Skip to main content
Journal of Bacteriology logoLink to Journal of Bacteriology
. 2005 Jan;187(1):304–319. doi: 10.1128/JB.187.1.304-319.2005

pH Regulates Genes for Flagellar Motility, Catabolism, and Oxidative Stress in Escherichia coli K-12

Lisa M Maurer 1, Elizabeth Yohannes 1, Sandra S Bondurant 2, Michael Radmacher 1, Joan L Slonczewski 1,*
PMCID: PMC538838  PMID: 15601715

Abstract

Gene expression profiles of Escherichia coli K-12 W3110 were compared as a function of steady-state external pH. Cultures were grown to an optical density at 600 nm of 0.3 in potassium-modified Luria-Bertani medium buffered at pH 5.0, 7.0, and 8.7. For each of the three pH conditions, cDNA from RNA of five independent cultures was hybridized to Affymetrix E. coli arrays. Analysis of variance with an α level of 0.001 resulted in 98% power to detect genes showing a twofold difference in expression. Normalized expression indices were calculated for each gene and intergenic region (IG). Differential expression among the three pH classes was observed for 763 genes and 353 IGs. Hierarchical clustering yielded six well-defined clusters of pH profiles, designated Acid High (highest expression at pH 5.0), Acid Low (lowest expression at pH 5.0), Base High (highest at pH 8.7), Base Low (lowest at pH 8.7), Neutral High (highest at pH 7.0, lower in acid or base), and Neutral Low (lowest at pH 7.0, higher at both pH extremes). Flagellar and chemotaxis genes were repressed at pH 8.7 (Base Low cluster), where the cell's transmembrane proton potential is diminished by the maintenance of an inverted pH gradient. High pH also repressed the proton pumps cytochrome o (cyo) and NADH dehydrogenases I and II. By contrast, the proton-importing ATP synthase F1Fo and the microaerophilic cytochrome d (cyd), which minimizes proton export, were induced at pH 8.7. These observations are consistent with a model in which high pH represses synthesis of flagella, which expend proton motive force, while stepping up electron transport and ATPase components that keep protons inside the cell. Acid-induced genes, on the other hand, were coinduced by conditions associated with increased metabolic rate, such as oxidative stress. All six pH-dependent clusters included envelope and periplasmic proteins, which directly experience external pH. Overall, this study showed that (i) low pH accelerates acid consumption and proton export, while coinducing oxidative stress and heat shock regulons; (ii) high pH accelerates proton import, while repressing the energy-expensive flagellar and chemotaxis regulons; and (iii) pH differentially regulates a large number of periplasmic and envelope proteins.


Escherichia coli and related enteric bacteria respond to a wide range of pH stresses by regulating gene expression (for reviews see references 21 and 68) and protein profiles (73, 82). Enteric bacteria encounter a wide range of external pHs in their natural habitat, the human digestive tract (17). Colonization of the intestine requires transient survival through the stomach at pH 1 to 2 (fasting) or 2 to 7 (transiently, during feeding) (18), as well as exposure to pancreatic secretions at pH 10 (25) followed by growth and persistence at a range of external pHs of 5 to 8 (20). Growth at a pH substantially higher or lower than the cytoplasmic pH 7.6 induces protective responses with two fundamental aims: to maintain internal pH homeostasis and to prepare the cell to survive future exposure to more extreme pH conditions (below pH 5 or above pH 9) that no longer permit growth (11, 41, 70).

The effects of pH on enteric bacteria contribute to disease. Low pH enhances expression of numerous virulence factors, such as the ToxR-ToxT virulence regulon in Vibrio cholerae (7), the phoP-phoQ regulon of Salmonella enterica (6), and the pH 6 antigen of Yersinia pestis (50). Acid stress contributes to food preservation; many food preservatives are membrane-permeant acids whose uptake is enhanced by acid (60), and acid interacts in complex ways with both temperature and organic food preservatives (65).

While growth in acid challenges pH homeostasis, the pH difference across the inner cell membrane (ΔpH) nevertheless contributes cell energy in the form of proton potential or proton motive force (Δp). The proton potential powers motility, ATP synthesis, and catabolite transport (for a review see reference 29). But low pH also amplifies the uptake of membrane-permeant acids that dissipate the proton potential (59). Thus, we expect low pH to induce a combination of positive and negative responses.

Much of bacterial catabolism affects pH, and in E. coli a growing number of catabolic enzymes and catabolite transporters are known to be regulated by pH (21, 73). Sugar fermentation initially generates short-chain acids that are excreted but accumulate and reenter the cytoplasm, causing acidification. Thus, it is not surprising that sugar transporters such as OmpF and the maltose regulon are down-regulated at low pH (13). Consumption of acids by the tricarboxylic acid (TCA) cycle causes alkalinization, a common result of growth to stationary phase in tryptone-based media (66, 73). Catabolism of amino acids by decarboxylases generates alkaline amines, which help the cell counteract external acidification, for example, the lysine and arginine decarboxylases (4, 27, 45, 47, 71). High pH, however, induces deaminases that generate acids, such as tryptophan deaminase (tnaAB) and serine deaminase (sda) (9, 73, 82).

A complicated case is that of the glutamic acid decarboxylase genes gadA and gadBC (12, 44). The gad system enables cells to survive extreme acid (77), but its expression is induced mainly at high pH, or in Luria-Bertani medium grown to stationary phase, where pH naturally increases (73, 82). An alternative role of gad, particularly under anaerobiosis, may be to channel its product γ-aminobutyric acid into fermentation acids.

Even mild acid (pH 6 to 7) greatly amplifies the uptake of membrane-permeant weak acids such as acetate. Permeant acids pass through the bacterial membrane and dissociate in the cytoplasm, causing accumulation of anions and depression of internal pH (34, 56). Acetate concentrations rise as cell density increases, and acetate induces a large number of genes and proteins (3, 35). Growth inhibition occurs as a result of both lower internal pH and the differential ability of anions to inhibit metabolism (60). The effect of permeant acids is critical in the human colon, where the concentration of short-chain fatty acids totals approximately 100 mM (15).

While numerous responses to pH stress are known, the mechanisms by which E. coli maintains its internal pH at 7.6 remain poorly understood. The electron transport chain pumps protons outside the cell, and the H+-ATPase either exports or imports protons, but mutants in these components maintain pH homeostasis. There is evidence that potassium exchange contributes to pH homeostasis in external acid (5, 10, 52, 80), but the precise mechanisms remain unclear. At high pH, the electrical potential (Δψ) is diminished in order to compensate for the inverted ΔpH. The sodium-proton antiporterNhaA contributes to internal pH maintenance under sodium stress (24, 75). High pH also induces major stress systems such as heat shock response (1, 28, 74), the SOS regulon (63), and the CpxP envelope stress response (16).

At more extreme pH values, well below the growth range (as low as pH 1.5 for clinical isolates) E. coli can retain viability for many hours, a phenomenon termed acid survival or acid resistance. Acid resistance is enhanced by many genes induced during growth at the acid end of the pH range (pH 5) or growth to stationary phase. Acid-induced acid resistance factors include periplasmic chaperones such as the hdeA product (23), envelope proteins such as OsmY, and redox modulators such as Tpx (73, 78). A complex acid resistance regulon including the gad system is regulated by transcription factors GadX-GadW and EvgA-YdeO, as well as by RpoS, H-NS, and cyclic AMP (11, 12, 44, 79). E. coli also exhibits base resistance, the ability to survive at or above pH 10 (58, 70). Base resistance requires rpoS and components of the gad system (30).

Finally, pH may affect flagellar motility, although the present picture is unclear. According to one report, growth in acid represses flagellar genes and eliminates motility (72), whereas another group finds motility enhanced by acetate and propionate, which cause acid stress (53).

To investigate acid and base response, we used microarrays to compare E. coli gene expression at low, neutral, or high external pH. Past microarray studies of pH response have been limited by their absence of pH conditions above pH 7 (44, 78); their use of glucose minimal medium (78), in which many catabolic genes are repressed; and their focus on only a single acid resistance regulon (44). Our experimental design included both acid and base conditions, as well as pH 7.0. For each growth condition, five independent cultures were hybridized separately, a number of replicates that ensured detection of virtually all expression ratios of at least twofold. The coregulation of numerous genes within operons confirmed the biological relevance of our expression ratios. Our study revealed unexpected patterns of pH response and clarified the overlap of pH stress with other stress responses.

MATERIALS AND METHODS

Growth conditions.

E. coli K-12 strain W3110 (R. VanBogelen and F. Neidhardt) was grown overnight in unbuffered potassium-modified Luria broth (LBK) (10 g of tryptone/liter, 5 g of yeast extract/liter, 7.45 g of KCl/liter). For pH-controlled growth, media were buffered with 100 mM homopiperazine-N,N′-bis-2-(ethanesulfonic acid) (HOMOPIPES) (pKa, 4.55 and 8.12). The pH of the media were adjusted to 5.0, 7.0, or 8.7 with KOH solution to avoid extra sodium ions, which stress cells at high pH (24). To maximize aeration and maintain logarithmic growth, the overnight culture was diluted 1,000-fold into 12 ml of buffered medium in a 125-ml baffled flask and rotated at 240 rpm. Cultures were grown at 37°C to an optical density at 600 nm of 0.3. For all cultures, the pH was tested after growth to ensure that the values were maintained at ±0.2 pH unit of the pH of the original uninoculated medium.

To observe motility, we used E. coli K-12 strain RP437 and S. enterica serovar Typhimurium SJW1103 from a laboratory in which strains are maintained for motility (M. Macnab). Culture was spotted on tryptone-KCl soft-agar plates (0.35% Bacto Agar) and incubated at 37°C until cells swam out. Culture was picked from the leading edge of the swimming cells and inoculated into LBK for overnight growth. For quantitative assay of motility, 5 μl of culture was spotted in triplicate on plates containing tryptone-KCl with 100 mM sulfonate buffer of appropriate pKa (73). After growth for 8 h, the diameter of motile cell growth was measured.

RNA isolation.

Bacterial RNA was isolated using the Qiagen RNeasy kit with on-column DNA digestion (Qiagen), with additional DNA removal with Ambion DNase. To perform this additional DNase digestion, RNA was precipitated and redissolved in 85 μl of nuclease-free water. We then added 10 μl of 10× DNase I buffer and 5 μl of (1-U/μl) DNase I (Ambion). The DNase reaction mixture was incubated at 37°C for 30 min and then chilled on ice. A second RNeasy column purification was performed.

cDNA preparation and array hybridization.

For microarrays, standard methods were used for cDNA synthesis, fragmentation, and end-terminus biotin labeling, based on Affymetrix protocols. Labeled cDNA was hybridized to E. coli Affymetrix Antisense Genome Arrays. Hybridized arrays were stained with streptavidin-phycoerythrin with the use of the Affymetrix Fluidic Station. After staining, arrays were scanned with a GC2500 scanner.

Statistical analysis of gene expression.

The experiment was designed so as to minimize both false-positive and false-negative results for expressed genes. Five full replicates (with respect to E. coli growth, RNA isolation, sample preparation, and array hybridization) were performed for each pH condition.

The median within-group variance in expression for all genes in the data set was 0.031 (or standard deviation, 0.175). To test for significant differences in expression between the pH classes, one-way analysis of variance (ANOVA) was performed at a significance level of 0.001; thus, of every thousand genes tested, only one false positive would be expected. For a gene with average within-group variability, our sample size provided statistical power of 98% to detect a twofold difference in gene expression among pH groups. That is, only 2% of genes that show a twofold difference in expression between any two pH groups would be missed (false negatives).

Model-based expression analysis with dChip software (40) was performed on the probe-level data from Affymetrix's DAT files. The model relates target RNA levels to the probe signals by a linear function that weights the significance of all oligonucleotide probes for each gene. The analysis includes normalization, which rescales data from different arrays so that comparisons can be made among arrays. Each array was normalized to a baseline array from a pH 7 culture, by using local regression on an invariant set of probes (62). Model-based expression indices were calculated for each gene on each array by using only the perfect match probes (61), and outlier detection was performed (39). Only probe sets that received an Affymetrix call of “present” on greater than 50% of the arrays were used in subsequent analyses. “Present” or “absent” calls use information from paired perfect-match and single-base-mismatch probes. Four thousand six hundred fifty probe sets passed this criterion.

For genes whose probe sets passed the 50% screen, one-way ANOVA was performed on the log2-transformed model-based expression indices, on a gene-by-gene basis. For each gene that displayed significant differences in expression among the classes, pairwise comparisons of pH classes were determined using Tukey's multiple comparisons procedure to control the familywise error rate for the t test.

Additional analyses were performed to explore categories of differential gene expression. Global relationships among arrays were visualized by performing a principal component analysis (81) on the expression data and plotting arrays in two-dimensional space corresponding to the first two principal components. The gene expression profiles of the arrays were visualized in two-dimensional Euclidian space, by using BRB ArrayTools software. In addition, categories of differential expression profiles across the pH classes were generated by a hierarchical cluster analysis of differentially expressed genes, based on the average linkage method (19) with BRB ArrayTools.

RESULTS

Growth range of pH.

To study the full range of pH response, we selected the widest pH range (pH 5.0 to 8.7) in which cultures maintained reasonable doubling times and approximately constant pH throughout growth. Culture media were adjusted to pH 5.0, 7.0, and 8.7. The doubling time for E. coli cultured at pH 5.0 and 8.7 was approximately 25 min and at pH 7.0 was 18 min. All cultures were grown to an optical density of 0.3 in order to facilitate at least five complete replications. The final pH of growth cultures was found to be within ±0.2 of the initial pH. The internal pH of the cytoplasm is approximately 7.6 (69); thus, growth at external pH 7.0 might induce some acid response.

Probe hybridization.

To determine differential gene expression, the log2 transforms of normalized model-based expression values of genes were compared. Of the 7,231 genes and intergenic regions (IGs) on the array, 4,650 loci were detected on more than half (eight or more) of the 15 arrays. These loci, constituting about 70% of the total array, were taken for further analysis.

Principal component analysis.

Global relationships among arrays were visualized by performing a principal component analysis (81) on the expression data (Fig. 1). Before dimensional reduction, each array existed in 4,650-dimensional space (one dimension for each of the 4,650 intensity values). The array comparisons were plotted in two-dimensional space, corresponding to the first and second principal components of variation. The first principal component for each array is the weighted linear combination of intensity values that shows maximum variation, whereas the second principal component is a weighted linear combination orthogonal to the first component that has maximum variance.

FIG. 1.

FIG. 1.

Principal component analysis. The gene expression profiles of the arrays were visualized in two-dimensional Euclidian space, by using BRB ArrayTools software as described under Materials and Methods. The first and second principal components are shown. pH 5.0, squares; pH 7.0, circles; pH 8.7, triangles.

The principal component analysis indicated that the microarrays from each of the three pH conditions appeared in distinct groups (Fig. 1). Within-class variability was small relative to variability among pH levels. The pH 8.7 arrays showed the greatest degree of separation, clustering into two groups based on the date on which the arrays were hybridized, but this difference was small compared to the differences between pH classes.

ANOVA for significance of expression profiles.

We compared gene expression among the three pH groups on a gene-by-gene basis using one-way ANOVA at a significance level of 0.001. The significance level indicates the probability of a false positive, and we therefore expect 0.001 × 4,650 = 4.65 false-positive genes (i.e., genes that are not truly differentially expressed but that appear in our differentially expressed list) in our full analysis. Of the 4,650 loci with eight or more “present” calls on arrays, 761 genes and 353 IGs showed a significant F value for differential expression among the three pH classes. Thus, about 17% of E. coli genes showed significant modulation of expression as a function of pH.

Cluster analysis.

As a first attempt at categorizing differentially expressed genes, we performed a hierarchical cluster analysis (19) of differentially expressed genes (Fig. 2). We used average linkage and one minus the centered Pearson correlation as the distance metric. At a correlation value of approximately 0.6, the dendrogram generated six clusters of gene expression profiles.

FIG. 2.

FIG. 2.

Cluster analysis of differentially expressed genes. The dendrogram was generated based on the average linkage method (19) with BRB ArrayTools. At a correlation of 0.6, six clusters of related gene expression were designated Acid High (AH), Acid Low (AL), Base High (BH), Base Low (BL), Neutral High (NH), and Neutral Low (NL).

Within each of the six clusters, the average profiles were determined for all the gene expression indices (log2 intensity values) across the three pH conditions (Fig. 3). The clusters were defined by their mean expression profiles across the three pH conditions. The Acid High cluster showed highest expression at pH 5.0, declining at pH 7.0 and 8.7. It included 160 genes and 49 IGs. Acid Low (113 genes, 57 IGs) showed approximately the reverse profile, with its lowest expression at pH 5.0, rising at pH 7.0 and 8.7. Base High (93 genes, 70 IGs) showed low expression at pH 5.0 and 7.0 and higher expression at pH 8.7, whereas Base Low (123 genes, 40 IGs) showed the reverse, higher expression at pH 5.0 and 7.0 than at pH 8.7. The Neutral High cluster (93 genes, 14 IGs) showed highest expression at pH 7.0 and lower expression at both pH extremes. The Neutral Low cluster (181 genes 123 IGs) showed the lowest expression at pH 7.0 and higher expression at both pH extremes, although the mean expression was substantially greater at pH 5.0 than at pH 8.7; a number of acid-induced genes fell in this category.

FIG. 3.

FIG. 3.

Cluster mean expression profiles. The mean expression profiles over pH are plotted for the six clusters defined in Fig. 2.

Table 1 lists the genes that fell into each cluster; details of description and Blattner open reading frame (ORF) numbers are available online in Table S1 in the supplemental material. In many cases, all or most of the ORFs in a given operon were induced in the same cluster; see, for example, the atp operon (Base High cluster) and the flg and fli operons (Base Low cluster).

TABLE 1.

Clusters of pH-dependent genes

graphic file with name zjb00105435400t1.jpg
a

Acid induced (68, 73, 82).

b

Data from five arrays from pH 5.0; no significant expression at pH 7.0 and 8.7.

c

Acetate induced (3, 35).

d

Extreme acid resistance (44, 78, 79).

e

Base induced (68, 73, 82).

f

Base induced (16).

Known acid-induced genes and acid resistance genes such as sucBC and hdeA (73) generally fell under Acid High, Base Low, or Neutral Low, a cluster whose mean expression indices were actually twofold higher in acid than in base (Fig. 3). These results are generally consistent with the cluster pH profiles and with the structure of the cluster dendrogram, in which the Acid High profile correlates most closely with the Neutral Low profile. Most known base-induced genes, such as alx (ygjT) (8, 73) and tnaA (9), fell under Base High or Acid Low.

For IGs, the cluster assignment and expression ratios are presented online in Table S2 in the supplemental material. Expression of an IG may result from a small regulatory RNA that lies between protein-encoding genes (2, 43), or it may indicate the tail end of mRNA containing pH-regulated genes. For example, the IGs upstream of tnaC (tnaA leader peptide) and downstream of tnaB both were repressed in acid, as are tnaA and tnaB.

Individual gene expression ratios.

For genes whose overall expression profile yielded a significant F value (one-way ANOVA), we used the Tukey procedure to determine ratios of average model-based expression indices from cultures at pH 5.0 versus pH 7.0, at pH 8.7 versus pH 7.0, and at pH 8.7 versus pH 5.0. The full list of individual log2 expression ratios for all analyzed genes is presented in Table S1 in the supplemental material and for IGs is presented in Table S2 in the supplemental material; for genes of particular interest grouped in functional categories, the data are presented in Tables 3 through 7. Expression ratios that are significant at α = 0.001 are shown in boldface.

TABLE 3.

Flagellar and chemotaxis genes

Gene Function Log2 pH ratioa
Classb
5/7 8.7/7 8.7/5
cheA Chemotaxis sensor kinase −1.125 3.199 −2.0474 BL
cheB Protein methylesterase −1.050 −1.578 −0.528 BL
cheR Chemotaxis MCPc methyltransferase −0.564 −1.013 −0.448 BL
cheW Chemotaxis signal transducer −1.336 −2.785 −1.449 BL
cheY Response regulator for chemotactic signal −1.089 −1.310 −0.221 BL
cheZ CheY-P phosphatase −1.505 −3.199 −1.694 BL
flgA Flagellar synthesis −0.261 −1.056 −0.795 BL
flgB Basal body rod subunit −0.192 −1.120 −0.928 BL
flgC Basal body rod subunit −0.257 −1.241 −0.984 BL
flgD Basal body rod modification 0.133 −1.241 −1.107 BL
flgE Hook subunit 0.272 −0.856 −1.128 BL
flgF Basal body rod subunit 0.109 −1.239 −1.348 BL
flgG Basal body rod major subunit 0.220 −1.116 −1.335 BL
flgI Basal body P-ring −0.011 −1.218 −1.207 BL
flgJ Flagellum-specific muramidase 0.030 −0.857 −0.886 BL
flgK Flagellar synthesis −0.211 −1.875 −1.664 BL
flgL Flagellar synthesis 0.111 −1.165 −1.276 BL
flgM Anti-sigma 28 (FliA); regulates FlhD −0.292 −1.424 −1.132 BL
flgN Flagellar synthesis −0.295 −1.567 −1.272 BL
flhA Flagellar export pore protein 0.333 −0.528 −0.861 AH
flhE Function unknown 0.492 −0.454 −0.946 AH
fliA Sigma 28; regulates class III flagellar genes −0.150 −1.127 −0.976 BL
fliC Flagellin subunit, H-antigen −1.165 −4.561 −3.396 BL
fliD Hook-associated protein −0.311 −1.953 −1.641 BL
fliE Flagellar synthesis; basal body component −0.478 −1.861 −1.203 BL
fliF Flagellar basal body M-ring −0.159 −1.216 −1.057 BL
fliG Motor switching and energizing 0.182 −1.313 −1.495 BL
fliH Negative regulator of FliI; flagellar assembly and export 0.389 −1.245 −1.634 BL
fliI Membrane ATPase, flagellar, axial subunit export 0.270 −1.238 1.508 BL
fliJ Flagellar biosynthesis −0.068 −1.404 −1.336 BL
fliK Hook filament junction 0.278 −1.186 −1.464 BL
fliL Rotational direction of flagella −0.102 −1.083 −0.981 BL
fliM Flagellar synthesis, motor switching and energizing 0.029 −1.053 −1.082 BL
fliN Flagellar switch 0.273 −0.875 −1.148 BL
fliO Flagellar synthesis 0.191 −0.921 −1.112 BL
fliP Flagellar synthesis 0.419 −0.669 −1.087 AH
fliQ Flagellar synthesis 0.463 −0.980 −1.444 AH
fliR Flagellar synthesis 0.924 −0.899 −1.822 AH
fliS Cytosolic chaperone inhibits premature FliC assembly −0.174 −1.947 −1.772 BL
fliT Flagellar synthesis −0.323 −1.021 −0.698 BL
fliY Cystine-binding protein, periplasmic; may regulate FliA (sigma 28) 0.233 −0.252 −0.484 AH
fliZ Not required for motility; may regulate FliA (sigma 28) 0.144 −1.165 −1.309 BL
motA Flagellar rotation −0.682 −1.847 −1.166 BL
motB Flagellar rotation −1.241 −3.238 −1.997 BL
tap Dipeptide chemoreceptor −0.622 −1.089 −0.466 BL
tar Aspartate, maltose chemoreceptor −0.519 −2.213 −1.694 BL
tsr Serine chemoreceptor −1.061 −1.826 −0.765 BL
ycgR Suppresses hns motility defect −0.531 −0.783 −0.252 BL
yhjH Suppresses hns motility defect −0.972 −2.522 −1.550 BL
a

Boldface for ratios indicates significance (α = 0.001).

b

BL, Base Low; AH, Acid High.

c

MCP, methyl-accepting chemotaxis protein.

The genes most strongly regulated by pH are summarized in Table 2. These genes each showed an expression ratio of at least fourfold (log2 = 2) between two of the pH classes. Note that the two genes most strongly induced in acid are ORFs with no known function, yhcN and yagU. Other acid-induced genes include those for catabolic enzymes in pathways that consume acids, such as sdhCD (succinate dehydrogenase). Genes repressed at high pH include several members of the flagellar regulon, including the main flagellar subunit fliC (for a review see reference 42).

TABLE 2.

Strongest pH-dependent expression ratios (fourfold or higher)

pH dependence and pH ratio Gene Log2 ratio pH dependence and pH ratio Gene Log2 ratio
Acid induced
    5.0/7.0 yagU 3.220
yhcN 3.064
sdhC 2.728
lysP 2.662
sdhD 2.349
cfa 2.075
nemA 2.060
    5.0/8.7 yhcN 4.199
yagU 3.962
fliC 3.396
fimA 2.579
cfa 2.555
gltB 2.271
ydiY 2.193
ycdN 2.147
yncD 2.117
mqo 2.075
dhaH 2.074
cheA 2.074
motB 1.997
    7.0/8.7 fliC 4.561
malM 3.780
malK 3.748
lamB 3.735
malP 3.373
motB 3.238
cheA 3.199
cheZ 3.199
flxA 3.007
malE 2.933
malQ 2.883
cheW 2.785
ibpB 2.618
yhjH 2.522
htpG 2.439
deoC 2.267
pstS 2.262
dnaK 2.249
tar 2.213
yjdA 2.208
dnaJ 2.155
yrfG 2.055
yheL 2.031
deoA 2.003
Base induced
    8.7/7.0 yifO 2.769
ymgD 2.221
    8.7/5.0 tnaC 5.517
cpxP 4.234
tnaA 4.028
nmpC 3.961
treB 3.895
yjiY 3.665
treC 3.233
b3913 3.176
yifO 3.095
borD 3.088
tnaB 2.993
ycfS 2.820
yghJ 2.762
ymgD 2.433
yccA 2.378
yfiA 2.364
yebE 2.343
yjiX 2.292
nrdD 2.182
dniR 2.081
alx 2.009
mutL 2.000
    7.0/5.0 tnaC 5.026
lamB 4.881
malK 4.790
malM 4.643
yjiY 4.359
nmpC 4.012
malP 4.000
tnaA 3.805
malE 3.425
borD 3.378
malQ 3.322
cpxP 3.232
treB 3.156
yghJ 3.081
yjiX 2.959
fecB 2.842
ompF 2.834
treC 2.613
fecA 2.608
pstS 2.479
b3913 2.177
fecE 2.098

The genes most strongly induced at high pH included tnaC, encoding the tryptophanase leader peptide (26), as well as tnaA (tryptophanase) and the Trp transporter gene tnaB, with its leader peptide gene tnaC. Previously in proteomic gels, we found tryptophanase to be the most highly expressed protein observed at high pH (9). The alkali-inducible protease gene cpxP (16) was also strongly induced. Members of the maltose transport regulon (malEKM) were strongly repressed by acid, consistent with previous reports (31, 73). But proteins strongly induced by base also included those from genes of unknown function, such as yifO and ymgD.

Flagellar and chemotaxis regulons.

Motility in E. coli is governed by the flagellar chemotaxis regulon including 50 components in 19 operons, governed by the major regulators FlhC and FlhD (42, 76). The expression of the regulatory operon flhCD is controlled by numerous environmental response systems, such as adenylate cyclase (37), RcsCDB (22), and ClpXP (76).

Nearly all the genes of the flagellar regulons (47 genes) were repressed at high pH (Table 2). Forty-one genes fell in the Base Low cluster, which means that the bulk of significant expression difference occurred between pH 7.0 and 8.7. (The other six genes were Acid High.) These genes were among the most strongly base-repressed genes in the arrays (Table 2); for instance, fliC, encoding the flagellin monomer, had the lowest pH 8.7/pH 7.0 ratio observed, down-regulated about 20-fold (Table 3). Some of the che and mot genes showed a relatively small degree of repression in acid compared to that at pH 7.0 but overall were repressed at high pH.

The major regulator operon flhCD, however, showed no effect of pH. Thus, either the flhCD probes failed to show up in our arrays or pH may affect expression posttranscriptionally.

Motility assays.

The effect of pH on motility was tested by spotting motile cultures of E. coli K-12 RP437 and S. enterica serovar Typhimurium SJW1103 on motility agar buffered at a range of pH values (Fig. 4). Both species showed a steady decline of motility as pH increased. The decline was particularly steep between pH 7.5 and 8.7.

FIG. 4.

FIG. 4.

Swimming distance as a function of pH. E. coli K-12 RP437 and S. enterica serovar Typhimurium SJW1103 were spotted on soft-agar plates as described under Materials and Methods. Error bars represent standard errors of the means (n = 3); in most cases their size was smaller than the symbol.

Catabolism and proton transport.

Several enzymes for catabolism of sugars and amino acids show a pH dependence that minimizes acid production at low external pH or maximizes acids at high pH (68, 73). Our microarrays revealed many new components, showing the broad scope of pH regulation of catabolism (Table 4).

TABLE 4.

Catabolism and respiration

Group Gene Function Log2 pH ratioa
Classb
5/7 8.7/7 8.7/5
Sugar catabolism and TCA cycle aceE Pyruvate dehydrogenase 1.928 0.527 −1.401 NL
aceF Pyruvate dehydrogenase dihydrolipoamide acetyltransferase 1.802 0.402 1.401 NL
aceK Isocitrate dehydrogenase kinase/phosphatase 0.667 0.038 −0.629 AH
acnA Aconitase A, stationary phase induced 0.769 −0.143 −0.912 AH
acnB Aconitase B; 2-methylaconitate hydratase 1.036 −0.041 −1.077 NL
acrR Regulator for acrA and acrB −0.873 0.164 1.037 AL
dcuR (yjdG) Fumurate respiration regulator (anaerobic) −0.220 0.055 0.275 AL
dcuS (yjdH) Fumurate respiration regulator (anaerobic) −0.324 0.065 0.390 AL
dhaK Dihydroxyacetone kinase, subunit I 1.041 −0.379 −1.420 AH
dhaL Dihydroxyacetone kinase, subunit II 0.742 −0.385 −1.127 AH
dld d-Lactate dehydrogenase 0.919 0.566 −0.353 NL
eno Enolase; RNA degradosome 0.299 0.052 −0.247 AH
fucI l-Fucose isomerase −0.693 −0.173 0.520 AL
fucK l-Fuculose kinase −1.259 −0.519 0.740 NH
fucR Positive regulator, fuc operon −0.046 0.495 0.541 BH
galF Putative regulator of galU NL
galK Galactokinase −0.616 −0.562 0.053 BL
galM Galactose mutarotase; aldose-1-epimerase −0.619 −0.647 −0.028
gapA Glyceraldehyde 3-P dehydrogenase A 0.093 −0.752 −0.844 BL
gatA Galactitol-specific enzyme IIA of PTSd 0.929 −0.626 −1.555 AH
gatB Galactitol-specific enzyme IIB of PTS 0.632 −0.895 −1.528 AH
gatC Galactitol-specific enzyme IIC of PTS 0.973 −0.949 −1.922 AH
gatD Galactitol-1-phosphate dehydrogenase 1.037 −0.893 −1.930 AH
gatY d-Tagatose-1,6-bisphosphate aldolase, class II 0.846 −0.483 −1.329 AH
gatZ Enhances GatY activity 0.774 −0.605 −1.380 AH
glpA Glycerol-3-phosphate dehydrogenase large subunit (anaerobic) −0.200 0.558 0.758 AL
glpB Glycerol-3-phosphate membrane anchor (anaerobic) −0.124 0.437 0.562 BH
glpC Glycerol-3-phosphate dehydrogenase (anaerobic) small subunit −0.169 0.469 0.638 BH
glpX Fructose 1,6-bisphosphatase −0.162 0.375 0.537 BH
gltA Citrate synthase 0.288 −0.559 −0.846 AH
gnd Gluconate-6-phosphate dehydrogenase 0.895 0.351 −0.544 NL
gntT High-affinity gluconate transport −0.613 −0.362 0.251 NH
gpsA Glycerol-3-phosphate dehydrogenase −0.492 −0.534 −0.041 NH
gpmA Phosphoglycerate mutase I 0.303 −0.118 −0.421 AH
icdA Isocitrate dehydrogenase 1.211 0.061 −1.272 AH
lldD l-Lactate dehydrogenase 0.613 −0.554 −1.167 AH
lldP l-Lactate permease; glycolate uptake 1.527 0.219 −1.307 NL
lpdA Lipoamide dehydrogenase; E3 component of pyruvate and 2-oxoglutarate dehydrogenase complexes 1.507 0.281 −1.226 NL
malE Maltose-binding protein, periplasmic −3.425 −2.933 0.491 NH
malF Maltose transport, inner membrane −1.016 −0.935 0.082 NH
malG Maltose transport, inner membrane subunit −1.577 −1.502 0.075 NH
malK Maltose transport, ATP-binding subunit −4.790 −3.748 1.041 NH
malM Periplasmic protein, mal regulon −4.643 −3.780 0.863 NH
malP Maltodextrin phosphorylase −4.000 −3.373 0.627 NH
malQ Amylomaltase −3.322 −2.883 0.439 NH
malT mal positive regulator −0.103 0.617 0.720 BH
pdhR Pyruvate dehydrogenase operon repressor 0.755 −0.109 −0.863 AH
pfkB 6-Phosphofructokinase-2 0.363 0.095 −0.268 NL
pflA Pyruvate formate lyase I activase 0.964 0.667 −0.297 NL
pflB Pyruvate formate lyase I (anaerobic) 0.812 0.578 −0.234 NL
pgi Glucose phosphate isomerase 0.654 0.177 −0.477 NL
pta Phosphotransacetylase 1.167 0.639 −0.528 NL
ptsG Glucose PTS enzyme IIBC 0.841 0.544 −0.287 NL
ptsH PTS system histidine phosphocarrier protein Hpr 0.369 0.564 0.195 NL
ptsI PTS system enzyme I 0.241 0.339 0.098 NL
ptsO NPr, N-regulated HPr-like protein 0.639 0.168 −0.470 NL
rpiA Ribose-5-phosphate isomerase A 0.180 0.364 0.184 BH
srlA Sorbitol-specific enzyme II of PTS −1.156 −1.720 −0.564 BL
srlB Sorbitol-specific enzyme III of PTS −0.908 −1.140 −0.232 BL
srlD Sorbitol-6-phosphate dehydrogenase −0.660 −1.334 −0.674 BL
srlE srl operon protein −1.013 −1.613 −0.599 BL
srlR srl regulator −0.151 0.804 −0.654 BL
sucA 2-Oxoglutarate dehydrogenase, E1 component 0.850 −0.851 −1.701 AH
sucB Dihydrolipoamide succinyltransferase component of 2-oxoglutarate dehydrogenase complex (E2) 0.566 −1.073 −1.638 AH
sucC Succinyl-CoAc synthase beta subunit 0.494 −1.019 −1.513 AH
sucD Succinyl-CoA synthase alpha subunit 0.526 −1.225 −1.751 AH
tpiA Triosephosphate isomerase 0.772 0.245 −0.526 NL
treB Trehalose-specific PTS enzyme II −3.156 0.739 3.895 AL
treC Trehalose-6-phosphate hydrolase −2.613 0.620 3.233 AL
Proton transport and electron transport chain atpA ATP synthase subunit alpha, F1 0.111 0.460 0.349 BH
atpB ATP synthase subunit a, F0 0.125 0.484 0.609 BH
atpC ATP synthase subunit epsilon, F1 0.266 1.005 0.739 BH
atpD ATP synthase subunit beta, F1 0.542 1.172 0.630 BH
atpE ATP synthase subunit c, F0 0.182 0.503 0.321 BH
atpF ATP synthase subunit b, F0 −0.055 0.321 0.377 BH
atpG ATP synthase subunit gamma, F1 0.372 0.632 0.260 BH
atpH ATP synthase subunit delta, F1 −0.172 0.186 0.358 AL
atpI ATP synthase subunit, F1F0-type proton-ATPase 0.017 0.499 0.482 BH
cydA Cytochrome d (bd-I) terminal oxidase subunit I (microaerobic) −0.268 0.872 1.140 BH
cydB Cytochrome d (bd-I) terminal oxidase subunit II (microaerobic) −0.097 0.777 0.874 BH
cydC Cysteine exporter to periplasm required for Cyd assembly 0.135 0.490 0.355 BH
cydD Cysteine exporter to periplasm required for cytochrome assembly 0.232 0.574 0.342 BH
cyoA Cytochrome o oxidase subunit II 1.160 0.217 −0.943 NL
cyoB Cytochrome o oxidase subunit I 0.812 −0.204 −1.016 NL
cyoC Cytochrome o oxidase subunit III 1.026 0.019 −1.007 NL
cyoD Cytochrome o oxidase subunit IV 0.829 −0.192 −1.021 NL
cyoE Cytochrome o oxidase subunit protoheme IX farnesyltransferase 1.094 0.225 −0.869 NL
fdoG Formate dehydrogenase-O, major selenopeptide subunit 0.353 −0.138 −0.491 AH
fdoH Formate dehydrogenase-O Fe-S subunit 0.003 −0.414 −0.416 BL
frdA Fumarate reductase flavoprotein subunit −0.305 0.046 0.351 AL
frdC Fumarate reductase membrane anchor polypeptide −0.378 −0.008 0.370 AL
fumA Fumarase A 0.595 −0.647 −1.242 AH
nfsA (mdaA) Nitroreductase A 0.561 −0.419 −0.980 AH
mdaB Probable nitroreductase or quinone reductase 0.444 −0.223 −0.667 AH
napC Cytochrome electron source for NapAB, membrane bound −0.797 −0.382 0.416 NH
ndh Respiratory NADH dehydrogenase II; NADH:ubiquinone oxidoreductase II 0.789 0.161 −0.628 NL
nuoC NADH:ubiquinone oxidoreductase subunit C 0.440 −0.084 −0.524 AH
nuoG NADH:ubiquinone oxidoreductase subunit G; NADH dehydrogenase I 0.525 0.103 −0.442 NL
nuoH NADH:ubiquinone oxidoreductase subunit H; NADH dehydrogenase I 0.564 0.185 −0.379 NL
nuoI NADH:ubiquinone oxidoreductase subunit I; NADH dehydrogenase I 0.792 0.310 −0.481 NL
nuoJ NADH:ubiquinone oxidoreductase subunit J; NADH dehydrogenase I 0.468 0.234 −0.234 NL
nuoK NADH:ubiquinone oxidoreductase subunit K; NADH dehydrogenase I 0.766 0.460 −0.306 NL
nuoL NADH:ubiquinone oxidoreductase subunit L; NADH dehydrogenase I 0.311 0.078 −0.233 NL
nuoN NADH:ubiquinone oxidoreductase subunit N; NADH dehydrogenase I 0.126 −0.210 −0.336 AH
sdhA Succinate dehydrogenase flavoprotein subunit 1.438 −0.018 −1.458 AH
sdhB Succinate dehydrogenase iron-sulfur protein 1.833 0.329 −1.505 AH
sdhC Succinate dehydrogenase membrane anchor subunit, cytochrome b556 2.728 0.890 −1.838 NL
sdhD Succinate dehydrogenase hydrophobic subunit 2.349 0.595 −1.754 NL
Amino acid catabolism and transport artI Arginine periplasmic binding protein 0.286 0.655 0.369 BH
artM Arginine periplasmic binding protein −0.115 0.293 0.407 BH
cadA Lysine decarboxylase, degradative 1.024 0.137 −0.887 AH
cysK O-Acetylserine sulfhydrylase A (cysteine synthase) 1.204 1.351 0.147 NL
dadA d-Amino acid dehydrogenase 1.273 −0.229 −1.501 AH
dadX d-Amino acid dehydrogenase 0.683 −0.393 −1.076 AH
dppC Dipeptide permease system −0.346 −0.032 0.315 AL
gdhA Glutamate dehydrogenase −0.080 −0.496 −0.416 BL
hisC Histidinol-phosphate aminotransferase 0.253 0.984 0.731 BH
hisF Imidazole glycerol phosphate synthase (cyclase) 0.150 0.703 0.553 BH
hisH Amidotransferase of imidazole glycerol phosphate synthase 0.034 0.393 0.359 BH
hisI PR-ATP pyrophosphatase and PR-AMP cyclohydrolase 0.234 0.641 0.407 BH
hisJ Histidine-binding protein −0.149 0.400 0.549 BH
lysC Aspartokinase III 1.464 −0.226 −1.690 AH
lysP Lysine permease 2.662 0.963 −1.698 NL
lysU Lysine-tRNA ligase 0.544 −0.077 −0.621 AH
potD Putrescine-ornithine transporter 0.053 0.509 0.456 BH
sdaA l-Serine deaminase, degradative −0.409 0.638 1.048 BH
sdaB l-Serine deaminase −1.111 −0.434 0.678 AL
sdaC H+/serine symporter; regulator of serine deaminase −1.205 −0.410 0.794 AL
tnaA Tryptophan deaminase, degradative; also deaminases serine and cysteine −3.805 0.223 4.028 AL
tnaB Tryptophan transporter −1.840 1.153 2.993 AL
tnaC tnaA leader peptide −5.026 0.490 5.517 AL
tdcB Threonine dehydratase, degradative −0.849 −0.296 0.553 NH
ydfG l-allo-Threonine, l-serine, d-serine dehydrogenase 0.407 0.287 −0.120 NL
a

Values in boldface are significant (α = 0.001).

b

NL, Neutral Low; AH, Acid High; AL, Acid Low; BH, Base High; BL, Base Low; NH, Neutral High.

c

CoA, coenzyme A.

d

PTS, phosphotransferase.

Many operons encoding processes of glycolysis and the TCA cycle, such as aceEF (pyruvate dehydrogenase), dhaKL (dihydroxyacetone kinase), pta (phosphotransacetylase), and pts (glucose phosphotransferase), showed elevated expression in acid. Others, however, were elevated at high pH. Operons elevated at high pH tended to be those induced by anaerobiosis, such as glpABC (anaerobic glycerol-3-phosphate dehydrogenase), pflBA (anaerobic pyruvate formate lyase), and dcu (anaerobic fumarate respiration). The mal system, however, is strongly repressed by acid (13, 31) and showed up as such in our arrays.

Membrane-bound systems for proton and electron transport were regulated by acid or base along lines largely consistent with their relative degree of export or import of H+. An example is the atp operon encoding F1Fo ATP synthase (32), which imports H+ during oxidative respiration. Most of the atp genes were strongly upregulated at high pH, whereas ndh and nuo (the NADH dehydrogenases I and II), which export H+, were down-regulated. The sdh gene (succinate dehydrogenase), which contributes electrons for proton export, is also down-regulated at high pH. On the other hand, cytochrome d oxidase (cyd) is expressed in preference to cytochrome o oxidase (cyo) at high pH, presumably because it exports half as many H+ per electron (14).

Enzymes for degradation of amino acids showed pH regulation as expected, with high pH favoring deaminase operons such as tna (tryptophan deaminase), sda (serine deaminase), and tdcB (threonine dehydratase). Acid induced only one of the decarboxylase operons, cad (lysine decarboxylase). Several decarboxylases are known to be induced by acid, but their induction is repressed by oxygen (4, 30), which may explain their absence in our highly aerobic cultures.

Oxidative stress and salicylate stress.

Several acid stress genes are known to overlap with oxidative stress, for example, the alkyl hydroperoxide reductase ahpC (9, 84), and certain permeant acids such as salicylate are considered oxidative stress agents (54). We surveyed our pH-regulated genes for overlap with response to H2O2, paraquat, and salicylate, as reported in references 54 and 84 (Table 5).

TABLE 5.

pH-regulated oxidative stress responsea

Gene Function Log2 pH ratiob
PQ, Sal, or H2O2 Classc
5/7 8.7/7 8.7/5
acnA Aconitase A 0.769 −0.143 −0.912 Sal AH
adhE Acetaldehyde-coenzyme A dehydrogenase 0.043 −0.380 −0.422 Sal BL
ahpC Alkyl hydroperoxide reductase small subunit 1.003 0.436 −0.568 PQ, H2O2 NL
ahpF Alkyl hydroperoxide reductase large subunit 0.777 −0.222 −0.999 H2O2 AH
aldA Aldehyde dehydrogenase, NAD linked 0.773 0.764 −0.009 PQ NL
alx (ygjT) Membrane protein, alkali induced −1.317 0.692 2.009 PQ− AL
artI Periplasmic arginine binding protein 0.286 0.655 0.369 PQ BH
aspA Aspartate ammonia-lyase (aspartase) −0.592 0.770 1.362 PQ− BH
carA Carbamoylphosphate synthase small subunit 0.029 1.059 1.030 PQ− BH
cfa Cyclopropane fatty acid synthase 2.075 −0.480 −2.555 Sal AH
cyaA Adenylate cyclase 0.691 −0.156 −0.846 Sal AH
cyoD Cytochrome o oxidase subunit IV 0.829 −0.192 −1.021 PQ NL
cysK Cysteine synthase 1.204 1.351 0.147 PQ, Sal, H2O2 NL
dadX Alanine racemase 0.683 −0.393 −1.076 PQ AH
deoA Thymidine phosphorylase −1.031 −2.003 −0.972 Sal BL
deoB Deoxyribouratase, phosphopentomutase −0.731 −1.502 −0.771 PQ, Sal BL
dhaH Dihydroxyacetone phosphoryl donor 1.547 −0.527 −2.074 Sal AH
dhaK Dihydroxyacetone kinase 1.041 −0.379 −1.420 Sal AH
dnaK HSP-70-type molecular chaperone −0.894 −2.249 −1.356 Sal BL
dps Stress response DNA-binding protein 1.130 0.105 −1.025 PQ, Sal, H2O2 NL
fliS Flagellar synthesis; flagellar regulon member −0.174 −1.947 −1.772 PQ− BL
fpr Ferredoxin NADP+ reductase; anaerobic 0.565 0.275 −0.289 PQ, H2O2 NL
gapA GAPDHd A 0.093 −0.752 −0.844 Sal BL
gatA Galactitol-specific enzyme IIA of PTSe 0.929 −0.626 −1.555 PQ, Sal AH
gatB Galactitol-specific enzyme IIB of PTS 0.632 −0.895 −1.528 PQ, Sal AH
gatC Galactitol-specific enzyme IIC of PTS 0.973 −0.949 −1.922 Sal AH
gatD Galactitol-1-phosphate dehydrogenase 1.037 −0.893 −1.930 PQ, Sal AH
gatZ Tagatose 6-phosphate aldolase 2 0.774 −0.605 −1.380 Sal AH
gltA Citrate synthase 0.288 −0.559 −0.846 PQ, Sal AH
gltB Glutamate synthase, large subunit 0.846 −1.425 −2.271 Sal AH
grxA Glutaredoxin 1 1.666 −0.106 −1.772 H2O2 AH
gshB Glutathione synthetase 0.768 −0.081 −0.848 Sal AH
hdeA Periplasmic acid chaperone 0.841 −0.326 −1.167 Sal AH
hdeB Periplasmic acid chaperone 0.782 −0.622 −1.404 Sal AH
hisF Cyclase component of IGP synthase 0.150 0.703 0.553 PQ− BH
ibpB Chaperone, HSP20 family −1.691 −2.618 −0.928 H2O2 BL
katG Catalase hydrogen peroxidase 1 −0.578 −0.313 0.265 H2O2 AL
lamB Maltose high-affinity uptake −4.881 −3.735 1.146 PQ NH
lldP l-Lactate permease 1.527 0.219 −1.307 Sal NL
lysU Lysyl tRNA synthetase, inducible 0.544 −0.077 −0.621 Sal AH
malE Maltose-binding protein, periplasmic −3.425 −2.933 0.491 PQ NH
malK Maltose transport complex, ATP-binding subunit −4.790 −3.748 1.041 PQ NH
manX PTS family, mannose-specific enzyme IIA component 0.000 0.517 0.517 Sal AL
map Methionine aminopeptidase 0.510 0.056 −0.454 PQ AH
marA Multiple antibiotic resistance 1.321 −0.516 −1.836 PQ, Sal AH
marB Regulator for mar 0.894 −0.172 −1.066 Sal AH
marR Repressor of mar 0.352 −0.437 −0.789 Sal AH
mdaB Drug activity modulator 0.444 −0.223 −0.667 Sal AH
murF d-Alanyl:d-alanine adding to cell wall 0.216 −0.170 −0.386 PQ AH
nfnB Nitrofurantoin resistance; nitroreductase 0.818 −0.204 −1.022 PQ, Sal AH
nuoI NADH dehydrogenase I subunit 0.792 0.310 −0.481 PQ NL
nuoK NADH dehydrogenase I subunit 0.766 0.460 −0.306 PQ NL
ompF Outer membrane porin −2.834 −0.892 1.942 PQ− AL
ompT Outer membrane protease VII −0.815 −1.204 −0.389 Sal− BL
pdhR Repressor of pdh 0.755 −0.109 −0.863 PQ AH
pepN Aminopeptidase N 0.671 0.118 −0.553 Sal NL
pflB Pyruvate formate lyase I (anaerobic) 0.812 0.578 −0.234 Sal NL
pgi Glucose phosphate isomerase 0.654 0.177 −0.477 PQ NL
ptsG PTS family IIC, glucose specific 0.841 0.554 −0.287 PQ NL
putA Proline dehydrogenase −1.062 −0.353 0.708 Sal AL
pyrB Aspartate transcarbamylase 0.133 0.608 0.475 PQ−, Sal− BH
sdhB Succinate dehydrogenase 1.833 0.329 −1.505 PQ AH
tnaA Tryptophanase −3.805 0.223 4.028 H2O2 AL
treB Tre-specific PTS enzyme II −3.156 0.739 3.895 Sal− AL
yahA Putative repressor −0.667 0.422 1.089 PQ− AL
yaiA Function unknown 1.154 0.416 −0.738 H2O2 NL
ybjC Function unknown 0.463 −0.447 −0.910 PQ, Sal AH
ycfR Function unknown 0.537 −0.819 −1.356 H2O2 AH
yfiA Stabilizes ribosome against dissociation −1.782 0.581 2.364 PQ−, Sal, H2O2 AL
yggJ Function unknown 0.576 −0.178 −0.755 Sal AH
yqjD Function unknown −0.347 0.378 0.725 Sal AL
yncE Function unknown −0.476 0.331 0.808 PQ, Sal AL
a

Oxidative response is based on data in references 54 and 84. Induction was by H2O2, paraquat (PQ), or sodium salicylate (Sal). Repression is indicated by minus sign (Sal−, PQ−).

b

Values in boldface indicate significance (α = 0.001).

c

AH, Acid High; AL, Acid Low; BH, Base High; BL, Base Low; NH, Neutral High; NL, Neutral Low.

d

GAPDH, glyceraldehyde-3-phosphate dehydrogenase.

e

PTS, phosphotransferase.

Of the 73 pH-dependent genes known to be induced by H2O2, paraquat, or salicylate, virtually all were induced by acid or repressed by base. This finding confirms our hypothesis of a strong connection between acid stress and oxidative stress. It may be that low pH amplifies the toxicity of oxygen radicals. Genes repressed by paraquat or salicylate were repressed in acid or induced at high pH, such as the base-inducible membrane protein gene alx, the histidine cyclase gene hisF, and outer membrane protein gene ompF. An exception to these generalizations was the maltose regulon (lamB, malE, and malK), which was repressed by acid but induced by paraquat.

Envelope and periplasmic stress.

A large part of E. coli function takes place in the outer membrane and envelope (48) and the periplasm (49), compartments essentially exposed to “extracellular” pH. Thus, it is not surprising that several envelope and periplasmic components show pH-dependent expression (16, 23, 73, 82). Our microarrays revealed an even greater number of such responses (Table 6). Both acid and base induction were observed. Acid-induced periplasmic proteins included the well-known acid chaperone from hdeAB (23), as well as the newly observed TolA-binding protein (ybgF) and the lipoprotein from pal. High pH induced the ferric transporters from fecAB and fhuD, possibly due to low iron solubility at high pH. At high pH, various transport proteins and redox modulators such as that from dsbA are known to be induced. In addition, several additional base-induced periplasmic and envelope proteins appeared, including the vitamin B12 transporter from btuB, the outer membrane protein from nmpC, and the peptidylprolyl-cis-trans-isomerase from ppiA.

TABLE 6.

Envelope and periplasmic genes

Gene Function Log2 pH ratioa
Classb
5/7 8.7/7 8.7/5
artI Periplasmic binding protein of Arg transport system 0.286 0.655 0.369 BH
artM Arginine periplasmic binding protein −0.115 0.293 0.407 BH
btuB B-12 transporter, outer membrane receptor −0.667 0.037 0.704 AL
cirA Colicin I receptor production −0.885 0.260 1.145 AL
clpX ATPase subunit of ClpXP protease −0.361 −0.625 −0.264 BL
cpxA Periplasmic stress sensor (CpxAR) −0.461 0.230 0.691 AL
cpxP CpxAR-regulated periplasmic stress protein −3.232 1.002 4.234 AL
cpxR Periplasmic stress response regulator (CpxAR) −0.662 0.590 1.251 AL
dsbA Thiol:disulfide interchange, periplasmic 0.086 1.117 1.031 BH
dsbC Disulfide bond isomerase, periplasmic chaperone −0.288 −0.534 −0.246 BL
fadL Fatty acid transport, outer membrane −0.861 1.069 1.931 BH
fecA Outer membrane ferric citrate receptor −2.608 −1.128 1.480 NH
fecB Periplasmic ferric citrate-binding protein −2.842 −1.538 1.304 NH
fepA Ferrienterobactin outer membrane receptor −0.966 −0.289 0.676 NH
fhuD Ferric hydroxamate binding protein; hydroxamate-dependent iron uptake −0.767 −0.180 0.587 AL
fliY Cystine-binding protein, periplasmic 0.233 −0.252 −0.484 AH
hdeA Acid periplasmic chaperone 0.841 −0.326 −1.167 AH
hdeB Acid periplasmic protein 0.782 −0.622 −1.404 AH
hisJ High-affinity histidine-binding protein −0.149 0.400 0.549 BH
hlpA Periplasmic chaperone for OMPsc 0.099 −0.661 −0.759 BL
lamB Maltoporin, maltose high-affinity uptake; phage lambda receptor −4.881 −3.735 1.146 NH
lon DNA-binding, ATP-dependent protease −0.600 −1.666 −1.065 BL
malE Maltose-binding protein, periplasmic −3.425 −2.933 0.491 NH
malM Maltose operon periplasmic protein −4.643 −3.780 0.863 NH
mltB Membrane-bound murein hydrolase −0.739 −0.419 0.320 NH
nmpC Outer membrane −4.012 −0.051 3.961 AL
ompF Outer membrane porin protein 1a −2.834 −0.892 1.942 AL
ompT Outer membrane protease VII −0.815 −1.204 −0.389 BL
ompX OMP, induces RNAP-sigma E 1.523 0.439 −1.083 NL
oppA Periplasmic oligopeptide binding protein 1.358 0.406 −0.953 NL
pal Lipoprotein associated with peptidoglycan 0.532 0.379 −0.153 NL
potD Spermidine-binding membrane protein; regulates pot 0.053 0.509 0.456 BH
ppiA Rotamase; peptidylprolyl-cis-trans-isomerase A −0.526 0.231 0.757 AL
pstS High-affinity, periplasmic phosphate binding protein −2.479 −2.262 0.217 NH
rbsB d-Ribose binding protein, periplasmic −0.749 −1.150 −0.401 BL
rseB Periplasmic, binds RseA; enhances RpoE-RseA cytoplasmic complex formation 0.631 −0.105 −0.736 AH
secD SecDF-YajC inner membrane secretion complex −0.165 0.287 0.452 AL
surA Periplasmic outer membrane porin chaperone, stationary phase 0.310 0.209 −0.101 NL
tatA Twin arginine translocation 0.043 0.679 −0.636 BH
tatB Twin arginine translocation 0.075 0.446 0.372 BH
tolB Group A colicin uptake and tolerance 0.498 0.421 −0.077 NL
tpX Thiol peroxidase, antioxidant 0.637 0.178 −0.459 NL
tsx Phage T6, colicin K resistance; nucleoside channel −0.707 −0.904 −0.196 BL
ybgF TolA-binding periplasmic protein 0.312 −0.007 −0.318 AH
yceI Function unknown; periplasmic protein −0.036 1.038 1.074 BH
yhcN Periplasmic protein 3.064 −1.136 −4.199 AH
a

Values in boldface indicate significance (α = 0.001).

b

AH, Acid High; AL, Acid Low; BH, Base High; BL, Base Low; NH, Neutral High; NL, Neutral Low.

c

OMPs, outer membrane proteins.

d

RNAP, RNA polymerase.

Universal stress and heat shock.

Various heat shock and universal stress proteins are inducible by the permeant acid benzoate, such as the products of clpB, htpG, dnaK, groS, and uspA (38). Some of these showed pH response in our microarrays (Table 7). The DNA damage response gene uspD was acid induced, as was dps, encoding the DNA-binding protein involved in stationary phase and acid resistance. Acid induced rseAB, the antisigma regulators of the rpoE envelope heat stress system (1). High pH induced the rpoH heat shock sigma 32 gene (28) as well as heat shock proteasome genes hslUV and regulators hslOR.

TABLE 7.

Universal stress and heat shock response genes

Gene Function Log2 pH ratioa
Classb
5/7 8.7/7 8.7/5
ahpC Alkyl hydroperoxide reductase 1.003 0.436 −0.568 NL
ahpF NAD(P)H:peroxiredoxin oxidoreductase 0.777 −0.222 −0.999 AH
cfa Cyclopropane fatty acid synthase; acid resistance in stationary phase 2.075 −0.480 −2.555 AH
clpB ClpB protease, ATP-dependent chaperone −0.219 −1.963 −1.744 BL
cysK Cysteine synthase, o-acetylserine sulfhydrylase A 1.204 1.351 0.147 NL
cysZ Unknown function 0.440 0.212 −0.228 NL
dinI Inhibits RecA coprotease 0.132 0.568 0.436 BH
dinJ Induced by DNA damage 0.683 1.192 0.508 BH
dnaJ DnaK cochaperone −0.908 −2.155 −1.247 BL
dnaK HSP-70-type molecular chaperone −0.894 −2.249 −1.356 BL
dps Stress response DNA-binding protein 1.130 0.105 −1.025 NL
grpE Nucleotide exchange factor for DnaKJ −0.491 −1.141 −0.650 BL
grxA Glutaredoxin 1 1.666 −0.106 −1.772 AH
hdeA Acid periplasmic chaperone 0.841 −0.326 −1.167 AH
hdeB Acid periplasmic chaperone 0.782 −0.622 −1.404 AH
hslJ Heat-inducible novobiocin resistance 0.368 −0.676 −1.044 AH
hslU Heat-inducible ATP-dependent protease −0.745 −1.688 −0.946 BL
hslV Heat-inducible ATP-dependent protease −1.125 −1.913 −0.788 BL
ibpB Heat-inducible chaperone, HSP20 family −1.691 −2.618 −0.928 BL
hslO Hsp33, cytoplasmic heat shock chaperone activated by disulfide bond formation −1.453 −1.737 −0.284 BL
hslR (yrfH) Hsp15, heat shock, binds RNA and DNA −1.773 −1.940 −0.168 NH
katG Catalase-hydrogen peroxidase I −0.578 −0.313 0.265 AL
rpoE RNAPc sigma E, envelope heat stress 0.310 −0.427 −0.738 AH
rpoH RNAP sigma 32, heat shock regulons −0.378 0.084 0.462 AL
rseA Anti-RpoE sigma factor, spans inner membrane 0.419 −0.159 −0.578 AH
rseB Binds periplasmic domain of anti-RpoE sigma RseA 0.631 −0.105 −0.736 AH
sodB Superoxide dismutase, Fe; acid inducible 0.752 0.280 −0.472 NL
ycdB Function unknown, peroxidase homolog 0.654 −0.329 −0.983 AH
ycdO Acid inducible, function unknown 0.842 −0.625 −1.467 AH
uspD (yiiT) UV resistance 1.333 −0.497 −1.830 AH
a

Values in boldface indicate significance (α = 0.001).

b

AH, Acid High; AL, Acid Low; BH, Base High; BL, Base Low; NH, Neutral High; NL, Neutral Low.

c

RNAP, RNA polymerase.

DISCUSSION

Overall, our work revealed a large number of genes not previously known to be regulated by pH. Furthermore, many of these genes had no previously known function or response, such as yhcN and yagU (induced by acid) and yifO and ymcG (induced by base).

An important question is to assess the biological relevance of the expression ratios reported (36, 51). Most of the ratios we reported as significant (boldface in Tables 3 through 7) are greater than twofold (log2 = 1). In many cases, all or most members of an operon fell in the same cluster and show similar expression profiles; the flagellar regulon was particularly consistent (Table 3). The gene probes are synthesized on the array independently of their operon map; thus, parallel expression profiles within operons do not reflect array position. Note that even genes with significant expression ratios of less than 2 (log2 = 1) tend to group with their operons. In previous studies, comparison with quantitative reverse transcriptase real-time PCR shows that microarray ratios, while quantitatively consistent, generally underestimate the actual differences in mRNA levels between the biological systems compared (83).

Flagellar biosynthesis and motility.

The effects of pH in flagellar biosynthesis and motility remain poorly understood. It has long been known that low external pH (thus, large ΔpH) contributes to the proton motive force that drives flagellar rotation (33). The cytoplasmic pH, however, must remain high; permeant acids such as acetate and benzoate, which depress internal pH and decrease proton motive force, are chemotactic repellents (67) and impair rotation of the flagellar motor (46). Low pH elicits negative chemotaxis (55, 67), whereas a pH increase up to 8.3 elicits a positive response (55).

In recent reports acid stress is associated with low motility (72), yet acetate has been reported to induce the flagellar regulon and enhance motility (53). We believe that the previous reports are limited in several ways. Reference 72 does not compare pH conditions directly but notes repression of flagellar genes in an hns mutant in which acid resistance is increased. The motility assay is not clearly described, and the acid dependence of flhDC-cat expression was observed on plasmids, not in the genome. Reference 53 reports induction of chromosomal flhDC-lacZ fusions by acetate. Those authors' assays of motility, however, show relatively small differences between pH conditions.

Our microarrays showed strong evidence for suppression of motility and chemotaxis at high pH. This evidence was supported by the decrease in motility at high pH, observed for both E. coli and S. enterica serovar Typhimurium, which swims twice as fast as E. coli. We also found weaker evidence for repression of che and mot genes at pH 5, but the flagellar synthesis genes were strongly induced at low pH. Overall, our data point to alkaline suppression of flagellar motility. Work in progress shows that, at high pH, the number of flagella per cell is decreased to one to three per cell (about 20% of normal) (S. Aizawa and J. Slonczewski, unpublished data).

No pH dependence was observed for the flagellar regulators flhD and flhC. On the other hand, in a microarray study of anaerobic cultures, flhD and flhC are induced by acid (E. Hayes and J. L. Slonczewski, unpublished data). Acid induction of these regulators would be consistent with the report of their induction by acetate (53). We did see acid induction of two known activators of flhDC: adenylate cyclase cyaA (37) (Acid High) and dnaK-dnaJ-grpE (64) (Base Low). We saw no acid induction of other flagellar activators such as crp (37), nor did we see alkaline induction of the negative flagellar regulator rcsCDB (22).

An alternative model is that pH regulation of the flagellar regulon is mediated by proteolysis, as in the case of ClpXP proteolysis of FlhD and FlhC (76). We find that ClpX is down-regulated at high pH (Base Low cluster), but a different protease could be involved.

Catabolism.

The picture of catabolism is more complicated, but in general our expression ratios confirm our present hypotheses of pH regulation while extending our knowledge to many more components. Systems that consume acids are enhanced at low pH. On the other hand, initial import and breakdown of some sugars, such as maltose, are favored at high pH, where they may quickly generate a large burst of fermentation acids.

With respect to proton export, E. coli appears to prefer components such as ATP synthase that import protons at high pH (counteracting the alkaline stress on cytoplasmic pH) and prefers to minimize proton export associated with the terminal oxidase cyd in preference to cyo. This observation is consistent with the previous report that cyd expression is higher at pH 7.5 than at pH 5.0 in an fnr mutant (14), although in those experiments cyo expression also increased with pH. It is likely that our broader range of pH classes (up to pH 8.7) provided a clearer picture of pH regulation of cyo and cyd.

Under amino acid catabolism, relatively few new components of pH response were observed. This makes sense, because most amino acid decarboxylases are repressed by oxygen (4, 68), as are deaminases such as sdaA (82). In preliminary experiments, we have repeated our microarray study on cultures grown anaerobically. Under anaerobiosis, several amino acid decarboxylases and deaminases show pH-dependent expression (Hayes and Slonczewski, unpublished).

Stress responses.

Several stress responses are known to interact with pH stress and pH resistance, including oxidative stress, heat shock, and envelope stress (for reviews see references 21 and 68). The overlap with salicylate stress could be explained in part by salicylate's effect as a permeant acid, stressing internal pH (60). The mar drug resistance operon is known to be coinduced by aromatic permeant acids and low pH (69) under regulation by MarR as well as by the superoxide regulator SoxRA (57).

Beyond salicylate, however, a large number of oxidative stress genes inducible by H2O2 or by paraquat showed significant pH-dependent expression, nearly all induced by acid or repressed by base. This finding confirms our hypothesis of a strong connection between acid stress and oxidative stress. Since so much of aerobic respiration is stepped up at pH 5, including cytochrome o oxidase, it is likely that acid conditions accelerate the production of oxygen radicals, thus inducing a partial oxidative stress response.

Various envelope and periplasmic stress responses are induced by acid, contributing to acid resistance; the best characterized in terms of mechanism is the acid-induced periplasmic chaperone HdeA (23). Extracellular acid induces a dimer-to-monomer transition in HdeA, which then suppresses aggregation by acid-denatured proteins. Our study reveals additional potential contributors to acid resistance and base resistance, including genes of unknown function such as yhcN, induced by acid, and yceI, induced by base.

Our study presents the most comprehensive picture to date of acid and base response by E. coli grown aerobically in complex medium. Overall, low pH accelerates acid consumption and proton export, while coinducing oxidative stress, possibly through increased production of oxygen radicals. High pH accelerates proton import while repressing the energy-expensive systems of flagellar biosynthesis and chemotaxis. Finally, pH differentially regulates a large number of periplasmic and envelope stress systems, as well as transporters, chaperones, and redox regulators.

Supplementary Material

[Supplemental material]

Acknowledgments

The class comparison and cluster analysis were performed using BRB ArrayTools v3.1 developed by Richard Simon and Amy Peng Lam. We thank Bryan Lin and Ariel Kahrl for excellent technical assistance.

This work was supported by grant MCB-0234732 from the National Science Foundation and by undergraduate research funds from the Kenyon College grant from the Howard Hughes Medical Institute Biological Sciences Education Program.

Footnotes

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

REFERENCES

  • 1.Ades, S. E., I. L. Grigorova, and C. A. Gross. 2003. Regulation of the alternative sigma factor σE during initiation, adaptation, and shutoff of the extracytoplasmic heat shock response in Escherichia coli. J. Bacteriol. 185:2512-2519. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 2.Argaman, L., R. Hershberg, J. Vogel, G. Bejerano, E. Gerhart, H. Wagner, H. Margalit, and S. Altuvia. 2001. Novel small RNA-encoding genes in the intergenic regions of Escherichia coli. Curr. Biol. 11:941-950. [DOI] [PubMed] [Google Scholar]
  • 3.Arnold, C. N., J. McElhanon, A. Lee, R. Leonhart, and D. A. Siegele. 2001. Global analysis of Escherichia coli gene expression during the acetate-induced acid tolerance response. J. Bacteriol. 183:2178-2186. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 4.Auger, E. A., K. E. Redding, T. Plumb, L. C. Childs, S. Y. Meng, and G. N. Bennett. 1989. Construction of lac fusions to the inducible arginine and lysine decarboxylase genes of Escherichia coli K-12. Mol. Microbiol. 3:609-620. [DOI] [PubMed] [Google Scholar]
  • 5.Bakker, E. P., and W. E. Mangerich. 1981. Interconversion of components of the bacterial proton motive force by electrogenic potassium transport. J. Bacteriol. 147:820-826. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 6.Bearson, B. L., L. Wilson, and J. W. Foster. 1998. A low pH-inducible, PhoPQ-dependent acid tolerance response protects Salmonella typhimurium against inorganic acid stress. J. Bacteriol. 180:2409-2417. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 7.Behari, J., L. Stagon, and S. B. Calderwood. 2001. pepA, a gene mediating pH regulation of virulence genes in Vibrio cholerae. J. Bacteriol. 183:178-188. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 8.Bingham, R. J., K. S. Hall, and J. L. Slonczewski. 1990. Alkaline induction of a novel gene locus, alx, in Escherichia coli. J. Bacteriol. 172:2184-2186. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 9.Blankenhorn, D., J. Phillips, and J. L. Slonczewski. 1999. Acid- and base-induced proteins during aerobic and anaerobic growth of Escherichia coli revealed by two-dimensional gel electrophoresis. J. Bacteriol. 181:2209-2216. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 10.Buurman, E. T., D. McLaggan, J. Naprstek, and W. Epstein. 2004. Multiple paths for nonphysiological transport of K+ in Escherichia coli. J. Bacteriol. 186:4238-4245. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 11.Castanie-Cornet, M., T. A. Penfound, D. Smith, J. Elliott, and J. W. Foster. 1999. Control of acid resistance in Escherichia coli. J. Bacteriol. 181:3525-3535. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 12.Castanie-Cornet, M., and J. W. Foster. 2001. Escherichia coli acid resistance: cAMP receptor protein and a 20 bp cis-acting sequence control pH and stationary phase expression of the gadA and gadBC glutamate decarboxylase genes. Microbiology 147:709-715. [DOI] [PubMed] [Google Scholar]
  • 13.Chagneau, C., M. Heyde, S. Alonso, R. Portalier, and P. Laloi. 2001. External-pH-dependent expression of the maltose regulon and ompF gene in Escherichia coli is affected by the level of glycerol kinase, encoded by glpK. J. Bacteriol. 183:5675-5683. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 14.Cotter, P. A., V. Chepuri, R. B. Gennis, and R. P. Gunsalus. 1990. Cytochrome o (cyoABCDE) and d (cydAB) oxidase gene expression in Escherichia coli is regulated by oxygen, pH, and the fnr gene product. J. Bacteriol. 172:6333-6338. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 15.Cummings, J. H., E. W. Pomare, W. J. Branch, C. P. Naylor, and G. T. Macfarlane. 1997. Short chain fatty acids in human large intestine, portal, hepatic and venous blood. Gut 28:1221-1227. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 16.DiGiuseppe, P. A., and T. J. Silhavy. 2003. Signal detection and target gene induction by the CpxRA two-component system. J. Bacteriol. 185:2432-2440. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 17.Drasar, B. S., M. Shiner, and G. M. McLeod. 1969. Studies on the intestinal flora. I. The bacterial flora of the gastrointestinal tract in healthy and achlorhydric persons. Gastroenterology 56:71-79. [PubMed] [Google Scholar]
  • 18.Dressman, J. B., R. R. Berardi, L. C. Dermentzoglou, T. L. Russel, S. P. Schmaltz, J. L. Barnett, and K. M. Jarvenpaa. 1990. Upper gastrointestinal (GI) pH in young, healthy men and women. Pharm. Res. 7:756-761. [DOI] [PubMed] [Google Scholar]
  • 19.Eisen, M. B., P. T. Spellman, P. O. Brown, and D. Botstein. 1998. Cluster analysis and display of genome-wide expression patterns. Proc. Natl. Acad. Sci. USA 95:14863-14868. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 20.Evans, D. F., G. Pye, R. Bramley, A. G. Clark, T. J. Dyson, and J. D. Hardcastle. 1988. Measurement of gastrointestinal pH profiles in normal ambulant human subjects. Gut 29:1035-1041. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 21.Foster, J. W. 2000. Microbial responses to acid stress, p. 99-115. In G. Storz and R. Hengge-Aronis (ed.), Bacterial stress responses. ASM Press, Washington, D.C.
  • 22.Francez-Charlot, A., B. Laugel, A. Van Gemert, 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]
  • 23.Gajiwala, K. S., and S. K. Burley. 2000. HDEA, a periplasmic protein that supports acid resistance in pathogenic enteric bacteria. J. Mol. Biol. 295:605-612. [DOI] [PubMed] [Google Scholar]
  • 24.Gerchman, Y., Y. Olami, A. Rimon, D. Taglicht, and S. Schuldiner. 1993. Histidine-226 is part of the pH sensor of NhaA, a Na+/H+ antiporter in Escherichia coli. Proc. Natl. Acad. Sci. USA 90:1212-1216. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 25.Giannella, R. A., S. A. Broitman, and N. Zamcheck. 1973. Influence of gastric acidity on bacterial and parasitic enteric infections. A perspective. Ann. Intern. Med. 78:271-276. [DOI] [PubMed] [Google Scholar]
  • 26.Gong, F., and C. Yanofsky. 2002. Instruction of translating ribosome by nascent peptide. Science 297:1864-1867. [DOI] [PubMed] [Google Scholar]
  • 27.Gong, S., H. Richard, and J. W. Foster. 2003. YjdE (AdiC) is the arginine:agmatine antiporter essential for arginine-dependent acid resistance in Escherichia coli. J. Bacteriol. 185:4402-4409. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 28.Gross, C. A. 1996. Function and regulation of the heat shock proteins, p. 1382-1399. 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, 2nd ed., vol. 1. ASM Press, Washington, D.C. [Google Scholar]
  • 29.Harold, F. M., and P. C. Maloney. 1996. Energy transduction by ion currents, p. 283-306. 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, 2nd ed. ASM Press, Washington, D.C.
  • 30.Hersh, B. M., F. T. Farooq, D. N. Barstad, D. Blankenhorn, and J. L. Slonczewski. 1996. A glutamate-dependent acid resistance gene in Escherichia coli. J. Bacteriol. 178:3978-3981. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 31.Heyde, M., P. Laloi, and R. Portalier. 2000. Involvement of carbon source and acetyl phosphate in the external-pH-dependent expression of porin genes in Escherichia coli. J. Bacteriol. 182:198-202. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 32.Kasimoglu, E., S. Park, J. Malek, C. P. Tseng, and R. P. Gunsalus. Transcriptional regulation of the proton-translocating ATPase (atpIBEFHAGDC) operon of Escherichia coli: control by cell growth. J. Bacteriol. 178:5563-5567. [DOI] [PMC free article] [PubMed]
  • 33.Khan, S., and R. Macnab. 1980. Proton chemical potential, proton electrical potential and bacterial motility. J. Mol. Biol. 138:599-614. [DOI] [PubMed] [Google Scholar]
  • 34.Kihara, M., and R. Macnab. 1981. Cytoplasmic pH mediates pH taxis and weak-acid repellent taxis of bacteria. J. Bacteriol. 145:1209-1221. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 35.Kirkpatrick, C., L. M. Maurer, N. E. Oyelakin, Y. N. Yoncheva, R. Maurer, and J. L. Slonczewski. 2001. Acetate and formate stress: opposite responses in the proteome of Escherichia coli. J. Bacteriol. 183:6466-6477. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 36.Kothapalli, R., S. J. Yoder, S. Mane, and T. P. Loughran, Jr. 2002. Microarray results: how accurate are they? BMC Bioinformatics 3:22. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 37.Kutsukake, K. 1997. Autogenous and global control of the flagellar master operon flhD, in Salmonella typhimurium. Mol. Gen. Genet. 254:440-448. [DOI] [PubMed] [Google Scholar]
  • 38.Lambert, L. A., K. Abshire, D. Blankenhorn, and J. L. Slonczewski. 1997. Proteins induced in Escherichia coli by benzoic acid. J. Bacteriol. 179:7595-7599. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 39.Li, C., and W. H. Wong. 2001. Model-based analysis of oligonucleotide arrays: expression index computation and outlier detection. Proc. Natl. Acad. Sci. USA 98:31-36. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 40.Li, C., and W. H. Wong. 2003. DNA-chip analyzer (dChip), p. 120-141. In G. Parmigiani, E. S. Garrett, R. Irizarry, and S. L. Zeger (ed.), The analysis of gene expression data: methods and software. Springer-Verlag, New York, N.Y.
  • 41.Ma, Z., H. Richard, D. L. Tucker, T. Conway, and J. W. Foster. 2002. Collaborative regulation of Escherichia coli glutamate-dependent acid resistance by two AraC-like regulators, GadX and GadW (YhiW). J. Bacteriol. 184:7001-7012. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 42.Macnab, R. M. 2003. How bacteria assemble flagella. Annu. Rev. Microbiol. 57:77-100. [DOI] [PubMed] [Google Scholar]
  • 43.Masse, E., N. Majdalani, and S. Gottesman. 2003. Regulatory roles for small RNAs in bacteria. Curr. Opin. Microbiol. 6:120-124. [DOI] [PubMed] [Google Scholar]
  • 44.Masuda, N., and G. Church. 2003. Regulatory network of acid resistance in Escherichia coli. Mol. Microbiol. 48:699-712. [DOI] [PubMed] [Google Scholar]
  • 45.McFall, E., and E. B. Newman. 1996. Amino acids as carbon sources, p. 358-379. 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, 2nd ed. ASM Press, Washington, D.C.
  • 46.Minamino, T., Y. Imae, F. Oosawa, Y. Kobayashi, and K. Oosawa. 2003. Effect of intracellular pH on rotational speed of bacteria flagellar motors. J. Bacteriol. 185:1190-1194. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 47.Neely, M. N., C. L. Dell, and E. R. Olson. 1994. Roles of LysP and CadC in mediating the lysine requirement for acid induction of the Escherichia coli cad operon. J. Bacteriol. 176:3278-3285. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 48.Nikaido, H. 1996. Outer membrane, p. 29-47. 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, 2nd ed. ASM Press, Washington, D.C.
  • 49.Oliver, D. B. 1996. Periplasm, p. 88-103. 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, 2nd ed. ASM Press, Washington, D.C.
  • 50.Payne, D., D. Tatham, E. D. Williamson, and R. W. Titball. 1998. The pH 6 antigen of Yersinia pestis binds to β1-linked galactosyl residues in glycosphingolipids. Infect. Immun. 66:4545-4548. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 51.Piper, M. D. W., P. Daran-Lipujade, C. Bro, B. Regenberg, S. Knudsen, J. Nielsens, and J. T. Pronk. 2002. Reproducibility of oligonucleotide microarray transcriptome analyses. J. Biol. Chem. 277:37001-37008. [DOI] [PubMed] [Google Scholar]
  • 52.Plack, R. H., Jr., and B. P. Rosen. 1980. Cation/proton antiport systems in Escherichia coli. Absence of potassium/proton antiporter activity in a pH-sensitive mutant. J. Biol. Chem. 255:3824-3825. [PubMed] [Google Scholar]
  • 53.Polen, T., D. Rittmann, V. F. Wendisch, and H. Sahm. 2003. DNA microarray analysis 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]
  • 54.Pomposiello, P., M. H. J. Bennik, and B. Demple. 2001. Genome-wide transcriptional profiling of Escherichia coli responses to superoxide stress and sodium salicylate. J. Bacteriol. 183:3890-3920. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 55.Repaske, D., and J. Adler. 1981. Change in intracellular pH of Escherichia coli mediates the chemotactic response to certain attractants and repellents. J. Bacteriol. 145:1196-1208. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 56.Roe, A. J., D. McLaggan, I. Davidson, C. O'Byrne, and I. R. Booth. 1998. Perturbations of anion balance during inhibition of growth of Escherichia coli by weak acids. J. Bacteriol. 180:767-772. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 57.Rosner, J. L., and J. L. Slonczewski. 1994. Dual regulation of inaA by the multiple antibiotic resistance (mar) and superoxide (soxRS) stress response systems of Escherichia coli. J. Bacteriol. 176:6262-6269. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 58.Rowbury, R. J., Z. Lazim, and M. Goodson. 1996. Regulatory aspects of alkali tolerance induction in Escherichia coli. Lett. Appl. Microbiol. 22:429-432. [DOI] [PubMed] [Google Scholar]
  • 59.Russell, J. B., and F. Diez-Gonzalez. 1998. The effect of fermentation acids on bacterial growth. Adv. Microb. Physiol. 39:205-234. [DOI] [PubMed] [Google Scholar]
  • 60.Salmond, C. V., R. G. Kroll, and I. Booth. 1984. The effect of food preservatives on pH homeostasis in Escherichia coli. J. Gen. Microbiol. 130:2845-2850. [DOI] [PubMed] [Google Scholar]
  • 61.Schadt, E. E., C. Li, B. Ellis, and W. Wong. 2001. Feature extraction and normalization algorithms for high-density oligonucleotide gene expression array data. J. Cell. Biochem. 84(S37):120-125. [DOI] [PubMed] [Google Scholar]
  • 62.Schadt, E. E., C. Li, C. Su, and W. H. Wong. 2001. Analyzing high-density oligonucleotide gene expression array data. J. Cell. Biochem. 80:192-202. [PubMed] [Google Scholar]
  • 63.Schuldiner, S., V. Agmon, J. Brandsma, A. Cohen, E. Friedman, and E. Padman. 1986. Induction of SOS function by alkaline intracellular pH in Escherichia coli. J. Bacteriol. 168:936-939. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 64.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]
  • 65.Skandamis, P. N., and G.-J. E. Nychas. 2000. Development and evaluation of a model predicting the survival of Escherichia coli O157:H7 NCTC 12900 in homemade eggplant salad at various temperatures, pHs, and oregano essential oil concentrations. Appl. Environ. Microbiol. 66:1646-1653. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 66.Slonczewski, J. L., T. N. Gonzalez, F. M. Bartholomew, and N. J. Holt. 1987. Mu d-directed lacZ fusions regulated by low pH in Escherichia coli. J. Bacteriol. 169:3001-3006. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 67.Slonczewski, J. L., R. M. Macnab, J. R. Alger, and A. M. Castle. 1982. Effects of pH and repellent tactic stimuli on protein methylation levels in Escherichia coli. J. Bacteriol. 152:384-399. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 68.Slonczewski, J. L., and J. W. Foster. 1996. pH-related genes and survival at extreme pH, p. 1539-1552. 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, 2nd ed. ASM Press, Washington, D.C.
  • 69.Slonczewski, J. L., B. P. Rosen, J. R. Alger, and R. M. Macnab. 1981. pH homeostasis in Escherichia coli: measurement by 31P nuclear magnetic resonance of methylphosphonate and phosphate. Proc. Natl. Acad. Sci. USA 78:6271-6275. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 70.Small, P., D. Blankenhorn, D. Welty, E. Zinser, and J. L. Slonczewski. 1994. Acid and base resistance in Escherichia coli and Shigella flexneri: role of rpoS and growth pH. J. Bacteriol. 176:1729-1737. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 71.Soksawatmaekhin, W., A. Kuraishi, K. Skata, K. Kashiwagi, and K. Igarahi. 2004. Excretion and uptake of cadaverine by CadB and its physiological functions in Escherichia coli. Mol. Microbiol. 51:1401-1412. [DOI] [PubMed] [Google Scholar]
  • 72.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]
  • 73.Stancik, L. M., D. M. Stancik, B. Schmidt, D. M. Barnhart, Y. N. Yoncheva, and J. L. Slonczewski. 2002. pH-dependent expression of periplasmic proteins and amino acid catabolism in Escherichia coli. J. Bacteriol. 184:4246-4258. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 74.Taglicht, D., E. Padan, A. B. Oppenheim, and S. Schuldiner. 1987. An alkaline shift induces the heat shock response in Escherichia coli. J. Bacteriol. 169:885-887. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 75.Taglicht, D., E. Padan, and S. Schuldiner. 1991. Overproduction and purification of a functional Na+/H+ antiporter coded by nhaA (ant) from Escherichia coli. J. Biol. Chem. 266:11289-11294. [PubMed] [Google Scholar]
  • 76.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]
  • 77.Tramonti, A., M. Visca, M. De Canio, M. Falconi, and D. De Biase. 2002. Functional characterization and regulation of gadX, a gene encoding an AraC/XylS-like transcriptional activator of the Escherichia coli glutamic acid decarboxylase system. J. Bacteriol. 184:2603-2613. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 78.Tucker, D. L., N. Tucker, and T. Conway. 2002. Gene expression profiling of the pH responses in Escherichia coli. J. Bacteriol. 184:6551-6558. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 79.Tucker, D. L., N. Tucker, Z. Ma, J. W. Foster, R. L. Miranda, P. S. Cohen, and T. Conway. 2003. Genes of the GadX-GadW regulon in Escherichia coli. J. Bacteriol. 185:3190-3201. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 80.White, S., F. E. Tuttle, D. Blankenhorn, D. C. Dosch, and J. L. Slonczewski. 1992. pH dependence and gene structure of inaA in Escherichia coli. J. Bacteriol. 174:1537-1543. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 81.Yeung, K. Y., and W. L. Ruzzo. 2001. Principal component analysis for clustering gene expression data. Bioinformatics 17:763-774. [DOI] [PubMed] [Google Scholar]
  • 82.Yohannes, E., D. M. Barnhart, and J. L. Slonczewski. 2004. pH-dependent catabolic protein expression during anaerobic growth of Escherichia coli K-12. J. Bacteriol. 186:192-199. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 83.Yuen, T., E. Wurmbach, R. L. Pfeffer, B. J. Ebersole, and S. C. Sealfon. 2002. Accuracy and calibration of commercial oligonucleotide and custom cDNA microarrays. Nucleic Acids Res. 30:1-9. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 84.Zheng, M., X. Wang, L. J. Templeton, D. R. Smulski, R. A. LaRossa, and G. Storz. 2001. DNA microarray-mediated transcriptional profiling of the Escherichia coli response to hydrogen peroxide. J. Bacteriol. 183:4562-4570. [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