin, and P. Bertin. 2001. Large-scale monitoring of pleiotropic regulation of gene expression by the prokaryotic nucleoid-associated protein, H-NS. Mol. Microbiol. 40:20-36. [DOI] [PubMed] [Google Scholar]
  • 54.Huang, B., C. B. Whitchurch, and J. S. Mattick. 2003. FimX, a multidomain protein connecting environmental signals to twitching motility in Pseudomonas aeruginosa. J. Bacteriol. 185:7068-7076. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 55.Huang, K. J., and M. M. Igo. 1996. Identification of the bases in the ompF regulatory region, which interact with the transcription factor OmpR. J. Mol. Biol. 262:615-628. [DOI] [PubMed] [Google Scholar]
  • 56.Hung, S. P., P. Baldi, and G. W. Hatfield. 2002. Global gene expression profiling in Escherichia coli K12. The effects of leucine-responsive regulatory protein. J. Biol. Chem. 277:40309-40323. [DOI] [PubMed] [Google Scholar]
  • 57.Jenal, U. 2004. Cyclic di-guanosine-monophosphate comes of age: a novel secondary messenger involved in modulating cell surface structures in bacteria? Curr. Opin. Microbiol. 7:185-191. [DOI] [PubMed] [Google Scholar]
  • 58.Jenal, U., R. E. Silversmith, L. Sogaard-Andersen, and L. Sockett. 2005. Sense and sensibility in bacteria. VIIIth International Conference on Bacterial Locomotion and Sensory Transduction. EMBO Rep. 6:615-619. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 59.Jubelin, G., A. Vianney, C. Beloin, J. M. Ghigo, J. C. Lazzaroni, P. Lejeune, and C. Dorel. 2005. CpxR/OmpR interplay regulates curli gene expression in response to osmolarity in Escherichia coli. J. Bacteriol. 187:2038-2049. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 60.Kelm, O., C. Kiecker, K. Geider, and F. Bernhard. 1997. Interaction of the regulator proteins RcsA and RcsB with the promoter of the operon for amylovoran biosynthesis in Erwinia amylovora. Mol. Gen. Genet. 256:72-83. [DOI] [PubMed] [Google Scholar]
  • 61.Kenney, L. J. 2002. Structure/function relationships in OmpR and other winged-helix transcription factors. Curr. Opin. Microbiol. 5:135-141. [DOI] [PubMed] [Google Scholar]
  • 62.Kenney, L. J., M. D. Bauer, and T. J. Silhavy. 1995. Phosphorylation-dependent conformational changes in OmpR, an osmoregulatory DNA-binding protein of Escherichia coli. Proc. Natl. Acad. Sci. USA 92:8866-8870. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 63.Kikuchi, T., Y. Mizunoe, A. Takade, S. Naito, and S. Yoshida. 2005. Curli fibers are required for development of biofilm architecture in Escherichia coli K-12 and enhance bacterial adherence to human uroepithelial cells. Microbiol. Immunol. 49:875-884. [DOI] [PubMed] [Google Scholar]
  • 64.Kirillina, O., J. D. Fetherston, A. G. Bobrov, J. Abney, and R. D. Perry. 2004. HmsP, a putative phosphodiesterase, and HmsT, a putative diguanylate cyclase, control Hms-dependent biofilm formation in Yersinia pestis. Mol. Microbiol. 54:75-88. [DOI] [PubMed] [Google Scholar]
  • 65.Kitamura, E., Y. Nakayama, H. Matsuzaki, K. Matsumoto, and I. Shibuya. 1994. Acidic-phospholipid deficiency represses the flagellar master operon through a novel regulatory region in Escherichia coli. Biosci. Biotechnol. Biochem. 58:2305-2307. [DOI] [PubMed] [Google Scholar]
  • 66.Klemm, P. 1986. Two regulatory fim genes, fimB and fimE, control the phase variation of type 1 fimbriae in Escherichia coli. EMBO J. 5:1389-1393. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 67.Ko, M., and C. Park. 2000. H-NS-dependent regulation of flagellar synthesis is mediated by a LysR family protein. J. Bacteriol. 182:4670-4672. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 68.Komeda, Y. 1982. Fusions of flagellar operons to lactose genes on a Mu lac bacteriophage. J. Bacteriol. 150:16-26. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 69.Komeda, Y. 1986. Transcriptional control of flagellar genes in Escherichia coli K-12. J. Bacteriol. 168:1315-1318. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 70.Komeda, Y., K. Kutsukake, and T. Iino. 1980. Definition of additional flagellar genes in Escherichia coli K12. Genetics 94:277-290. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 71.Kunin, C. M., T. H. Hua, and L. O. Bakaletz. 1995. Effect of salicylate on expression of flagella by Escherichia coli and Proteus, Providencia, and Pseudomonas spp. Infect. Immun. 63:1796-1799. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 72.Kunin, C. M., T. H. Hua, R. L. Guerrant, and L. O. Bakaletz. 1994. Effect of salicylate, bismuth, osmolytes, and tetracycline resistance on expression of fimbriae by Escherichia coli. Infect. Immun. 62:2178-2186. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 73.Lehnen, D., C. Blumer, T. Polen, B. Wackwitz, V. F. Wendisch, and G. Unden. 2002. LrhA as a new transcriptional key regulator of flagella, motility and chemotaxis genes in Escherichia coli. Mol. Microbiol. 45:521-532. [DOI] [PubMed] [Google Scholar]
  • 74.Li, C., C. J. Louise, W. Shi, and J. Adler. 1993. Adverse conditions which cause lack of flagella in Escherichia coli. J. Bacteriol. 175:2229-2235. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 75.Li, X., D. A. Rasko, C. V. Lockatell, D. E. Johnson, and H. L. Mobley. 2001. Repression of bacterial motility by a novel fimbrial gene product. EMBO J. 20:4854-4862. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 76.Lipmann, F. 1941. Metabolic generation and utilization of phosphate bond energy. Adv. Enzymol. 1:99-162. [Google Scholar]
  • 77.Liu, X., and P. Matsumura. 1994. The FlhD/FlhC complex, a transcriptional activator of the Escherichia coli flagellar class II operons. J. Bacteriol. 176:7345-7351. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 78.Liu, X., and P. Matsumura. 1995. An alternative sigma factor controls transcription of flagellar class-III operons in Escherichia coli: gene sequence, overproduction, purification and characterization. Gene 164:81-84. [DOI] [PubMed] [Google Scholar]
  • 79.Maeda, S., and T. Mizuno. 1990. Evidence for multiple OmpR-binding sites in the upstream activation sequence of the ompC promoter in Escherichia coli: a single OmpR-binding site is capable of activating the promoter. J. Bacteriol. 172:501-503. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 80.Majdalani, N., and S. Gottesman. 2005. The Rcs phosphorelay: a complex signal transduction system. Annu. Rev. Microbiol. 59:379-405. [DOI] [PubMed] [Google Scholar]
  • 81.Marc, D., and M. Dho-Moulin. 1996. Analysis of the fim cluster of an avian O2 strain of Escherichia coli: serogroup-specific sites within fimA and nucleotide sequence of fimI. J. Med. Microbiol. 44:444-452. [DOI] [PubMed] [Google Scholar]
  • 82.Martinez, J. J., M. A. Mulvey, J. D. Schilling, J. S. Pinkner, and S. J. Hultgren. 2000. Type 1 pilus-mediated bacterial invasion of bladder epithelial cells. EMBO J. 19:2803-2812. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 83.Mattison, K., R. Oropeza, N. Byers, and L. J. Kenney. 2002. A phosphorylation site mutant of OmpR reveals different binding conformations at ompF and ompC. J. Mol. Biol. 315:497-511. [DOI] [PubMed] [Google Scholar]
  • 84.Maurer, J. J., T. P. Brown, W. L. Steffens, and S. G. Thayer. 1998. The occurrence of ambient temperature-regulated adhesins, curli, and the temperature-sensitive hemagglutinin Tsh among avian Escherichia coli. Avian Dis. 42:106-118. [PubMed] [Google Scholar]
  • 85.McCleary, W. R., J. B. Stock, and A. J. Ninfa. 1993. Is acetyl phosphate a global signal in Escherichia coli? J. Bacteriol. 175:2793-2798. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 86.Mizuno, T. 1997. Compilation of all genes encoding two-component phosphotransfer signal transducers in the genome of Escherichia coli. DNA Res. 4:161-168. [DOI] [PubMed] [Google Scholar]
  • 87.Mizuno, T., M. Kato, Y. L. Jo, and S. Mizushima. 1988. Interaction of OmpR, a positive regulator, with the osmoregulated ompC and ompF genes of Escherichia coli. Studies with wild-type and mutant OmpR proteins. J. Biol. Chem. 263:1008-1012. [PubMed] [Google Scholar]
  • 88.Mizushima, T., R. Koyanagi, T. Katayama, T. Miki, and K. Sekimizu. 1997. Decrease in expression of the master operon of flagellin synthesis in a dnaA46 mutant of Escherichia coli. Biol. Pharm. Bull. 20:327-331. [DOI] [PubMed] [Google Scholar]
  • 89.Mizushima, T., R. Koyanagi, E. Suzuki, A. Tomura, K. Kutsukake, T. Miki, and K. Sekimizu. 1995. Control by phosphatidylglycerol of expression of the flhD gene in Escherichia coli. Biochim. Biophys. Acta 1245:397-401. [DOI] [PubMed] [Google Scholar]
  • 90.Mizushima, T., A. Tomura, T. Shinpuku, T. Miki, and K. Sekimizu. 1994. Loss of flagellation in dnaA mutants of Escherichia coli. J. Bacteriol. 176:5544-5546. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 91.Nevesinjac, A. Z., and T. L. Raivio. 2005. The Cpx envelope stress response affects expression of the type IV bundle-forming pili of enteropathogenic Escherichia coli. J. Bacteriol. 187:672-686. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 92.O'Gara, J. P., and C. J. Dorman. 2000. Effects of local transcription and H-NS on inversion of the fim switch of Escherichia coli. Mol. Microbiol. 36:457-466. [DOI] [PubMed] [Google Scholar]
  • 93.Ogasawara, H., J. Teramoto, K. Hirao, K. Yamamoto, A. Ishihama, and R. Utsumi. 2004. Negative regulation of DNA repair gene (ung) expression by the CpxR/CpxA two-component system in Escherichia coli K-12 and induction of mutations by increased expression of CpxR. J. Bacteriol. 186:8317-8325. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 94.Ohnishi, K., K. Kutsukake, H. Suzuki, and T. Iino. 1990. Gene fliA encodes an alternative sigma factor specific for flagellar operons in Salmonella typhimurium. Mol. Gen. Genet. 221:139-147. [DOI] [PubMed] [Google Scholar]
  • 95.Olsen, A., A. Arnqvist, M. Hammar, S. Sukupolvi, and S. Normark. 1993. The RpoS sigma factor relieves H-NS-mediated transcriptional repression of csgA, the subunit gene of fibronectin-binding curli in Escherichia coli. Mol. Microbiol. 7:523-536. [DOI] [PubMed] [Google Scholar]
  • 96.Olsen, A., A. Jonsson, and S. Normark. 1989. Fibronectin binding mediated by a novel class of surface organelles on Escherichia coli. Nature 338:652-655. [DOI] [PubMed] [Google Scholar]
  • 97.Oshima, T., H. Aiba, Y. Masuda, S. Kanaya, M. Sugiura, B. L. Wanner, H. Mori, and T. Mizuno. 2002. Transcriptome analysis of all two-component regulatory system mutants of Escherichia coli K-12. Mol. Microbiol. 46:281-291. [DOI] [PubMed] [Google Scholar]
  • 98.Ostrow, K. S., T. J. Silhavy, and S. Garrett. 1986. cis-Acting sites required for osmoregulation of ompF expression in Escherichia coli K-12. J. Bacteriol. 168:1165-1171. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 99.Park, K., S. Choi, M. Ko, and C. Park. 2001. Novel sigmaF-dependent genes of Escherichia coli found using a specified promoter consensus. FEMS Microbiol. Lett. 202:243-250. [DOI] [PubMed] [Google Scholar]
  • 100.Parkinson, J. S. 1993. Signal transduction schemes of bacteria. Cell 73:857-871. [DOI] [PubMed] [Google Scholar]
  • 101.Paul, R., S. Weiser, N. C. Amiot, C. Chan, T. Schirmer, B. Giese, and U. Jenal. 2004. Cell cycle-dependent dynamic localization of a bacterial response regulator with a novel di-guanylate cyclase output domain. Genes Dev. 18:715-727. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 102.Pei, J., and N. V. Grishin. 2001. GGDEF domain is homologous to adenylyl cyclase. Proteins 42:210-216. [DOI] [PubMed] [Google Scholar]
  • 103.Perry, R. D., A. G. Bobrov, O. Kirillina, H. A. Jones, L. Pedersen, J. Abney, and J. D. Fetherston. 2004. Temperature regulation of the hemin storage (Hms+) phenotype of Yersinia pestis is posttranscriptional. J. Bacteriol. 186:1638-1647. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 104.Polen, T., D. Rittmann, V. F. Wendisch, and H. Sahm. 2003. DNA microarray analyses of the long-term adaptive response of Escherichia coli to acetate and propionate. Appl. Environ. Microbiol. 69:1759-1774. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 105.Postma, P. W., J. W. Lengeler, and G. R. Jacobson. 1996. Phosphoenolpyruvate:carbohydrate phosphotransferase systems, p. 1149-1174. In F. C. Neidhardt, R. Curtiss III, J. L. Ingraham, E. C. C. Lin, K. B. Low, B. Magasanik, W. S. Reznikoff, M. Riley, M. Schaechter, and H. E. Umbarger (ed.), Escherichia coli and Salmonella: cellular and molecular biology. ASM Press, Washington, DC.
  • 106.Prigent-Combaret, C., E. Brombacher, O. Vidal, A. Ambert, P. Lejeune, P. Landini, and C. Dorel. 2001. Complex regulatory network controls initial adhesion and biofilm formation in Escherichia coli via regulation of the csgD gene. J. Bacteriol. 183:7213-7223. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 107.Pristovsek, P., K. Sengupta, F. Lohr, B. Schafer, M. W. von Trebra, H. Ruterjans, and F. Bernhard. 2003. Structural analysis of the DNA-binding domain of the Erwinia amylovora RcsB protein and its interaction with the RcsAB box. J. Biol. Chem. 278:17752-17759. [DOI] [PubMed] [Google Scholar]
  • 108.Prüß, B. M. 2000. FlhD, a transcriptional regulator in bacteria. Recent Res. Dev. Microbiol. 4:31-42. [Google Scholar]
  • 109.Prüß, B. M. 1998. Acetyl phosphate and the phosphorylation of OmpR are involved in the regulation of the cell division rate in Escherichia coli. Arch. Microbiol. 170:141-146. [DOI] [PubMed] [Google Scholar]
  • 110.Prüß, B. M., J. W. Campbell, T. K. Van Dyk, C. Zhu, Y. Kogan, and P. Matsumura. 2003. FlhD/FlhC is a regulator of anaerobic respiration and the Entner-Doudoroff pathway through induction of the methyl-accepting chemotaxis protein Aer. J. Bacteriol. 185:534-543. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 111.Prüß, B. M., X. Liu, W. Hendrickson, and P. Matsumura. 2001. FlhD/FlhC-regulated promoters analyzed by gene array and lacZ gene fusions. FEMS Microbiol. Lett. 197:91-97. [DOI] [PubMed] [Google Scholar]
  • 112.Prüß, B. M., D. Markovic, and P. Matsumura. 1997. The Escherichia coli flagellar transcriptional activator flhD regulates cell division through induction of the acid response gene cadA. J. Bacteriol. 179:3818-3821. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 113.Prüß, B. M., and A. J. Wolfe. 1994. Regulation of acetyl phosphate synthesis and degradation, and the control of flagellar expression in Escherichia coli. Mol. Microbiol. 12:973-984. [DOI] [PubMed] [Google Scholar]
  • 114.Romling, U. 2002. Molecular biology of cellulose production in bacteria. Res. Microbiol. 153:205-212. [DOI] [PubMed] [Google Scholar]
  • 115.Romling, U., Z. Bian, M. Hammar, W. D. Sierralta, and S. Normark. 1998. Curli fibers are highly conserved between Salmonella typhimurium and Escherichia coli with respect to operon structure and regulation. J. Bacteriol. 180:722-731. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 116.Romling, U., M. Gomelsky, and M. Y. Galperin. 2005. C-di-GMP: the dawning of a novel bacterial signalling system. Mol. Microbiol. 57:629-639. [DOI] [PubMed] [Google Scholar]
  • 117.Romling, U., W. D. Sierralta, K. Eriksson, and S. Normark. 1998. Multicellular and aggregative behaviour of Salmonella typhimurium strains is controlled by mutations in the agfD promoter. Mol. Microbiol. 28:249-264. [DOI] [PubMed] [Google Scholar]
  • 118.Rose, I. A., M. Grunberg-Manago, S. R. Korey, and S. Ochoa. 1954. Enzymatic phosphorylation of acetate. J. Biol. Chem. 211:737-756. [PubMed] [Google Scholar]
  • 119.Ross, P., R. Mayer, and M. Benziman. 1991. Cellulose biosynthesis and function in bacteria. Microbiol. Rev. 55:35-58. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 120.Ryjenkov, D. A., M. Tarutina, O. V. Moskvin, and M. Gomelsky. 2005. Cyclic diguanylate is a ubiquitous signaling molecule in bacteria: insights into biochemistry of the GGDEF protein domain. J. Bacteriol. 187:1792-1798. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 121.Sauer, K., A. K. Camper, G. D. Ehrlich, J. W. Costerton, and D. G. Davies. 2002. Pseudomonas aeruginosa displays multiple phenotypes during development as a biofilm. J. Bacteriol. 184:1140-1154. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 122.Schembri, M. A., K. Kjaergaard, and P. Klemm. 2003. Global gene expression in Escherichia coli biofilms. Mol. Microbiol. 48:253-267. [DOI] [PubMed] [Google Scholar]
  • 123.Schembri, M. A., D. W. Ussery, C. Workman, H. Hasman, and P. Klemm. 2002. DNA microarray analysis of fim mutations in Escherichia coli. Mol. Genet. Genomics 267:721-729. [DOI] [PubMed] [Google Scholar]
  • 124.Schwan, W. R., M. T. Beck, S. J. Hultgren, J. Pinkner, N. L. Woolever, and T. Larson. 2005. Down-regulation of the kps region 1 capsular assembly operon following attachment of Escherichia coli type 1 fimbriae to d-mannose receptors. Infect. Immun. 73:1226-1231. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 125.Shi, W., M. Bogdanov, W. Dowhan, and D. R. Zusman. 1993. The pss and psd genes are required for motility and chemotaxis in Escherichia coli. J. Bacteriol. 175:7711-7714. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 126.Shi, W., C. Li, C. J. Louise, and J. Adler. 1993. Mechanism of adverse conditions causing lack of flagella in Escherichia coli. J. Bacteriol. 175:2236-2240. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 127.Shi, W., Y. Zhou, J. Wild, J. Adler, and C. A. Gross. 1992. DnaK, DnaJ, and GrpE are required for flagellum synthesis in Escherichia coli. J. Bacteriol. 174:6256-6263. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 128.Shin, S., and C. Park. 1995. Modulation of flagellar expression in Escherichia coli by acetyl phosphate and the osmoregulator OmpR. J. Bacteriol. 177:4696-4702. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 129.Silverman, M., and M. Simon. 1973. Genetic analysis of bacteriophage Mu-induced flagellar mutants in Escherichia coli. J. Bacteriol. 116:114-122. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 130.Silverman, M., and M. Simon. 1973. Genetic analysis of flagellar mutants in Escherichia coli. J. Bacteriol. 113:105-113. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 131.Simm, R., M. Morr, A. Kader, M. Nimtz, and U. Romling. 2004. GGDEF and EAL domains inversely regulate cyclic di-GMP levels and transition from sessility to motility. Mol. Microbiol. 53:1123-1134. [DOI] [PubMed] [Google Scholar]
  • 132.Soutourina, O., A. Kolb, E. Krin, C. Laurent-Winter, S. Rimsky, A. Danchin, and P. Bertin. 1999. Multiple control of flagellum biosynthesis in Escherichia coli: role of H-NS protein and the cyclic AMP-catabolite activator protein complex in transcription of the flhDC master operon. J. Bacteriol. 181:7500-7508. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 133.Soutourina, O. A., E. Krin, C. Laurent-Winter, F. Hommais, A. Danchin, and P. N. Bertin. 2002. Regulation of bacterial motility in response to low pH in Escherichia coli: the role of H-NS protein. Microbiology 148:1543-1551. [DOI] [PubMed] [Google Scholar]
  • 134.Sperandio, V., A. G. Torres, J. A. Giron, and J. B. Kaper. 2001. Quorum sensing is a global regulatory mechanism in enterohemorrhagic Escherichia coli O157:H7. J. Bacteriol. 183:5187-5197. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 135.Sperandio, V., A. G. Torres, and J. B. Kaper. 2002. Quorum sensing Escherichia coli regulators B and C (QseBC): a novel two-component regulatory system involved in the regulation of flagella and motility by quorum sensing in E. coli. Mol. Microbiol. 43:809-821. [DOI] [PubMed] [Google Scholar]
  • 136.Stafford, G. P., T. Ogi, and C. Hughes. 2005. Binding and transcriptional activation of non-flagellar genes by the Escherichia coli flagellar master regulator FlhD2C2. Microbiology 151:1779-1788. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 137.Stoodley, P., K. Sauer, D. G. Davies, and J. W. Costerton. 2002. Biofilms as complex differentiated communities. Annu. Rev. Microbiol. 56:187-209. [DOI] [PubMed] [Google Scholar]
  • 138.Stout, V. 1994. Regulation of capsule synthesis includes interactions of the RcsC/RcsB regulatory pair. Res. Microbiol. 145:389-392. [DOI] [PubMed] [Google Scholar]
  • 139.Stout, V., A. Torres-Cabassa, M. R. Maurizi, D. Gutnick, and S. Gottesman. 1991. RcsA, an unstable positive regulator of capsular polysaccharide synthesis. J. Bacteriol. 173:1738-1747. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 140.Sturny, R., K. Cam, C. Gutierrez, and A. Conter. 2003. NhaR and RcsB independently regulate the osmCp1 promoter of Escherichia coli at overlapping regulatory sites. J. Bacteriol. 185:4298-4304. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 141.Sukupolvi, S., A. Edelstein, M. Rhen, S. J. Normark, and J. D. Pfeifer. 1997. Development of a murine model of chronic Salmonella infection. Infect. Immun. 65:838-842. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 142.Suzuki, K., X. Wang, T. Weilbacher, A. K. Pernestig, O. Melefors, D. Georgellis, P. Babitzke, and T. Romeo. 2002. Regulatory circuitry of the CsrA/CsrB and BarA/UvrY systems of Escherichia coli. J. Bacteriol. 184:5130-5140. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 143.Takeda, S., Y. Fujisawa, M. Matsubara, H. Aiba, and T. Mizuno. 2001. A novel feature of the multistep phosphorelay in Escherichia coli: a revised model of the RcsC → YojN → RcsB signalling pathway implicated in capsular synthesis and swarming behaviour. Mol. Microbiol. 40:440-450. [DOI] [PubMed] [Google Scholar]
  • 144.Tal, R., H. C. Wong, R. Calhoon, D. Gelfand, A. L. Fear, G. Volman, R. Mayer, P. Ross, D. Amikam, H. Weinhouse, A. Cohen, S. Sapir, P. Ohana, and M. Benziman. 1998. Three cdg operons control cellular turnover of cyclic di-GMP in Acetobacter xylinum: genetic organization and occurrence of conserved domains in isoenzymes. J. Bacteriol. 180:4416-4425. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 145.Tamayo, R., A. D. Tischler, and A. Camilli. 2005. The EAL domain protein VieA is a cyclic diguanylate phosphodiesterase. J. Biol. Chem. 280:33324-33330. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 146.Tomoyasu, T., T. Ohkishi, Y. Ukyo, A. Tokumitsu, A. Takaya, M. Suzuki, K. Sekiya, H. Matsui, K. Kutsukake, and T. Yamamoto. 2002. The ClpXP ATP-dependent protease regulates flagellum synthesis in Salmonella enterica serovar Typhimurium. J. Bacteriol. 184:645-653. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 147.Tomoyasu, T., A. Takaya, E. Isogai, and T. Yamamoto. 2003. Turnover of FlhD and FlhC, master regulator proteins for Salmonella flagellum biogenesis, by the ATP-dependent ClpXP protease. Mol. Microbiol. 48:443-452. [DOI] [PubMed] [Google Scholar]
  • 148.Tsung, K., R. E. Brissette, and M. Inouye. 1989. Identification of the DNA-binding domain of the OmpR protein required for transcriptional activation of the ompF and ompC genes of Escherichia coli by in vivo DNA footprinting. J. Biol. Chem. 264:10104-10109. [PubMed] [Google Scholar]
  • 149.Van Houdt, R., and C. W. Michiels. 2005. Role of bacterial cell surface structures in Escherichia coli biofilm formation. Res. Microbiol. 156:626-633. [DOI] [PubMed] [Google Scholar]
  • 150.Vidal, O., R. Longin, C. Prigent-Combaret, C. Dorel, M. Hooreman, and P. Lejeune. 1998. Isolation of an Escherichia coli K-12 mutant strain able to form biofilms on inert surfaces: involvement of a new ompR allele that increases curli expression. J. Bacteriol. 180:2442-2449. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 151.Wanner, B. L. 1993. Gene regulation by phosphate in enteric bacteria. J. Cell Biochem. 51:47-54. [DOI] [PubMed] [Google Scholar]
  • 152.Wehland, M., and F. Bernhard. 2000. The RcsAB box. Characterization of a new operator essential for the regulation of exopolysaccharide biosynthesis in enteric bacteria. J. Biol. Chem. 275:7013-7020. [DOI] [PubMed] [Google Scholar]
  • 153.Wehland, M., C. Kiecker, D. L. Coplin, O. Kelm, W. Saenger, and F. Bernhard. 1999. Identification of an RcsA/RcsB recognition motif in the promoters of exopolysaccharide biosynthetic operons from Erwinia amylovora and Pantoea stewartii subspecies stewartii. J. Biol. Chem. 274:3300-3307. [DOI] [PubMed] [Google Scholar]
  • 154.Wei, B. L., A. M. Brun-Zinkernagel, J. W. Simecka, B. M. Prüß, P. Babitzke, and T. Romeo. 2001. Positive regulation of motility and flhDC expression by the RNA-binding protein CsrA of Escherichia coli. Mol. Microbiol. 40:245-256. [DOI] [PubMed] [Google Scholar]
  • 155.Weinhouse, H., S. Sapir, D. Amikam, Y. Shilo, G. Volman, P. Ohana, and M. Benziman. 1997. c-di-GMP-binding protein, a new factor regulating cellulose synthesis in Acetobacter xylinum. FEBS Lett. 416:207-211. [DOI] [PubMed] [Google Scholar]
  • 156.West, A. H., and A. M. Stock. 2001. Histidine kinases and response regulator proteins in two-component signaling systems. Trends Biochem. Sci. 26:369-376. [DOI] [PubMed] [Google Scholar]
  • 157.Wolfe, A. J. 2005. The acetate switch. Microbiol. Mol. Biol. Rev. 69:12-50. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 158.Wolfe, A. J., D. E. Chang, J. D. Walker, J. E. Seitz-Partridge, M. D. Vidaurri, C. F. Lange, B. M. Prüß, M. C. Henk, J. C. Larkin, and T. Conway. 2003. Evidence that acetyl phosphate functions as a global signal during biofilm development. Mol. Microbiol. 48:977-988. [DOI] [PubMed] [Google Scholar]
  • 159.Xia, Y., D. Gally, K. Forsman-Semb, and B. E. Uhlin. 2000. Regulatory cross-talk between adhesin operons in Escherichia coli: inhibition of type 1 fimbriae expression by the PapB protein. EMBO J. 19:1450-1457. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 160.Yamamoto, K., R. Nagura, H. Tanabe, N. Fujita, A. Ishihama, and R. Utsumi. 2000. Negative regulation of the bolA1p of Escherichia coli K-12 by the transcription factor OmpR for osmolarity response genes. FEMS Microbiol. Lett. 186:257-262. [DOI] [PubMed] [Google Scholar]
  • 161.Yokota, T., and J. S. Gots. 1970. Requirement of adenosine 3′,5′-cyclic phosphate for flagella formation in Escherichia coli and Salmonella typhimurium. J. Bacteriol. 103:513-516. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 162.Zheng, D., C. Constantinidou, J. L. Hobman, and S. D. Minchin. 2004. Identification of the CRP regulon using in vitro and in vivo transcriptional profiling. Nucleic Acids Res. 32:5874-5893. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 163.Zhou, L., X. H. Lei, B. R. Bochner, and B. L. Wanner. 2003. Phenotype microarray analysis of Escherichia coli K-12 mutants with deletions of all two-component systems. J. Bacteriol. 185:4956-4972. [DOI] [PMC free article] [PubMed] [Google Scholar]

Associated Data

This section collects any data citations, data availability statements, or supplementary materials included in this article.

Supplementary Materials

[Supplemental material]

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

RESOURCES