Skip to main content
Journal of Bacteriology logoLink to Journal of Bacteriology
. 2006 May;188(10):3449–3462. doi: 10.1128/JB.188.10.3449-3462.2006

Characterization of the ArsRS Regulon of Helicobacter pylori, Involved in Acid Adaptation

Michael Pflock 1,, Nadja Finsterer 1,, Biju Joseph 1, Hans Mollenkopf 2, Thomas F Meyer 3, Dagmar Beier 1,*
PMCID: PMC1482845  PMID: 16672598

Abstract

The human gastric pathogen Helicobacter pylori is extremely well adapted to the highly acidic conditions encountered in the stomach. The pronounced acid resistance of H. pylori relies mainly on the ammonia-producing enzyme urease; however, urease-independent mechanisms are likely to contribute to acid adaptation. Acid-responsive gene regulation is mediated at least in part by the ArsRS two-component system consisting of the essential OmpR-like response regulator ArsR and the nonessential cognate histidine kinase ArsS, whose autophosphorylation is triggered in response to low pH. In this study, by global transcriptional profiling of an ArsS-deficient H. pylori mutant grown at pH 5.0, we define the ArsR∼P-dependent regulon consisting of 109 genes, including the urease gene cluster, the genes encoding the aliphatic amidases AmiE and AmiF, and the rocF gene encoding arginase. We show that ArsR∼P controls the acid-induced transcription of amiE and amiF by binding to extended regions located upstream of the −10 box of the respective promoters. In contrast, transcription of rocF is repressed by ArsR∼P at neutral, acidic, and mildly alkaline pH via high-affinity binding of the response regulator to a site overlapping the promoter of the rocF gene.


Helicobacter pylori is a human pathogen which is associated with gastric diseases like chronic active gastritis, peptic ulceration, adenocarcinoma, and mucosa-associated lymphoid-tissue lymphoma (4, 25, 37). The neutralophilic bacterium, which thrives in the mucus layer covering the gastric epithelium, is extremely well adapted to cope with the fluctuating low-pH conditions encountered in the human stomach and has evolved mechanisms both to survive severe acid shocks and to grow at moderately acidic pH. In accordance with the stomach being its unique habitat, sequencing of the H. pylori genome revealed a very restricted repertoire of transcriptional regulators (34); however, it is becoming increasingly evident that a considerable subset of these regulators is involved in the control of the acid response. In three independent studies performing genome-wide transcriptional profiling, between 101 and about 280 genes have been reported to be regulated in response to the exposure of H. pylori to low pH (6, 21, 44). Up to now, two transcriptional regulators, the metal-dependent regulators NikR and Fur and a two-component signal transduction system termed ArsRS, have been implicated in the transcriptional regulation of acid-responsive genes (26, 40, 41). The ArsRS system consists of the histidine kinase ArsS, comprising a periplasmic input domain of 111 amino acids and likely to be responsible for low-pH sensing, and the OmpR-like response regulator ArsR. Inactivation of the histidine kinase ArsS renders H. pylori unable to colonize mice (24), while the response regulator ArsR is essential for the in vitro growth of H. pylori, suggesting distinct functions for ArsR in its unphosphorylated and phosphorylated states (2, 29).

A major player in the acid resistance of H. pylori is the urease system, which is essential for colonization (11, 14, 22, 31, 36). By cleaving urea, the urease enzyme generates ammonia and carbon dioxide, the latter being subsequently converted to HCO3 by a periplasmic carbonic anhydrase. These buffering compounds capture protons that leak into the cytoplasm and periplasm and thereby maintain the cytoplasmic and periplasmic pH of the bacteria near neutrality (18, 33). The enzymatic activity of urease is controlled by the inner membrane, pH-gated channel UreI, which regulates the access of the substrate urea to the bacterial cell in response to acidic pH (28, 43). Transcription of both the ureAB operon encoding the enzymatic subunits of urease and the ureIEFGH operon encoding the channel protein UreI and accessory proteins which are required for the incorporation of the metal cofactor Ni2+ into the urease apoenzyme is positively regulated in response to low pH (6, 21, 27). This pH-responsive transcriptional induction is mediated by the ArsRS two-component system via the binding of the phosphorylated response regulator ArsR (ArsR∼P) to the promoters of the ureAB and ureIEFGH operons (27). Several other target genes of the ArsRS two-component system encoding mostly H. pylori-specific proteins of unknown function have already been identified (10, 15) and are likely to be involved in urease-independent mechanisms of acid adaptation. Interestingly, the rocF gene encoding arginase was also shown to be regulated by ArsRS (15). Arginase produces urea and ornithine by the cleavage of arginine, and it was suggested that this enzyme might contribute to urease-dependent acid resistance when urea is scarce in the surrounding medium (20). Transcription of the ureAB genes is also positively controlled in response to increasing concentrations of Ni2+ ions by the pleiotropic, metal-dependent regulator NikR, which is also involved in the regulation of genes involved in Ni2+ uptake and storage, Fe3+ uptake and storage, stress responses, motility, and encoding outer membrane proteins (7, 13, 38). The acid stimulon and the NikR regulon defined by transcriptome analyses overlap to some extent, underlining the role of NikR in the process of acid adaptation (6, 7).

Besides that of urease, the expression of other ammonia-producing enzymes was demonstrated to be induced in response to acidic pH (6, 21, 44). Most notably, transcription of the genes amiE and amiF, which encode aliphatic amidases, was strongly increased upon exposure of H. pylori to low pH. Amidase genes which were known so far to be predominantly associated with environmental bacteria are present in several stomach-colonizing Helicobacter species, like H. pylori, H. acinonychis, and H. felis, suggesting an important role for the enzymes in nitrogen metabolism and acid resistance for these Helicobacter species. However, so far the substrates cleaved by the amidases in vivo remain unknown (5, 32). Under iron-replete conditions, transcription of amiE has been shown to be repressed by Fur, which binds to the upstream region of amiE. To our knowledge, amiF transcription is not controlled by Fur (39). However, at low pH the activity of both amidases is conversely affected by the inactivation of Fur or the NikR protein, suggesting a role for both regulators in the acid-responsive regulation of the amidases (6, 40). Interestingly, Fur-deficient mutants are impaired in their acid tolerance, suggesting an important role for Fur in the acid response (3).

The aim of this study was to define the subset of acid-responsive genes which is transcriptionally regulated by the ArsRS two-component system. We show that at pH 5.0, compared to those of the wild type, 109 genes are differentially regulated in a mutant of H. pylori G27 lacking the histidine kinase ArsS. Furthermore, we demonstrate that acid-induced transcription of amiE and amiF is mediated directly by the ArsRS two-component system via binding of ArsR∼P to the promoter regions of the amidase genes. Unexpectedly, the rocF gene encoding arginase, whose transcription is repressed by ArsR∼P, did not show a pH-responsive transcription profile under the applied experimental conditions.

MATERIALS AND METHODS

Bacterial strains and growth conditions.

H. pylori strains G27 and 26695 are clinical isolates (34, 45). Strain H. pylori G27/HP165::km, lacking the histidine kinase ArsS, has been described previously (2). H. pylori strain G27/ΔarsSΔnikR, carrying deletions of both the arsS and nikR genes, was constructed through allelic replacement by transformation of strain G27/HP165::km with the suicide plasmid pNko::cm. pNko::cm is a derivative of pNko::km (27), in which the kanamycin resistance cassette was replaced by a chloramphenicol resistance cassette of Campylobacter coli (42). Chromosomal DNA of the transformants was checked by PCR with primers flanking the integration site for the correct replacement of the nikR gene. H. pylori strains were grown at 37°C under microaerophilic conditions on Columbia agar plates containing 5% horse blood, 0.2% cyclodextrin, and Dent's or Skirrow's antibiotic supplement. Liquid cultures were grown in brain heart infusion (BHI) broth containing Dent′s or Skirrow′s antibiotic supplement and 10% fetal calf serum. When required, blood agar plates or liquid broth were supplemented with kanamycin or chloramphenicol in a final concentration of 20 μg/ml. Acid exposure experiments were performed as follows: bacteria from a liquid culture were harvested at an optical density at 550 nm (OD550) of 0.7 by centrifugation and were resuspended in BHI broth containing fetal calf serum and an antibiotic supplement whose pH had been adjusted to pH 5.0 with hydrochloric acid. Incubation at 37°C under microaerophilic conditions was continued for 1 h. To study gene expression in bacteria grown under mildly alkaline conditions, cells from a logarithmic culture (OD550 = 0.7) were shifted to BHI broth whose pH had been adjusted to pH 8.0 or 8.5 with sodium hydroxide and cultivation was continued for 1 h. Escherichia coli DH5α was grown in Luria-Bertani broth. When necessary, antibiotics were added to the following final concentrations: ampicillin, 100 μg/ml; kanamycin, 50 μg/ml; and chloramphenicol, 30 μg/ml.

RNA isolation.

H. pylori RNA was isolated from bacteria grown to the logarithmic phase in liquid broth by using TRIzol reagent (Invitrogen) according to the manufacturer's protocol. RNA preparations intended to be used for global transcriptional profiling were further purified using the RNeasy mini kit (QIAGEN) according to the manufacturer's RNA-cleanup protocol. Residual DNA was removed by on-column digestion during RNA purification with QIAGEN RNase-free DNase (QIAGEN). The RNA concentration was quantified by determination of absorbance at 260 and 280 nm, and RNA integrity was checked by visualization on a 1.5% agarose gel.

Microarray hybridization and data analysis.

Transcriptome analyses were performed using a whole-genome microarray containing 1,649 PCR products generated with specific primer pairs derived from the genome sequences of H. pylori 26695 (34) and J99 (1), which were spotted in duplicate. Microarrays were produced as described previously (17). To determine genes which are differentially expressed in the ArsS-deficient mutant G27/HP165::km at pH 5.0, cDNA was prepared from RNA extracted from H. pylori G27 and G27/HP165::km after exposure of the bacteria to acidic pH for 1 h. A total of eight RNA samples from two independent preparations of RNA from strains G27 and G27/HP165::km were used for cDNA labeling and hybridization. Equal amounts of RNA (20 μg) were used to synthesize differentially labeled cDNA (Cy3-dCTP and Cy5-dCTP; Amersham Biosciences) during first-strand reverse transcription reactions with Superscript II reverse transcriptase and 6 μg of random primers (Invitrogen). Dye reversal color swaps were performed as follows: a cDNA sample using Cy3-dCTP and another using Cy5-dCTP were generated, resulting in four labeled cDNAs per color swap. Synthesized cDNAs were purified by using QiaQuick PCR purification columns (QIAGEN) according to the manufacturer's protocol. Cy5-dCTP- and Cy3-dCTP-labeled cDNAs were combined and concentrated by evaporation in a SpeedVac at 45°C. Samples were diluted to 3× SSC-0.1% sodium dodecyl sulfate (SDS) and hybridized to the H. pylori microarray (50°C, 18 h). The slides were washed for 5 min at room temperature with 0.1% SDS-0.1× SSC and then two times with 0.1× SSC. After being washed, the slides were scanned, using ScanArray HT, and analyzed by using ScanArray express software (Perkin Elmer). Spots were flagged and eliminated from analysis when the signal-to-background ratio was less than three or in obvious instances of high background or stray fluorescent signals. Median intensities of spots were background corrected, and differences in dye bias were normalized by using the LOWESS algorithm (46). The signal ratios used as measures of differential expression between the red and green channels were obtained from processed signal intensities. The ratios were further analyzed with Microsoft Excel (Microsoft) and SAM software for statistical significance (35). Only those genes which had at least seven unflagged replicates were considered after significance analysis by SAM (Table 1). The parameters used for significance analysis are indicated in the supplemental material (see Table S1 in the supplemental material).

TABLE 1.

ArsR∼P-activated and -repressed genes

ArsR∼P-regulated genesa Category ORF HP no.b ORF JHP no.b ΔarsS/WTc ratio Gene and functiond Genome organizatione pH-responsive transcription confirmed by indicated referencef
Activated
Ammonia production HP0067 JHP0062 0.36 Urease accessory protein, ureH op HP0071-HP0067 21
HP0068 JHP0063 0.22 Urease accessory protein, ureG op HP0071-HP0067 21
HP0069 JHP0064 0.39 Urease accessory protein, ureF op HP0071-HP0067 21
HP0070 JHP0065 0.25 Urease accessory protein, ureE op HP0071-HP0067 44
HP0071 JHP0066 0.21 Urea channel protein, ureI op HP0071-HP0067 21, 44
HP0072 JHP0067 0.24 Urease subunit B, ureB op HP0073-HP0072 21
HP0073 JHP0068 0.20 Urease subunit A, ureA op HP0073-HP0072 6, 21
HP0294 JHP0279 0.06 Aliphatic amidase, amiE m 6, 21, 44
HP1238 JHP1159 0.15 Formamidase, amiF m 6, 21
Amino acid biosynthesis HP0106 JHP0098 0.27 Cystathionine g-synthase, metB op HP0107-HP0105 ? 6, 21
HP0380 JHP1001 0.38 Glutamate dehydrogenase, dghA m 21, 44
Biosynthesis of cofactors, prosthetic groups, and carriers HP0755 JHP0692 0.43 Molybdopterin biosynthesis protein, moeB op HP0755-HP0757 6, 21
HP0240 JHP0225 0.34 Octaprenyl-diphosphate synthase, ispB op HP0243-HP0236 ?
HP0306* JHP0291 0.42 Glutamate-1-semialdehyde 2,1-aminomutase, hemL op HP0305-HP0308 21, 44
HP0824 JHP0763 0.37 Thioredoxin, trxA op HP0824-HP0825 44
HP0825 JHP0764 0.31 Thioredoxin reductase, trxB op HP0824-HP0825
HP1224 JHP1145 0.39 Uroporphyrinogen III cosynthase, hemD op HP1223-HP1225
Cell envelope HP1191 JHP1116 0.25 LPS heptosyltransferase, rfaF m 6
HP1105 JHP1031 0.47 LPS biosynthesis protein m
HP0078 JHP0073 0.19 Truncated outer membrane protein, omp3 op HP0078-HP0079 6, 44
HP0079 JHP0073 0.26 Outer membrane protein, omp3 op HP0078-HP0079 6, 44
HP0492 JHP0444 0.30 Predicted lipoprotein m 6
HP1456 JHP1349 0.48 Membrane-associated lipoprotein op HP1457-HP1454 44
HP1564 JHP1472 0.36 Outer membrane lipoprotein m
Cellular processes HP0389 JHP0992 0.45 Superoxide dismutase, sodB op HP0389-HP0388
HP0390 JHP0991 0.18 Predicted thiol peroxidase m
HP0875 JHP0809 0.43 Catalase, katA m 44
HP1563 JHP1471 0.29 Alkyl hydroperoxide reductase, ahpC m
HP1192 JHP1117 0.06 Secreted protein involved in flagellar motility m 6, 44
HP0243 JHP0228 0.28 Neutrophil-activating protein, napA op HP0243-HP0236 ? 44
DNA metabolism, restriction, and modification HP1022 JHP0402 0.30 Predicted 5′-3′ exonuclease m 21, 44
Energy metabolism HP0890 JHP0823 0.47 Predicted short-chain oxidoreductase op HP0890-HP0888
HP0924 JHP0858 0.44 4-oxalocrotonate tautomerase, dmpI m
HP0954 JHP0888 0.45 NADPH nitroreductase op HP0961-HP0953 ?
HP1104 JHP1030 0.34 Predicted mannitol dehydrogenease m 44
JHP0585* 0.27 Predicted 3-hydroxyacid dehydrogenase op JHP0583-JHP0586 21
JHP1429 0.29 Predicted mannitol dehydrogenease m
Fatty acid and phospholipid metabolism HP0891 JHP0824 0.41 Predicted acyl coenzyme A thioesterase m 44
Central intermediary metabolism HP0757 JHP0649 0.41 Predicted carbon nitrogen hydrolase op HP0755-HP0757 21
Protein fate HP1332* JHP1225 0.48 Cochaperone and heat shock protein op HP1332-HP1330 21
Transport and binding proteins HP0228 JHP0213 0.27 Predicted sulfate permease m 6, 21
HP0889 JHP0822 0.40 Iron (III) dicitrate ABC transporter, permease op HP0890-HP0888 44
HP1339 JHP1258 0.37 Biopolymer transport protein, exbB2 op HP1339-HP1341 21
HP1340* JHP1259 0.47 Biopolymer transport protein, exbD2 op HP1339-HP1341 21
HP1341 JHP1260 0.48 Siderophore-mediated iron transport protein, tonB2 op HP1339-HP1341
HP1331 JHP1251 0.47 Predicted branched-chain amino acid transport protein op HP1332-HP1330 21
HP1466 JHP1359 0.44 Predicted ABC transport system permease op HP1466-HP1462 ? 6
HP1427 JHP1320 0.20 Histidine-rich protein, hpn m
HP1432 JHP1321 0.07 Histidine-rich protein m 6, 21
Unknown HP1098 JHP1024 0.43 Cysteine-rich protein C, hcpC op HP1100-HP1097 ?
JHP1437 0.16 Cysteine-rich protein m
HP1225 JHP1146 0.27 Predicted CrcB integral membrane protein op HP1223-HP1225 21, 44
Hypothetical HP0189 JHP0175 0.47 Hypothetical integral membrane protein m
HP1286 JHP1206 0.48 Conserved hypothetical secreted protein m
HP0081 0.04 Hypothetical protein m
HP0118 JHP0110 0.19 Hypothetical protein m 6, 21
HP0120 0.17 Hypothetical protein m 6
HP0242* JHP0227 0.33 Hypothetical protein op HP0243-HP0236 ? 44
HP0305 JHP0290 0.24 Hypothetical protein op HP0305-HP0308 ? 44
HP0307* JHP0292 0.48 Hypothetical protein op HP0305-HP0308 ? 6
HP0423 0.35 Hypothetical protein op HP0427-HP0423 ? 6
HP0425 0.43 Hypothetical protein op HP0427-HP0423 ?
HP0426 0.40 Hypothetical protein op HP0427-HP0423 ? 6
HP0641 JHP0584 0.27 Hypothetical protein op HP640-HP0642 21
HP0731 JHP0668 0.29 Hypothetical protein op HP0733-HP0731
HP0733 JHP0670 0.47 Hypothetical protein op HP0733-HP0731
HP0963 JHP0897 0.38 Hypothetical protein op HP0966-HP0962
HP0964 JHP0898 0.37 Hypothetical protein op HP0966-HP0962
JHP0899
HP1154 JHP1081 0.48 Hypothetical protein op HP1154-HP1155 21
HP1187 JHP1113 0.28 Hypothetical protein m 21, 44
HP1188 0.12 Hypothetical protein m 6, 21, 44
HP1223 JHP1144 0.37 Hypothetical protein op HP1223-HP1225
HP1408 JHP1300 0.20 Hypothetical protein op HP1408-HP1412 ?
HP1412 JHP1307 0.32 Hypothetical protein op HP1408-HP1412 ? 44
JHP0959 0.32 Hypothetical protein m
Repressed
Cell envelope HP0229 JHP0214 2.51 Outer membrane protein, omp6 m 6, 21, 44
HP0722 JHP0659 3.87 Outer membrane protein, omp16 m 6, 21
HP0725 JHP0662 5.68 Outer membrane protein, omp17 m 6, 21
HP0788 JHP0725 2.64 Predicted outer membrane protein, hofF m
HP1083 JHP0342 2.79 Predicted outer membrane protein, hofB m
HP1167 JHP1094 5.33 Predicted outer membrane protein, hofH m 6, 21
Cellular processes HP1560* JHP1468 2.24 Cell division protein, ftsW m 6
HP0017 JHP0015 2.22 virB4 homolog involved in natural competence, comB4 op HP0015-HP0017
HP1527 JHP1416 3.86 Protein involved in natural competence, comH m 6, 21
HP1585 JHP1492 2.09 Flagellar basal-body rod protein
DNA metabolism, restriction, and modification HP1526 JHP1415 2.14 Exodeoxyribonuclease, lexA m
Energy metabolism HP1398 JHP1428 4.31 Alanine dehydrogenase, ald m
HP1399 JHP1427 3.23 Arginase, rocF m
Protein fate HP0109* JHP0101 3.00 Chaperone and heat shock protein 70, dnaK op HP0111-HP0109 21
HP0110 JHP0102 2.86 Cochaperone and heat shock protein, grpE op HP0111-HP0109 21
HP0382* JHP0999 2.36 Predicted zinc-metalloprotease op HP0382-HP0381 21
Purines, pyrimidines, nucleosides, and nucleotides HP1178 JHP1104 3.78 Purin nucleoside phosphorylase, deoD op HP1180-HP1178 21
HP1179 JHP1105 3.86 Phosphopentamutase op HP1180-HP1178
Transcription regulation HP0166 JHP0152 2.20 Two-component response regulator, arsR op HP0166-HP0162 6
Transport and binding proteins HP0686 JHP0626 4.06 Iron(III) dicitrate transport protein, fecA1 m
HP1129 JHP1057 2.81 Biopolymer transport protein, exbD1 op HP1137-HP1123 ?
HP1130 JHP1058 3.19 Biopolymer transport protein, exbB1 op HP1137-HP1123 ?
HP1174 JHP1101 5.65 Glucose/galactose transporter m
HP1180 JHP1106 4.17 Pyrimidine nucleoside transport protein op HP1180-HP1178 6, 21
Hypothetical HP1175 JHP1102 2.12 Hypothetical integral membrane protein m 6, 21
HP0681 JHP0622 10.40 Hypothetical protein op HP0682-HP0681
HP0682 JHP0623 16.56 Hypothetical protein op HP0682-HP0681
HP0688* JHP0628 2.65 Hypothetical protein op HP0688-HP0689
HP0689 JHP0628 2.02 Hypothetical protein op HP0688-HP0689
HP0709 JHP0648 3.97 Hypothetical protein m 6
HP0947 JHP0881 3.74 Hypothetical protein m
HP1288* JHP1208 4.48 Hypothetical protein op HP1288-HP1289 6
HP1289 JHP1209 22.77 Hypothetical protein op HP1288-HP1289
HP1322 JHP1242 2.13 Hypothetical protein m 21
a

Genes listed are those whose transcription at pH 5.0 differed more than twofold (ratio, >2.0 or <0.5) for the ArsS-deficient mutant H. pylori G27/HP165::km compared to that for the G27 wild type, according to microarray analysis.

b

ORF numbers and the prediction of transcriptional units are based on the genome sequences of H. pylori 26695 and J99 (1, 34). Asterisks indicate ORFs which were determined by the SAM algorithm to be not statistically significant (see Table S1 in the supplemental material) but which were retained for data evaluation because they are part of a predicted operon structure with other differentially expressed genes or because their pH-responsive transcription was confirmed in another study.

c

WT, wild type.

d

The functional annotation is that used by the PyloriGene database (http://www.pasteur.fr/english.html).

e

m, monocistronically transcribed genes; op, putative transcriptional unit. Question marks indicate that the proposed operon structures cannot be unambiguously deduced from the genome sequences.

f

Where a reference(s) is given, pH-responsive transcription was confirmed by the studies of Merrell et al. (21), Wen et al. (44), or Bury-Moné et al. (6).

To determine the significance of differential expression, RNA was isolated from the H. pylori G27 wild type grown in BHI broth (pH 5.0) and 20 μg of this RNA was labeled with either Cy3-dCTP or Cy5-dCTP. The two cDNA probes generated were hybridized onto the same slide, and the data were analyzed as mentioned above. Signal ratios of <0.5 and >2.0 were analyzed further.

Primer extension and RNA slot blot analysis.

Primer extension analysis was performed essentially as described previously (26), using 0.5 pMol of γ32P-end-labeled oligonucleotides amiE-PE, amiF-PE, rocF-PE, 682-PE, 1104-PE, 1174-PE, 1398-PE, and 1563-PE (Table 2) and 30 μg of RNA. Plasmids which were used as template DNA in the sequencing reactions performed with the aforementioned primers were constructed as follows. EcoRI-BamHI fragments of 328 bp and 517 bp were amplified from chromosomal DNA of H. pylori 26695 with primer pairs PamiE-5/PamiE-3 and PamiF-5/PamiF-3, respectively, and were ligated into the cloning vector pSL1180 (Amersham Biosciences) to yield plasmids pSL-PamiE and pSL-PamiF. Plasmid pSL-ProcF was obtained by cloning a 365-bp EcoRI-BamHI fragment that was amplified with primer pair ProcF-5/ProcF-3 into pSL1180 vector DNA. Primer extension experiments were performed three times with independently prepared RNAs. Quantification of the signals from the primer extension products was performed using a Typhoon 9200 Variable Mode imager (Amersham Biosciences) and ImageMaster TotalLab software (Amersham Biosciences). RNA slot blot analysis was performed as follows: RNA (20 to100 μg) was denatured in 1× MOPS (morpholinepropanesulfonic acid) buffer containing 50% formamide and 6% formaldehyde. The samples were incubated at 65°C for 5 min and cooled on ice before one volume of 20× SSC was added. The denatured samples were filtered through a positively charged nylon membrane (Hybond N+, Amersham Biosciences), using a Bio-Dot chamber (Bio-Rad). After UV cross-linking was performed, the nylon membrane was prehybridized for 1 h at 42°C in hybridization buffer (ECL gold hybridization buffer; Amersham Biosciences). The PCR products used as hybridization probes were amplified with primer pairs 229-5/229-3, 725-5/725-3, 686-5/686-3, and 16S-5/16S-3 and were labeled nonradioactively, using the ECL Direct nucleic acid-labeling system (Amersham Biosciences) according to the manufacturer's instructions.

TABLE 2.

Oligonucleotides used in this study

Oligonucleotide or primer Sequence (5′ to 3′)a Siteb Orientation of strand Positionc
PamiE-5 gattttgaattcTATTGTATTAACGCGCTATATGG EcoRI + 310783-310805
PamiE-3 attttcggatccTCATTCTTAGTGTGGAGTCTAGG BamHI 311089-311111
PamiF-5 caagctgaattcTGGTCATCATGGGAGCAACC EcoRI + 1311723-1311742
SamiF-5 ccagcagaattcAGAAAGTAGCCCAGGTCCTAA EcoRI + 1311918-1311938
PamiF-3 aatatcggatccGCTATTGACAATTGGCACAGG BamHI 1312219-1312239
ProcF-5 atccttgaattcGATTAGTGCCACATCATCAGG EcoRI + 1459647-1459667
ProcF-3 agccatggatccGCTTAAAGCCTCTCTCAAACG BamHI 1459992-1460012
amiE-PE TAGGCATCTTATAATTAACTACCGC 311068-311092
amiF-PE AGGAAACTGAATGGCTGCC 1312203-1312221
rocF-PE GCTTAAAGCCTCTCTCAAACG 1459992-1460012
682-PE ATCCACTCATAACAAATC 732444-732461
1104-PE TAGCAAAACCTTTAGATTGAAC 1165334-1165355
1174-PE GAATAGCGCTGTCAAACTCC + 1241646-1241665
1398-PE CTCGGGATTCTAAATCCATGC + 1459679-1459699
1563-PE GGCAGGTGCTTTAAAGTCTGG 1645245-1645265
229-5 ATGAAAAAAACGATTTTACTT 240139-240159
229-3 TTGATTAAGGTTTTTATTGAA + 239440-239460
725-5 AAGCAAAGCATTCAAAACGCC 780757-780777
725-3 GATGTCTTTAGCGAATTTAGG + 780058-780078
686-5 ATGAAAAGAATTTTAGTCTCT 737126-737146
686-3 GTTGTATCTAAACCCTTGCCC + 736427-736447
16S-5 GCTAAGAGATCAGCCTATGTCC 1208855-1208876
16S-3 TGGCAATCAGCGTCAGGTAATG + 1208356-1208377
a

Sequences in uppercase letters are derived from the genome sequences of H. pylori 26695 (34). Sequences introduced for cloning purposes are given in lowercase letters, and restriction recognition sequences are underlined.

b

Restriction recognition sites.

c

Nucleotide positions refer to the genome sequence of H. pylori 26695 (34).

The labeled probes were added to the hybridization solution, and hybridization was performed for 12 to 16 h at 42°C. The membrane was washed two times in prewarmed (42°C) wash solution I (6 M urea, 0.5× SSC, 0.4% SDS) for 20 min at 42°C and two times in wash solution II (2× SSC) at room temperature. For signal detection, the ECL detection system (ECL Direct Nucleic Acid Labeling and Detection system; Amersham Biosciences) and X-ray films (Konica Minolta) were used.

DNase I footprint analysis.

The recombinant N-terminally His6-tagged response regulator ArsR encoded on plasmid pQE-166 was overexpressed in E. coli M15[pREP4] (QIAGEN) and was purified by affinity chromatography on Ni2+-nitrilotriacetic acid agarose essentially as described previously (2). In vitro phosphorylation of His6-ArsR was performed as described by Dietz et al. (10). Plasmids pSL-PamiE, pSL-PSamiF, and pSL-ProcF were used for the generation of end-labeled DNA probes for DNase I footprint experiments. pSL-PSamiF was created by ligating a 318-bp EcoRI-BamHI fragment derived from the upstream region of the amiF gene, which was amplified with primer pair SamiF-5/PamiF-3 into pSL1180. The promoter DNA fragments were 5′-end labeled with [γ-32P] ATP and T4 polynucleotide kinase at one extremity and gel purified, and approximately 100,000 cpm of each probe was used for footprint experiments, which were performed essentially as described by Delany et al. (8). The binding reactions were performed for 20 min at room temperature in 50 μl binding buffer (50 mM Tris-HCl [pH 7.9], 40 mM KCl, 10 mM NaCl, 2 mM MgCl2, 0.1 mM CaCl2, 1 mM dithiothreitol).

Microarray accession numbers.

The microarray raw data were deposited in the NCBI Gene Expression Omnibus database (http://www.ncbi.nlm.nih.gov/geo) and are accessible with the accession number GSE4293.

RESULTS

Whole-genome transcriptional profiling of an ArsS-deficient mutant of H. pylori G27 exposed to pH 5.0.

Low pH has been demonstrated to be a signal triggering the autophosphorylation of the histidine kinase ArsS and the subsequent phosphorylation of the cognate response regulator ArsR (26). In order to define the complete regulon controlled by ArsR∼P, a transcriptome analysis was performed using a whole-genome microarray containing 1,649 PCR products generated with specific primer pairs derived from the genome sequences of H. pylori 26695 (34) and J99 (1) and comprising 98% of the coding sequences present in both genomes (17). RNA was extracted from two independent cultures of H. pylori G27 and G27/HP165::km, grown to an OD550 of 0.7, and then exposed to pH 5.0 for one hour. Cy5- and Cy3-labeled cDNA was prepared from these RNAs, and the differentially labeled cDNA pairs derived from the H. pylori wild type and arsS mutant were hybridized to four independent microarray slides, creating eight sets of hybridization data. A total of 109 genes was identified to be differentially expressed in the arsS mutant at pH 5.0 by using a signal ratio cut off of <0.5 and >2.0 (Table 1). Seventy-five genes were positively regulated by ArsR∼P, while the transcription of 34 genes was repressed by ArsR∼P. In accordance with our previous observation that the promoters of both ureA(PureA) and ureI(PureI) are positively regulated by ArsR∼P (26, 27), transcription of the complete urease gene cluster was reduced in the arsS mutant, as well as transcription of amiE and amiF, which encode additional ammonia-producing enzymes. Transcription of rocF, which encodes arginase, was found to be repressed by ArsR∼P. Furthermore, global transcriptional profiling confirmed that ArsR∼P acts as a negative autoregulator (10). Transcription of open reading frames (ORFs) HP0165 to HP0162 was reduced in the arsS mutant, due to the insertion of the kanamycin resistance cassette into ORF HP0165/HP0164 and its polar effect on expression of the downstream genes encoding δ-aminolevulinic acid dehydratase (HP0163) and a conserved hypothetical protein of unknown function (HP0162). The ArsR∼P-regulated genes were grouped into several categories according to their cellular functions and encode mainly proteins affecting the composition of the cell envelope, transport and binding proteins, detoxifying enzymes, and H. pylori-specific proteins of unknown function (Table 1).

To confirm the differential expression of the identified ORFs, seven genes not part of putative operon structures were selected and their transcription was monitored by Northern slot blot analysis or primer extension analysis performed with RNA extracted from the G27 wild type and the isogenic ArsS-deficient mutant exposed to pH 5.0. The selected genes encode two outer membrane proteins (omp6, HP0229; omp17, HP0725), the ferric dicitrate receptor FecA1 (HP0686), a glucose/galactose transporter (HP1174), alanine dehydrogenase (HP1398), mannitol dehyrogenase (HP1104), and alkyl hydroperoxide reductase (HP1563). As shown in Fig. 1 and in agreement with the results from the microarray analysis, in these experiments the transcription of omp6, omp17, fecA1, HP1174, and HP1398 was found to be derepressed in the absence of the histidine kinase ArsS, while transcription of HP1104 and HP1563 was reduced in the ArsS-deficient mutant.

FIG. 1.

FIG. 1.

Analysis of transcription of selected ArsR∼P target genes by slot blot Northern hybridization (A) and primer extension (B). Equal amounts of RNA extracted from H. pylori G27 (lane 1) and G27/HP165::km (lane 2) exposed to pH 5.0 for 1 h were used in the respective experiments performed with DNA probes or radioactively labeled oligonucleotides specific for the indicated ORFs. RNA slot blot analysis with a labeled probe specific for 16S RNA was performed as a control (not shown). +, presence of gene; −, absence of gene.

Acid-induced transcription of amiE and amiF is mediated directly by the binding of ArsR∼P to the promoters of the amidase genes.

From the genome organization of H. pylori it can be deduced that both amiE and amiF are monocistronically transcribed. To confirm the ArsR∼P-dependent transcription of the amidase genes observed in the transcriptome analyses, primer extension experiments with RNA extracted from H. pylori G27 and the isogenic ArsS-deficient mutant grown at neutral pH and exposed to pH 5.0 were performed. The transcriptional start site of amiE was mapped to position −44 with respect to the translational start codon (Fig. 2A). The upstream sequence revealed a box (TATAAA) at −10 with a high similarity to the E. coli promoter consensus sequence. The 5′ end of the amiF-specific transcript was mapped to position −48 with respect to the translational start codon, which corresponds to a −10 promoter element of the sequence TAGTAT (Fig. 2B), showing two mismatches from the consensus −10 promoter element from E. coli. In the G27 wild-type strain, transcription of both amiE and amiF was strongly induced at pH 5.0. At neutral pH, the detected amount of amiE-specific transcript was approximately fivefold lower in the ArsS-deficient mutant than in the G27 wild-type strain. Compared to the low basal level of ArsR∼P-independent expression detected in the mutant grown at pH 7.0, only a slight increase of amiE transcription was observed upon shifting the bacteria to pH 5.0 (Fig. 2A). No transcription of amiF was detected in the arsS mutant, irrespective of the pH of the growth medium (Fig. 2B). These data demonstrate that the pH-responsive transcription of the amidase genes is mediated mainly by the ArsRS two-component system. Since Fur was shown to act as a repressor of amiE expression (39), we investigated whether the acid-induced increase in amiE transcription in the ArsS-deficient mutant was due to the negative regulatory effect exerted by the NikR protein on fur transcription (7, 9, 40) by analyzing the amount of amiE-specific mRNA in an arsS nikR double mutant. As shown in Fig. 2A (compare lanes 3 and 5) transcription of amiE was similar in the ArsS-deficient mutant grown at pH 7.0 and the arsS nikR double mutant exposed to pH 5.0, suggesting a minor effect of the abovementioned repressor cascade on the acid-responsive regulation of amiE expression.

FIG. 2.

FIG. 2.

Analysis of transcription of the amidase genes in H. pylori G27 and the isogenic ArsS-deficient mutant G27/HP165::km grown at neutral pH and exposed to pH 5.0. (A) Primer extension experiments using the radiolabeled oligonucleotide amiE-PE were performed on equal amounts of RNAs extracted from H. pylori G27 (lanes 1 and 2) and G27/HP165::km (lanes 3 and 4) that were grown at neutral pH (lanes 1 and 3) or incubated at pH 5.0 for 1 h (lanes 2 and 4). In addition, RNA extracted from strain G27/ΔarsSΔnikR exposed to pH 5.0 (lane 5) was analyzed. (B) Primer extension analysis using the radiolabeled oligonucleotide amiF-PE was performed on RNAs prepared from strains G27 (lanes 1 and 2) and G27/HP165::km (lanes 3 and 4) grown at neutral pH (lanes 1 and 3) or exposed to pH 5.0 (lanes 2 and 4). The respective cDNAs are indicated by an arrow on the right. The sequences of the −10 element of the PamiE and PamiF promoter are given on the left. The sequencing ladders (lanes T, A, G, C) were obtained by annealing primers amiE-PE and amiF-PE to plasmids pSL-PamiE and pSL-PamiF, respectively. +, presence of gene; −, absence of gene.

Since besides arsR no genes encoding transcriptional regulators belong to the ArsR∼P regulon, it is likely that transcription of its constituents is controlled directly by the binding of ArsR∼P to the respective promoter regions. To define the binding sites of ArsR∼P in the promoter regions of amiE(PamiE) and amiF(PamiF), DNase I footprint analysis with the purified response regulator protein ArsR was performed. Figure 3 shows the results of a footprint experiment carried out on a 328-bp radioactively labeled DNA fragment derived from the upstream region of amiE. In the presence of 3 μM ArsR which was phosphorylated in vitro with acetylphosphate (ArsR∼P), a maximum region of 56 bp located immediately upstream of the −10 box of the PamiE promoter and ranging from position −14 to −69 with respect to the transcriptional start site of amiE was protected from DNase I digestion with a band of enhanced DNase I sensitivity appearing at position −21. In contrast, no footprinting could be detected when unphosphorylated ArsR protein was included in the binding reaction (data not shown).

FIG. 3.

FIG. 3.

Binding of ArsR∼P to the PamiE promoter. (A) DNase I footprint experiments were performed on a 328-bp EcoRI-BamHI fragment containing the PamiE promoter derived from plasmid pSL-PamiE, which was end labeled at the EcoRI terminus by adding increasing amounts of His6-ArsR phosphorylated in vitro by acetylphosphate. In lanes 2 to 6, His6-ArsR is present in concentrations of 0 (lane 2), 0.37 (lane 3), 0.75 (lane 4), 3.0 (lane 5), and 4.5 μM (lane 6). The numbers on the left indicate nucleotide positions with respect to the transcriptional start site, which is marked by an arrow. The solid bar on the right indicates the region of maximum DNase I protection. The arrow on the right indicates a band of hypersensitivity to DNase I digestion. Lane 1 contained a G+A sequence reaction mixture with the DNA probe used as a size marker (19). (B) Schematic representation of the PamiE promoter. The −10 promoter element is highlighted by black shading, and the transcriptional start site is indicated by an arrow above the double-stranded sequence. The gray box indicates the region with maximum protection from DNase I digestion by ArsR∼P binding to the PamiE promoter probe labeled at the EcoRI-terminus. The black bar below the sequence indicates the minimum size of the ArsR∼P binding site protected from DNaseI digestion. Numbers above the sequence indicate the nucleotide position with respect to the transcriptional start site (+1, not shown). The translational start codon of the amiE gene is shown in italics.

The results of footprint experiments carried out on a 318-bp DNA probe containing the PamiF promoter are presented in Fig. 4. Upon addition of in vitro phosphorylated ArsR at a concentration of 3 μM, a maximum region of 38 bp ranging from position −13 to −50 with respect to the transcriptional start site of amiF was protected from DNase I digestion. As already observed for the PamiE promoter, no binding to the upstream region of amiF could be detected when the footprint experiment was performed with unphosphorylated ArsR protein (data not shown). Therefore, we conclude that acid-responsive transcription of the amidase genes is regulated by the ArsRS two-component system via the direct binding of ArsR∼P to extended regions overlapping the −35 region of the PamiE and PamiF promoters.

FIG. 4.

FIG. 4.

Binding of ArsR∼P to the PamiF promoter. (A) DNase I footprint experiments were performed on a 318-bp EcoRI-BamHI fragment containing the PamiF promoter, which was end labeled at the EcoRI terminus by adding increasing amounts of His6-ArsR phosphorylated in vitro by acetylphosphate. In lanes 2 to 7, His6-ArsR is present in concentrations of 0 (lane 2), 0.37 (lane 3), 0.75 (lane 4), 1.5 (lane 5), 3.0 (lane 6), and 4.5 μM (lane 7). The numbers on the left indicate nucleotide positions with respect to the transcriptional start site, which is marked by an arrow. The solid bar on the right indicates the maximum region of DNase I protection. Lane 1 contained a G+A sequence reaction mixture with the DNA probe used as a size marker (19). (B) Schematic representation of the PamiF promoter. The −10 promoter element is highlighted by black shading, and the transcriptional start site is indicated by an arrow above the double-stranded sequence. The gray box indicates the region with maximum protection from DNase I digestion by response regulator binding to the PamiF promoter probe labeled at the EcoRI terminus. The black bar below the sequence indicates the minimum size of the ArsR∼P binding site protected from DNase I digestion. Numbers above the sequence indicate the nucleotide position with respect to the transcriptional start site (+1, not shown). The translational start codon of the amiF gene is given in italics.

ArsR∼P-dependent repression of rocF and HP0682 is not responsive to low pH.

Global transcriptional profiling revealed that at pH 5.0 transcription of the arginase gene rocF is derepressed in the ArsS-deficient mutant of H. pylori G27. However, in none of the previous transcriptome studies analyzing pH-responsive gene regulation in H. pylori (6, 21, 44) was expression of rocF reported to be repressed at low pH. To further investigate these conflicting observations, transcription of rocF was analyzed by primer extension experiments carried out on RNA extracted from H. pylori G27 and the isogenic ArsS-deficient mutant grown at neutral pH and exposed to pH 5.0. As shown in Fig. 5A, the 5′ end of the rocF-specific mRNA was mapped to position −29 with respect to the translational start site of the rocF gene, corresponding to a −10 promoter hexamer of the sequence TAGAAT. An almost equal amount of rocF-specific transcript was detected in the G27 wild type irrespective of whether the bacteria were cultivated at pH 7.0 or exposed to pH 5.0. In the arsS mutant, transcription of rocF was strongly derepressed at both growth conditions; however, an approximately threefold-higher amount of rocF-specific transcript was detected at pH 5.0. This result demonstrated that the promoter of the rocF gene (ProcF) is repressed by ArsR∼P; however, growth of H. pylori G27 at neutral or acidic pH, supposed to cause alterations of the cellular level of phosphorylated ArsR, did not result in differential transcription of rocF. To test whether cultivation of H. pylori at mildly basic pH might relieve the ArsR∼P-dependent repression of ProcF, transcription of the rocF gene was analyzed by primer extension performed on RNA extracted from strain G27, which was exposed to pH 8.0 or pH 8.5 for 1 h. It has been reported previously that a pH of 8.0 still allows growth of H. pylori in the absence of urea (30). No significant change in the amount of rocF-specific transcripts was observed under these conditions compared to that of strain G27 grown at neutral pH (data not shown).

FIG. 5.

FIG. 5.

Analysis of transcription of rocF and ORF HP0682 in H. pylori G27 and G27/HP165::km. (A) Primer extension experiments using the radiolabeled oligonucleotide rocF-PE were performed on RNAs extracted from the wild-type strain G27 grown at pH 7.0 and pH 5.0 (lanes 1 and 2) and the ArsS-deficient mutant G27/HP165::km grown at neutral pH and pH 5.0 (lanes 3 and 4). The elongated primer products are indicated by an arrow on the right. The sequences of the −10 element of the ProcF promoter is given on the left. The sequencing ladder (lanes T, A, G, C) was obtained by annealing primer rocF-PE to plasmid pSL-ProcF. (B) Primer extension experiments using the radiolabeled oligonucleotide 682-PE were performed on RNAs extracted from strains G27 (lanes 1 and 2) and G27/HP165::km (lanes 3 and 4) grown at pH 7.0 (lanes 1 and 3) and exposed to pH 5.0 for 1 h (lanes 2 and 4). The cDNAs specific for ORF hp682 are indicated by an arrow on the right. +, presence of gene; −, absence of gene.

In our transcriptome analysis, the paralogous ORFs HP0682 and HP1289, which encode H. pylori-specific proteins of unknown function, showed the highest ratio of ArsR∼P-dependent repression. Both genes are organized in bicistronic transcriptional units with the ORFs HP0681 and HP1288, respectively, and had been previously shown to be negatively regulated by ArsR∼P (15). However, as is the case with rocF, pH-dependent repression of these genes has not been noted in previous studies (6, 21, 44). The transcriptional start site of ORF HP0682 was mapped by Forsyth et al. (15) to position −6 with respect to the annotated translational start codon, corresponding to a TATAAA −10-box hexamer. When the transcription of ORF HP0682 was compared in the G27 wild-type strain and the ArsS-deficient mutant grown at neutral pH or exposed to pH 5.0, respectively, the same expression profile as observed for the rocF gene was detected (Fig. 5B). Similarly, no significant change in the amount of HP0682-specific transcript was detected when the bacteria were shifted from pH 7.0 to pH 8.0 or pH 8.5 (data not shown). Therefore, we conclude that the transcription of a subset of genes which are clearly regulated by ArsR∼P is not responsive to pH changes in a range from pH 5.0 to 8.5.

ArsR∼P binds to a high-affinity binding site overlapping the ProcF promoter.

To investigate whether transcription of rocF is regulated directly by ArsR∼P, DNase I footprint experiments were carried out on a 365-bp DNA probe comprising the upstream region of the rocF gene (Fig. 6). In the presence of 0.37 to 0.75 μM ArsR which was phosphorylated in vitro with acetylphosphate, a region protected from DNase I digestion, ranging from position −6 to −67 with respect to the transcriptional start site and overlapping the −10 promoter box, became visible. When the binding reactions were performed with unphosphorylated ArsR protein, footprinting of this region could be observed at a protein concentration of 4.5 μM (data not shown). Therefore, the results of the footprint experiments demonstrate that transcription of the rocF gene is repressed by ArsR∼P, which binds with high affinity to a region overlapping the −10 box of the ProcF promoter.

FIG. 6.

FIG. 6.

Binding of ArsR∼P to the ProcF promoter. (A) DNase I footprint experiments were carried out on the ProcF promoter fragment labeled at the BamHI terminus. In lanes 2 to 9, His6-ArsR was added in concentrations of 0 (lane 2), 0.09 (lane 3), 0.18 (lane 4), 0.37 (lane 5), 0.75 (lane 6), 1.5 (lane 7), 3.0 (lane 8) and 4.5 μM (lane 9). The numbers on the left indicate nucleotide positions with respect to the transcriptional start site, which is marked by an arrow. The solid bar on the right indicates the region of DNase I protection. Lane 1 contained a G+A sequence reaction mixture with the DNA probe used as a size marker (19). (B) Schematic representation of the ProcF promoter. The −10 promoter element is highlighted by black shading, and the transcriptional start site is indicated by an arrow above the double-stranded sequence. The gray box indicates the region protected from DNase I digestion by response regulator binding to the ProcF promoter probe. The black bar below the sequence highlights the complete ArsR∼P binding site, which covers the −10 promoter element. Numbers above the sequence indicate the nucleotide position with respect to the transcriptional start site (+1, not shown). The translational start codon of the rocF gene is given in italics.

DISCUSSION

H. pylori is a highly specialized bacterium which, to our knowledge, colonizes the human stomach as its unique habitat and, therefore, is endowed with highly efficient acid acclimation mechanisms distinguishing it from other neutralophilic bacteria. Due to the continuous alteration between starvation and digestive phases following food intake, H. pylori is likely to encounter considerable pH fluctuations within its niche, posing the necessity of efficient modulation of the pH-adaptive response. Three independent global transcriptome studies have been performed so far in order to completely define the pH-responsive stimulon comprising the genes involved in acid acclimation. However, there was only a rather low overlap among the data raised in the individual studies, reflected by a total of 429 genes reported to be differentially expressed upon acid exposure added from the three individual data sets comprising between 101 and 279 genes (6, 21, 44). This discrepancy might be explained by differences in the experimental setups, including the particular H. pylori strains analyzed. The ArsRS two-component system has been identified previously as an important regulator of the acid response, mediating the low-pH induction of urease expression upon phosphorylation of the response regulator ArsR (26, 27). Therefore, we intended to characterize the acid-responsive regulon controlled by ArsR∼P by transcriptional profiling of H. pylori G27 and an isogenic mutant lacking the histidine kinase gene arsS. The level of transcription of 109 genes differed more than twofold between the two strains, with 75 genes being positively regulated by ArsR∼P and 34 genes being repressed by ArsR∼P. Acid-responsive regulation of 64 of these genes (48 induced, 16 repressed) has previously been reported on the basis of global transcriptome analysis (6, 21, 44). From the 109 genes supposed to be regulated by ArsR∼P, 45 are predicted to be monocistronically transcribed (pH regulation of 25 of these genes was confirmed in the previous array studies) according to the genome sequences of H. pylori 26695 and J99 (1, 34), while the remaining genes are members of 33 predicted transcriptional units. In 10 of these transcriptional units, transcription of all the members of the respective operon was found to be altered in the ArsS-deficient mutant, and in 5 predicted operons comprising three or more genes, all but one of their members were found to be differentially regulated. Most convincingly, in accordance with our previous observation that the promoters located upstream of ureAB and ureIEFGH are controlled by ArsR∼P (26, 27), in our microarray analysis transcription of the complete urease gene cluster was found to be altered, which was not observed in previous global transcriptome studies (6, 21, 44). Differential expression of 11 genes identified in the microarray analysis was confirmed by Northern and primer extension analysis. pH-responsive or ArsR∼P-dependent regulation of four of these genes had not been reported so far. The highest ratio of transcriptional induction by ArsR∼P showed ORFs HP0081 (25-fold) and HP0079 (17-fold), encoding a hypothetical protein and a secreted protein supposed to be involved in flagellar motility, respectively, and the amidase gene amiE (16-fold) and ORF HP1432 (15-fold), encoding a histidine-rich protein predicted to be involved in nickel binding. amiF and ORF HP1427, encoding another histidine-rich protein, exhibited an approximately threefold-lower transcriptional induction than their respective paralogous genes. The highest ratio of ArsR∼P-dependent repression was observed with ORFs HP0682 and HP1289 (16- and 22-fold, respectively), encoding hypothetical proteins of unknown function (see below). Induction of amidase expression in response to low pH has been observed previously (6, 21, 40, 44) and suggests an important role for these ammonia-producing enzymes in the acid adaptation of H. pylori. However, the amidase genes are not essential for colonization in the mouse model of H. pylori infection (5). Here we demonstrate that expression of the three major ammonia-producing pathways—urease, AmiE, and AmiF—is under the control of the ArsRS two-component system. Interestingly, besides the nickel-binding proteins HP1432 and HP1427, expression of a number of detoxifying enzymes (thiol peroxidase, alkyl hydroperoxide reductase, catalase, and superoxide dismutase) was induced by ArsR∼P. This might reflect protection against the toxic effects caused by high intracellular concentrations of the bivalent metal ions Ni2+ and Fe2+, whose solubility and therefore bioavailability increase under acidic conditions. In particular, Fe2+ is known to be involved in the triggering of radical reactions whose products need to be detoxified. In this context, it should be emphasized that low pH was shown to induce the expression of NikR, which is a repressor of fur transcription (9, 40). As a consequence, the amount of Fur protein present under acidic conditions might be too low to allow efficient repression of the iron uptake systems, increasing the necessity of protection against the toxic effects of Fe2+. ArsR∼P is also involved in changing the composition of the outer membrane in response to low pH by regulating the expression of several outer membrane proteins and of enzymes involved in lipopolysaccharide (LPS) biosynthesis. Although an intriguing speculation, the pI of these outer membrane proteins and therefore their buffering capacity seem not to be correlated with their predominant expression at neutral or acidic pH. Changes in LPS composition upon exposure of H. pylori to acidic conditions are known to occur (23) and might involve ORFs HP1191 and HP1105, regulated by ArsR∼P. A number of transport-associated proteins were also found to be both positively and negatively regulated by ArsR∼P, possibly reflecting adaptation to changes in the supply of nutrients, which might be associated with the fluctuations in gastric pH. However, about one-third of the genes regulated by ArsR∼P encode hypothetical proteins whose roles in the acid acclimation of H. pylori remain to be elucidated.

As observed previously, the ArsR∼P regulon and the Ni2+-responsive NikR regulon overlap to some extent and have several genes involved in nickel homeostasis in common, such as the urease genes and HP1432 and HP1427, whose products are involved in nickel storage (7, 38). No significant effect of ArsR∼P on the transcription of nikR was observed, indicating that the acid-induced increase in NikR expression (40) is mediated by an additional pH-responsive regulatory system. In the transcriptome studies performed by Merrell et al. (21), Wen et al. (44), and Bury-Moné et al. (6), four genes (HP0624, HP1050, HP1440, and HP1457) were consistently found to be differentially expressed at low pH versus neutral pH, but no influence of ArsR∼P on their transcription was detected. Therefore, it is likely that such systems are present in H. pylori despite its general paucity of transcriptional regulators. Interestingly, of the 29 genes in common between our study and the transcriptome analysis performed by Bury-Moné et al. (6), 20 were reported to have lost their pH-responsive transcription profile in an H. pylori double mutant deficient in both NikR and Fur, suggesting a prominent role for these proteins as coregulators in the control of pH-responsive gene expression. The genes aberrantly transcribed in the nikR fur mutant at pH 5.0 included ureA, amiE, omp16, omp17, and arsR itself, while pH-responsive transcription of HP1432, amiF and omp3 was retained in the double mutant (6). Gancz et al. (16) reported recently that several members of the ArsR∼P regulon were aberrantly transcribed at pH 5.0 in a Fur-deficient mutant that includes amiE, HP1432, and arsR (16). However, we could not detect any significant effect on ureA transcription in a NikR-deficient mutant grown in standard broth at pH 7.0 and pH 5.0 (27), and no influence of Fur on pH-responsive transcription of the urease genes was observed in the study of Gancz et al. (16). Furthermore, no indications for metal-responsive regulation of arsR by Fur and NikR were obtained in recent macroarray analyses (7, 12). These partially conflicting data point out the complex interplay between ArsR and the metal-dependent regulators, which must be dissected in future studies. Note also that we and Bury-Moné et al. (6), in accordance with the mapping of an ArsR∼P binding site downstream of the promoter of the arsR gene (10), detected repression of arsR transcription at pH 5.0, while an increase in arsR transcription was reported in the transcriptome study of Wen et al. (44).

The finding that ArsR∼P footprinted the PamiE and PamiF promoters clearly demonstrated that the pH-responsive regulation of the amidase genes is mediated directly by the ArsRS two-component system (Fig. 3 and 4). While in the case of amiE we detected a basal level of ArsR∼P-independent transcription, as was previously observed for the ureA and ureI genes (26, 27), transcription from the PamiF promoter absolutely required ArsR∼P, since no amiF-specific transcript could be detected in the ArsS-deficient mutant, irrespective of the pH of the growth medium (Fig. 2). In the arsS mutant, we found a slight increase in amiE transcription at low pH, which might be caused by a repressor cascade affecting the PamiE promoter, which involves the metal-dependent regulators NikR and Fur (39, 40). Although no evidence for a role of NikR in transcriptional regulation of the amiF gene has been obtained (6, 7) and our data suggest a role for ArsR∼P as the prominent regulator of pH-responsive transcription of the amidase genes, the pH-responsive induction of the enzymatic activity of both AmiF and AmiE is clearly reduced in a nikR mutant (6, 40). This might suggest that a gene product of the NikR regulon contributes to the enzymatic activity of the amidases or is involved in providing their cellular substrates, which are unknown so far.

The regions protected from DNase I digestion by the binding of ArsR∼P to the PamiE and PamiF promoters spanned 56 bp and 38 bp, respectively, and overlapped with the −35 promoter elements. Extended binding regions for ArsR∼P have also been found in the PureA and PureI promoters (27). These observations might suggest that two consecutive ArsR∼P dimers are binding to the promoters of the respective genes or that the active response regulator forms a tetrameric protein complex. As noticed for the PamiE and PamiF promoters in this study, the binding regions of ArsR∼P mapped in the PureA and PureI promoters were not recognized by the unphosphorylated ArsR protein in DNase I footprint experiments (27). This similarity in promoter characteristics might be correlated with similar kinetics of pH-responsive transcription of the amidase and ureA and ureI genes. According to the microarray study of Merrell et al. (21), these genes reached their maximum level of transcription 90 min after the shift to low pH. In contrast, other ArsR∼P-regulated genes whose promoters harbor short binding sites for ArsR∼P (10) were maximally transcribed 15 min after the pH shift (21).

Most surprisingly, in this study we identified the rocF gene and ORF HP0682 as members of the ArsR∼P regulon which are not responsive to pH changes in a range of 5.0 to 8.5. As ArsR∼P footprinted the ProcF promoter (Fig. 6), it is likely that ArsS-dependent repression of the rocF gene is mediated by ArsR∼P and not by an alternative transcriptional regulator interacting with ArsS. Interestingly, a region of 62 bp overlapping the −10 box of the ProcF promoter exhibited the highest affinity for ArsR∼P noted so far in DNase I footprint experiments. While this high-affinity binding site interacted with ArsR∼P present in concentrations of 0.37 or 0.75 μM, depending on the particular experiment, binding to PureA, PureI, PamiE, PamiF, and the promoters of the ArsR∼P regulated ORFs HP1408 and HP0119 required ArsR∼P in concentrations of 1.5 μM and above (10, 27). Since from the comparison of the transcription of ArsR∼P-dependent genes in the G27 wild type and the ArsS-deficient mutant at pH 7.0 it is clear that at neutral pH ArsS autophosphorylates to some extent (10, 15), we hypothesize that the low cellular concentration of ArsR∼P already present at pH 7.0 is sufficient to repress the rocF gene. Increasing the pH of the surrounding medium up to pH 8.5 did not result in derepression of rocF. In the ArsS-deficient mutant, we observed an approximately threefold increase in rocF transcription at pH 5.0 compared to that at pH 7.0, suggesting pH-responsive positive regulation of rocF by an unknown mechanism which comes into play when the ArsR∼P-dependent repression of rocF is relieved. In this context, note that Wen et al. (44) observed a 1.7-fold increase in rocF transcription when H. pylori 26695 was exposed to pH 5.5. So far, the biological significance of this complex mode of regulation of rocF remains obscure. In addition to ORFs HP1288 and HP1289 being paralogs of HP0682 and HP0681, several genes, including HP0081, and HP1398, showed a rather high ratio of differential expression in the ArsS-deficient mutant at pH 5.0 but were not detected in the transcriptome studies investigating pH-responsive gene regulation (6, 21, 44). Therefore, for the moment it cannot be ruled out that these genes also belong to the class of ArsR∼P-dependent genes which do not exhibit a pH-responsive transcription profile.

In conclusion, we have shown that the major ammonia-producing pathways which are central to the acid resistance of H. pylori are regulated in response to low pH by the ArsRS two-component system. Several members of the ArsR∼P regulon defined here are under the additional control of the metal-dependent regulators NikR and Fur, indicating a complex regulatory interplay in the pH-responsive control of transcription. This complexity is further underlined by the identification of ArsR∼P-dependent genes whose transcription is not responsive to low pH.

Supplementary Material

[Supplemental material]

Acknowledgments

We acknowledge Roy Gross for helpful discussions and critically reading the manuscript. We thank Lucía Martín Gozalo for help with cloning.

This work was supported by the Competence Network PathoGenoMik of the German Federal Ministry of Education and Research and by a grant from the Deutsche Forschungsgemeinschaft (BE 1543/6-1).

Footnotes

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

REFERENCES

  • 1.Alm, R. A., L.-S. L. Ling, D. T. Moir, B. L. King, E. D. Brown, P. C. Doig, D. R. Smith, B. Noonan, B. C. Guild, B. L. deJonge, G. Carmel, P. J. Tummino, A. Caruso, M. Uria-Nickelsen, D. M. Mills, C. Ives, R. Gibson, D. Merberg, S. D. Mills, Q. Jiang, D. E. Taylor, G. F. Vovis, and T. J. Trust. 1999. Genomic-sequence comparison of two unrelated isolates of the human gastric pathogen Helicobacter pylori. Nature 397:176-180. [DOI] [PubMed] [Google Scholar]
  • 2.Beier, D., and R. Frank. 2000. Molecular characterization of two-component systems of Helicobacter pylori. J. Bacteriol. 182:2068-2076. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 3.Bijlsma, J. J. E., B. Waidner, A. H. M. van Vliet, N. J. Hughes, S. Häg, S. Bereswill, D. J. Kelly, C. M. J. E. Vandenbroucke-Grauls, M. Kist, and J. G. Kusters. 2002. The Helicobacter pylori homologue of the ferric uptake regulator is involved in acid resistance. Infect. Immun. 70:606-611. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 4.Blaser, M. J. 1992. Helicobacter pylori: its role in disease. Clin. Infect. Dis. 15:386-391. [DOI] [PubMed] [Google Scholar]
  • 5.Bury-Moné, S., S. Skouloubris, C. Dauga, J.-M. Thiberge, D. Dailidiene, D. E. Berg, A. Labigne, and H. De Reuse. 2003. Presence of active aliphatic amidases in Helicobacter species able to colonize the stomach. Infect. Immun. 71:5613-5622. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 6.Bury-Moné, S., J.-M. Thiberge, M. Contreras, A. Maitournam, A. Labigne, and H. De Reuse. 2004. Responsiveness to acidity via metal ion regulators mediates virulence in the gastric pathogen Helicobacter pylori. Mol. Microbiol. 53:623-638. [DOI] [PubMed] [Google Scholar]
  • 7.Contreras, M., J.-M. Thiberge, M.-A. Mandrand-Berthelot, and A. Labigne. 2003. Characterization of the roles of NikR, a nickel-responsive pleiotropic autoregulator of Helicobacter pylori. Mol. Microbiol. 49:947-963. [DOI] [PubMed] [Google Scholar]
  • 8.Delany, I., G. Spohn, R. Rappuoli, and V. Scarlato. 2001. The Fur repressor controls transcription of iron-activated and -repressed genes in Helicobacter pylori. Mol. Microbiol. 42:1297-1309. [DOI] [PubMed] [Google Scholar]
  • 9.Delany, I., R. Ieva, A. Soragni, M. Hilleringmann, R. Rappuoli, and V. Scarlato. 2005. In vitro analysis of protein-operator interactions of the NikR and Fur metal-responsive regulators of coregulated genes in Helicobacter pylori. J. Bacteriol. 187:7703-7715. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 10.Dietz, P., G. Gerlach, and D. Beier. 2002. Identification of target genes regulated by the two-component system HP166-HP165 of Helicobacter pylori. J. Bacteriol. 184:350-362. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 11.Eaton, K. A., and S. Krakowka. 1994. Effect of gastric pH on urease-dependent colonization of gnotobiotic piglets by Helicobacter pylori. Infect. Immun. 62:3604-3607. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 12.Ernst, F. D., S. Bereswill, B. Waidner, J. Stoof, U. Mader, J. G. Kusters, E. J. Kuipers, M. Kist, A. H. M. van Vliet, and G. Homuth. 2005. Transcriptional profiling of Helicobacter pylori Fur- and iron-regulated gene expression. Microbiology 151:533-546. [DOI] [PubMed] [Google Scholar]
  • 13.Ernst, F. D., E. J. Kuipers, A. Heijens, R. Sarwari, J. Stoof, C. W. Penn, J. G. Kusters, and A. H. M. van Vliet. 2005. The nickel-responsive regulator NikR controls activation and repression of gene transcription in Helicobacter pylori. Infect. Immun. 73:7252-7258. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 14.Ferrero, R. L., V. Cussac, P. Courcoux, and A. Labigne. 1992. Construction of isogenic urease-negative mutants of Helicobacter pylori by allelic exchange. J. Bacteriol. 174:4212-4217. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 15.Forsyth, M. H., P. Cao, P. P. Garcia, J. D. Hall, and T. L. Cover. 2002. Genome-wide transcriptional profiling in a histidine kinase mutant of Helicobacter pylori identifies members of a regulon. J. Bacteriol. 184:4630-4635. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 16.Gancz, H., S. Censini, and D. S. Merrell. 2006. Iron and pH homeostasis intersect at the level of Fur regulation in the gastric pathogen Helicobacter pylori. Infect. Immun. 74:602-614. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 17.Gressmann, H., B. Linz, R. Ghai, K.-P. Pleissner, R. Schlapbach, Y. Yamaoka, C. Kraft, S. Suerbaum, T. F. Meyer, and M. Achtmann. 2005. Gain and loss of multiple genes during the evolution of Helicobacter pylori. PLoS Genet. 1:e43. [Online.] [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 18.Marcus, E. A., A. P. Moshfegh, G. Sachs, and D. R. Scott. 2005. The periplasmic α-carbonic anhydrase activity of Helicobacter pylori is essential for acid acclimation. J. Bacteriol. 187:729-738. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 19.Maxam, A. M., and W. Gilbert. 1977. A new method for sequencing DNA. Proc. Natl. Acad. Sci. USA 74:560-564. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 20.McGee, D. J., F. J. Radcliff, G. L. Mendz, R. L. Ferrero, and H. L. T. Mobley. 1999. Helicobacter pylori rocF is required for arginase activity and acid protection in vitro but is not essential for colonization of mice or for urease activity. J. Bacteriol. 181:7314-7322. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 21.Merrell, D. S., M. L. Goodrich, G. Otto, L. S. Tompkins, and S. Falkow. 2003. pH-regulated gene expression of the gastric pathogen Helicobacter pylori. Infect. Immun. 71:3529-3539. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 22.Mollenhauer-Rektorschek, M., G. Hanauer, G. Sachs, and K. Melchers. 2002. Expression of UreI is required for intragastric transit and colonization of gerbil gastric mucosa by Helicobacter pylori. Res. Microbiol. 153:659-666. [DOI] [PubMed] [Google Scholar]
  • 23.Moran, A. P., Y. A. Knirel, S. N. Senchenkova, G. Widmalm, S. O. Hynes, and P.-E. Jansson. 2002. Phenotypic variation in molecular mimicry between Helicobacter pylori lipopolysaccharides and human gastric epithelial cell surface glycoforms. J. Biol. Chem. 277:5785-5795. [DOI] [PubMed] [Google Scholar]
  • 24.Panthel, K., P. Dietz, R. Haas, and D. Beier. 2003. Two-component systems of Helicobacter pylori contribute to virulence in a mouse infection model. Infect. Immun. 71:5381-5385. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 25.Peterson, W. L. 1991. Helicobacter pylori and peptic ulcer disease. N. Engl. J. Med. 324:1043-1048. [DOI] [PubMed] [Google Scholar]
  • 26.Pflock, M., P. Dietz, J. Schär, and D. Beier. 2004. Genetic evidence for histidine kinase HP165 being an acid sensor of Helicobacter pylori. FEMS Microbiol. Lett. 234:51-61. [DOI] [PubMed] [Google Scholar]
  • 27.Pflock, M., S. Kennard, I. Delany, V. Scarlato, and D. Beier. 2005. Acid-induced activation of the urease promoters is mediated directly by the ArsRS two-component system of Helicobacter pylori. Infect. Immun. 73:6437-6445. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 28.Scott, D. R., E. A. Marcus, D. L. Weeks, A. Lee, K. Melchers, and G. Sachs. 2000. Expression of the Helicobacter pylori ureI gene is required for acidic pH activation of cytoplasmic urease. Infect. Immun. 68:470-477. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 29.Schär, J., A. Sickmann, and D. Beier. 2005. Phosphorylation-independent activity of atypical response regulators of Helicobacter pylori. J. Bacteriol. 187:3100-3109. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 30.Sidebotham, R. L., M. L. Worku, Q. N. Karim, N. K. Dhir, and J. H. Baron. 2003. How Helicobacter pylori urease may affect external pH and influence growth and motility in the mucus environment: evidence from in vitro studies. Eur. J. Gastroenterol. Hepatol. 15:395-401. [DOI] [PubMed] [Google Scholar]
  • 31.Skouloubris, S., J.-M. Thiberge, A. Labigne, and H. De Reuse. 1998. The Helicobacter pylori UreI protein is not involved in urease activity but is essential for bacterial survival in vivo. Infect. Immun. 66:4517-4521. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 32.Skouloubris, S., A. Labigne, and H. De Reuse. 2001. The AmiE aliphatic amidase and AmiF formamidase of Helicobacter pylori: natural evolution of two enzyme paralogues. Mol. Microbiol. 40:596-609. [DOI] [PubMed] [Google Scholar]
  • 33.Stingl, K., E.-M. Uhlemann, G. Deckers-Hebestreit, R. Schmid, E. P. Bakker, and K. Altendorf. 2001. Prolonged survival and cytoplasmic pH homeostasis of Helicobacter pylori at pH 1. Infect. Immun. 69:1178-1180. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 34.Tomb, J.-F., O. White, A. R. Kerlavage, R. Clayton, G. G. Sutton, R. D. Fleischmann, K. A. Ketchum, H. P. Klenk, S. Gill, B. A. Dougherty, K. Nelson, J. Quackenbush, L. Zhou, E. F. Kirkness, S. Peterson, B. Loftus, D. Richardson, R. Dodson, H. G. Khalak, A. Glodek, K. McKenney, L. M. Fitzegerald, N. Lee, M. D. Adams, E. K. Hickey, D. E. Berg, J. D. Gocayne, T. R. Utterback, J. D. Peterson, J. M. Kelley, M. D. Cotton, J. M. Weidman, C. Fujii, C. Bowman, L. Watthey, E. Wallin, W. S. Hayes, M. Borodovsky, P. D. Karp, H. O. Smith, C. M. Fraser, and J. C. Venter. 1997. The complete genome sequence of the gastric pathogen Helicobacter pylori. Nature 388:539-547. [DOI] [PubMed] [Google Scholar]
  • 35.Tusher, V. G., R. Tibshirani, and G. Chu. 2001. Significance analysis of microarrays applied to the ionizing radiation response. Proc. Natl. Acad. Sci. USA 98:5116-5121. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 36.Tsuda, M., M. Karita, M. G. Morshed, K. Okita, and T. Nakazawa. 1994. A urease-negative mutant of Helicobacter pylori constructed by allelic exchange mutagenesis lacks the ability to colonize the nude mouse stomach. Infect. Immun. 62:3586-3589. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 37.Uemura, N. S., S. Okamoto, S. Yamamoto, N. Matsumura, S. Yamaguchi, M. Yamakido, K. Taniyama, N. Sasaki, and R. J. Schlemper. 2001. Helicobacter pylori infection and the development of gastric cancer. N. Engl. J. Med. 345:784-789. [DOI] [PubMed] [Google Scholar]
  • 38.van Vliet, A. H. M., S. W. Poppelaars, B. J. Davies, J. Stoof, S. Bereswill, M. Kist, C. W. Penn, E. J. Kuipers, and J. G. Kusters. 2002. NikR mediates nickel-responsive transcriptional induction of urease expression in Helicobacter pylori. Infect. Immun. 70:2846-2852. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 39.van Vliet, A. H. M., J. Stoof, S. W. Poppelaars, S. Bereswill, G. Homuth, M. Kist, E. J. Kuipers, and J. G. Kusters. 2003. Differential regulation of amidase- and formamidase-mediated ammonia production by the Helicobacter pylori Fur repressor. J. Biol. Chem. 278:9052-9057. [DOI] [PubMed] [Google Scholar]
  • 40.van Vliet, A. H. M., E. J. Kuipers, J. Stoof, S. W. Poppelaars, and J. G. Kusters. 2004. Acid-responsive gene induction of ammonia-producing enzymes in Helicobacter pylori is mediated via a metal-responsive repressor cascade. Infect. Immun. 72:766-773. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 41.van Vliet, A. H. M., F. D. Ernst, and J. G. Kusters. 2004. NikR-mediated regulation of Helicobacter pylori acid adaptation. Trends Microbiol. 12:489-494. [DOI] [PubMed] [Google Scholar]
  • 42.Wang, Y., and D. E. Taylor. 1990. Chloramphenicol resistance in Campylobacter coli: nucleotide sequence, expression, and cloning vector construction. Gene 94:23-28. [DOI] [PubMed] [Google Scholar]
  • 43.Weeks, D. L., S. Eskandari, D. R. Scott, and G. Sachs. 2000. A H+-gated urea channel: the link between Helicobacter pylori urease and gastric colonization. Science 287:482-485. [DOI] [PubMed] [Google Scholar]
  • 44.Wen, Y., E. A. Marcus, U. Matrubutham, M. A. Gleeson, D. R. Scott, and G. Sachs. 2003. Acid-adaptive genes of Helicobacter pylori. Infect. Immun. 71:5921-5939. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 45.Xiang, Z., S. Censini, P. F. Bayeli, J. L. Telford, N. Figura, R. Rappuoli, and A. Covacci. 1995. Analysis of expression of CagA and VacA virulence factors in 43 strains of Helicobacter pylori reveals that clinical isolates can be divided into two major types and that CagA is not necessary for expression of the vacuolating cytotoxin. Infect. Immun. 63:94-98. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 46.Yang, J. H., S. Dudoit, P. Luu, D. M. Lin, V. Peng, J. Ngai, and T. P. Speed. 2002. Normalization for cDNA microarray data: a robuste composite method addressing single and multiple slide systematic variation. Nucleic Acid Res. 30:e15. [Online.] [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