Urease of Enterohemorrhagic Escherichia coli

8 downloads 0 Views 184KB Size Report
Enzymatic assays confirmed a difference in urease expression of cloned EHEC ure clusters in E. coli MC3100 fur. Likewise, interruption of fur in O157:H7 isolate ...
INFECTION AND IMMUNITY, Feb. 2002, p. 1027–1031 0019-9567/02/$04.00⫹0 DOI: 10.1128/IAI.70.2.1027–1031.2002 Copyright © 2002, American Society for Microbiology. All Rights Reserved.

Vol. 70, No. 2

Urease of Enterohemorrhagic Escherichia coli: Evidence for Regulation by Fur and a trans-Acting Factor Susan R. Heimer,1 Rod A. Welch,2 Nicole T. Perna,3 György Pósfai,4 Peter S. Evans,3 James B. Kaper,1 Fred R. Blattner,3 and Harry L. T. Mobley1* Department of Microbiology and Immunology, University of Maryland School of Medicine, Baltimore, Maryland 212011; Department of Medical Microbiology and Immunology2 and Laboratory of Genetics,3 University of Wisconsin, Madison, Wisconsin 53706; and Institute of Biochemistry, Biological Research Center, Hungarian Academy of Science, Szeged, Hungary 67014 Received 29 June 2001/Returned for modification 8 October 2001/Accepted 12 November 2001

Recent genomic analyses of Escherichia coli O157:H7 strain EDL933 revealed two loci encoding urease gene homologues (ureDABCEFG), which are absent in nonpathogenic E. coli strain K-12. This report demonstrates that the cloned EDL933 ure gene cluster is capable of synthesizing urease in an E. coli DH5␣ background. However, when the gene fragment is transformed back into the native EDL933 background, the enzymatic activity of the cloned determinants is undetectable. We speculate that an unidentified trans-acting factor in enterohemorrhagic E. coli (EHEC) is responsible for this regulation of ure expression. In addition, Fur-like recognition sites are present in three independent O157:H7 isolates upstream of ureD and ureA. Enzymatic assays confirmed a difference in urease expression of cloned EHEC ure clusters in E. coli MC3100⌬fur. Likewise, interruption of fur in O157:H7 isolate IN1 significantly diminished urease activity. We propose that, similar to the function of Fur in regulating the acid response of Salmonella enterica serovar Typhimurium, it modulates urease expression in EHEC, perhaps contributing to the acid tolerance of the organism.

Enterohemorrhagic Escherichia coli (EHEC) O157:H7 is a highly infectious pathogen that is responsible for a large number of food-borne outbreaks (25, 46). Common symptoms in EHEC infections include nonbloody and bloody colitis as well as hemolytic-uremic syndrome (19, 25). As part of a comparative study of E. coli genomes, the chromosome of EHEC strain EDL933 (38) was recently sequenced at the E. coli genome center (http://www.genome.wisc .edu) (35). Amidst the 1.4 Mb of DNA that was not homologous with E. coli K-12 sequence, two urease gene clusters were identified in EDL933 at nonadjacent loci (35). Both gene clusters are predicted to encode the structural proteins and accessory polypeptides necessary for the assembly of urease. This complex nickel metalloenzyme catalyzes the hydrolysis of urea into ammonia and carbon dioxide (32). It has been demonstrated previously that ammonium ions accumulate in the proximal surroundings of urease-positive bacteria, mediating an increase in pH (32). It is postulated elsewhere that some acid-tolerant organisms like Helicobacter pylori, Streptococcus salivarius, and Yersinia enterocolitica exploit this enzyme to buffer themselves in acidic niches within a host (6, 13, 32, 41). Similar to these pathogens, E. coli O157:H7 is known to have a high tolerance to acid (1, 29). This phenotype is partially inhibited by chloramphenicol, suggesting that de novo protein synthesis is probably required (2, 11). Interestingly, urease production in H. pylori and S. salivarius has been demonstrated elsewhere to increase under acidic culture conditions (6, 42). We speculate that urease production could be part of a met-

abolic adaptation employed by EHEC to transiently neutralize acidity experienced during an oral route of infection (2, 11) or with exposure to organic acids present in the human intestinal tract (12). In this report, we demonstrate that EDL933 carries functional urease genes, which can be expressed in an E. coli K-12 background. EHEC ureases are preceded by Fur-like boxes, which appear to regulate ure expression as detected by enzymatic assays. This regulation is illustrated in both a recombinant system and an isogenic fur mutant generated in this study. The significance of Fur regulating acid tolerance is well documented elsewhere for Salmonella enterica serovar Typhimurium (17, 21, 36). Variations in ure regulation noted among O157:H7 strains used in this study suggest that additional mechanisms of regulation probably exist. Sequence analysis of E. coli EDL933 ure gene cluster. Despite the 390 kb of DNA separating the two ure gene clusters identified in E. coli EDL933, these clusters appear to be identical in sequence (35). Both loci of 4,900 bp encode three structural gene homologues, ureA, ureB, and ureC, as well as four accessory gene homologues, ureD, ureE, ureF, and ureG, similar to a majority of microbial ureases (32). Based on translated nucleotide sequences, the ure products of EDL933 are most similar to the homologues of Klebsiella aerogenes (M36068) (33). The structural proteins UreA, UreB, and UreC are predicted to share 80 to 96% similarity with the K. aerogenes polypeptides, whereas the translated accessory gene products of ureD, ureE, ureF, and ureG are predicted to share 80, 77, 76, and 93% similarity with K. aerogenes homologues, respectively. These urease accessory proteins are speculated to assist in procuring and coordinating the nickel ions in the active site of the apourease (10, 23, 27, 43). Urease expression from cloned EHEC ure gene clusters.

* Corresponding author. Mailing address: Department of Microbiology and Immunology, University of Maryland School of Medicine, 655 W. Baltimore St., Baltimore, MD 21201. Phone: (410) 706-0466. Fax: (410) 706-6751. E-mail: [email protected]. 1027

1028

NOTES

FIG. 1. Urease activities of cloned EHEC ure clusters expressed in E. coli DH5␣ and native EDL933 background. Strains were cultured overnight in M9 medium modified with 1.0 ␮M NiCl2. French-pressed lysates were standardized for protein concentration and assayed for urease activity, reported as nanomoles of NH3 per minute per milligram of protein (30, 48). Each column represents triplicate measurements averaged over three independent experiments. Error bars indicate standard deviations. Cloned EDL933 ure (pEDLure) was reisolated from EDL933 and transformed into E. coli DH5␣ (indicated by arrows). As a control, assay results with recombinant P. mirabilis ure (pMID1010) (24) transformants are included. E. coli DH5␣ (pBS SK⫹) routinely hydrolyzes ⬍10 nmol of urea/min/mg of protein. The inset is an anti-UreC immunoblot of lysates of E. coli DH5␣ carrying cloned EHEC ure clusters (3, 22). Relative molecular masses are indicated on the left. The UreC homologue is predicted to migrate at ⬃60 kDa (right arrow).

Initially, EDL933 cultures were examined for urease activity on Christensen’s agar, which contains a pH-sensitive, colorimetric indicator (8). In this assay, EDL933 cultures failed to generate pink colonies which are indicative of an alkaline pH (8; data not shown). Urease-specific enzymatic assays (30, 48) and immunoblotting with polyclonal rabbit sera raised against Proteus mirabilis homologues of UreC and UreD confirmed the absence of urease gene expression in L broth cultures (3, 22; data not shown). However, a 5,000-bp DNA fragment encoding ureDABCEFG (pEDL ure) derived from EDL933 conferred urease activity on E. coli DH5␣ (39) upon transformation (550 nmol of NH3 generated/min/mg of protein) (Fig. 1). This gene cluster was PCR amplified from genomic DNA preparations of EDL933 using the Expand Long Template PCR system (Boehringer Mannheim) with primer pairs 5⬘-TCGGAGCTCTCTGCCT GATTCACTGGATAA-3⬘ (upstream primer) and 5⬘-AAC GCCAACTTGGATCCTTCCTTCTGATAA-3⬘ (downstream primer) (3, 39). The gel-purified PCR product, which extends from 44 bp upstream of the first possible ureD translational start site to 53 bp downstream of the ureG stop codon (Fig. 2), was cloned into the EcoRV site of pBS SK⫹ for expression (3, 39). Because the two ure gene clusters of EDL933 are identical, it is not clear which of the two loci was cloned. For comparison, the ure clusters of O157:H7 isolates IN1

INFECT. IMMUN.

and MO28 were cloned and analyzed in a similar manner. These strains produce different DNA restriction patterns in pulsed-field gels probed with AvrII fragments containing the ure cluster and thus represent distinct clones (data not shown). However, DNA sequences (600 bp) upstream of the ure clusters were sequenced and found to be highly conserved (data not shown). Overall, comparable levels of urease activities were detected in E. coli DH5␣ cells transformed with the cloned ure clusters of EDL933, IN1, and MO28 (Fig. 1). Furthermore, cell lysates immunoblotted with antiserum to P. mirabilis UreC, the largest structural subunit, cross-reacted with appropriately sized polypeptides (Fig. 1, inset) (3, 22). These observations indicate that EHEC strains EDL933, IN1, and MO28 carry functional urease genes, despite the fact that ure gene expression in the native EDL933 cannot be detected on Christensen’s urea agar or stationary-phase L broth cultures. In contrast, O157:H7 strains IN1 and MO28 do produce an enzymatically active urease during stationary-phase growth in L broth cultures (200 ⫾ 10 and 85 ⫾ 20 nmol of NH3/min/mg of protein, respectively). Note that this activity is less than that expressed by the cloned determinants in E. coli DH5␣ (Fig. 1). Urease expression of the cloned ure genes in native EHEC strain EDL933. To address the disparity in urease expression between EDL933 and the cloned determinant in a DH5␣ background, pEDLure was transformed into EDL933 and subsequently analyzed for catalytic activity. As with the native strain, EDL933 (pEDLure) transformants do not produce urease. Thus, the cloned ure genes were repressed in their native background (Fig. 1). However, derepression could be demonstrated upon isolation of pEDLure and retransformation into E. coli DH5␣ (700 ⫾ 300 nmol of NH3/min/mg of protein) (Fig. 1). Based on this observation, we postulated that a transacting factor regulates EDL933 ure expression. Furthermore, the cloned EDL933 ure determinants were not expressed in EDL933⌬ure, in which DNA sequences extending 760 bp upstream and 840 bp downstream of both ure loci have been deleted by a method described in the work of Pósfai et al. (37; data not shown). Thus, we postulated that the sequence encoding this trans-acting factor is not immediately adjacent to either of the ure loci. In control experiments, P. mirabilis ure genes (pMID1010) (24) were shown to confer the same level of enzymatic activity upon EDL933 and DH5␣ (Fig. 1). Thus, the regulatory effect appears specific for the EDL933 ure genes. Fur-dependent regulation of O157:H7 ure gene clusters. In DNA sequence analyses of EDL933, three Fur-like (ferric uptake regulator) boxes were identified between the first and second possible translational start sites of ureD (Fig. 2) (16). The first of these putative recognition sequences lies immediately upstream of a ␴70-like promoter, situated halfway between the two possible start sites (Fig. 2). The second and third Fur boxes overlap and lie downstream of the putative promoter, respectively (Fig. 2). For comparison with other O157:H7 isolates, a 600-bp region immediately adjacent to the ure subclones of strains IN1 and MO28 was PCR amplified with primer pair 5⬘-TCATCG TACCCCCATTAGTTAGAGC-3⬘ (upstream primer) and 5⬘-GCTGTGATTATGCCTTCTCGCTCAG-3⬘ (downstream primer) and sequenced at the University of Maryland, Baltimore Biopolymer Core Facility (3, 40). This region was highly

VOL. 70, 2002

NOTES

1029

FIG. 2. A 300-bp DNA sequence upstream of the EHEC ure cluster in strains EDL933, IN1, and MO28. Recombinant EHEC ure constructs used in this study include DNA sequence extending to the region labeled as ure upstream primer. This stretch of DNA encodes three putative translational start sites for the ureD homologue (marked as “1st SC,” “2nd SC,” and “3rd SC”). A ␴70-like promoter (boldface) was identified between the first and second putative ureD start codons (SC). Based on the consensus sequence reported in the work of Escolar et al. (16), two Fur-like boxes were found upstream and overlapping the ␴70-like promoter, together with another recognition sequence identified further downstream. These Fur-like boxes are highlighted within the sequence (boxes) with an indication of their relative direction (arrows).

conserved between EDL933 and these two urease-positive strains (data not shown). To examine whether these Fur-like boxes had any effect on O157:H7 ure transcription, cloned EHEC ure determinants were transformed into E. coli MC3100 and an isogenic fur mutant (44). Cloned EDL933 ure expression, measured by enzymatic activity (30, 48), increased a modest fivefold in the fur mutant background (P ⫽ 0.042) (Fig. 3B). In contrast, the urease activities associated with the cloned IN1 and MO28 gene clusters decreased dramatically in a ⌬fur background. E. coli MC3100(pINure) produced 99.5 nmol of NH3/min/mg of protein, a 50-fold excess over that of the isogenic ⌬fur strain (2.1 nmol of NH3/min/mg of protein) (P ⬍ 0.001) (Fig. 3B). Likewise, E. coli MC3100(pMOure) urease activity exceeds that of the ⌬fur mutant by 25-fold (Fig. 3B) (P ⬍ 0.001). To verify the Fur phenotypes of these strains, a Fur-repressible gene fusion, chuA-phoA (31), was introduced in trans into both E. coli MC3100 and the isogenic ⌬fur strain (45). As an indicator of chuA transcription, alkaline phosphatase activity was measured in mid-exponential-phase L broth cultures according to the method of Brickman and Beckwith (5). As described in previous studies, chuA expression was fivefold greater in the ⌬fur strain than in the wild type (P ⫽ 0.014) (45) (Fig. 3A). These data suggest that Fur can regulate recombinant ure genes of multiple O157:H7 isolates (Fig. 3B). It is unclear why the phenotypes of the IN1 and MO28 cloned ure determinants were altered dramatically in a ⌬fur background while EDL933 ure showed only a small change in activity despite the high degree of homology upstream of the ure clusters, including the Fur-like boxes. To confirm that Fur influences EHEC ure expression, nonpolar insertional mutations were constructed within the fur loci of IN1 and EDL933. This was accomplished by PCR amplifying a 2,400-bp region encoding fur and flanking sequences from

genomic DNA with the primer pair 5⬘-GACTGCCTGTTCTG CTATGATTG-3⬘ (upstream primer) and 5⬘-TACTTCCTGC AACGTATGACTCC-3⬘ (downstream primer) (3). PCR products were cloned into the EcoRV site of pBS SK⫹ and subsequently interrupted at an internal XmnI site with a 1,200-bp nonpolar aphA cassette originating from pUC4K (47). After confirmation of the construction of fur::aphA by DNA sequence analysis (40), this insertional mutation was transferred into the SmaI site of the ␭pir-dependent suicide vector pCVD442 (15). Mutations in fur were generated in nalidixic acid-resistant isolates of IN1 and EDL933 by conjugation with E. coli S17␭pir transformed with the suicide construct (20). Transconjugants having undergone replacement recombination were confirmed by PCR amplification of genomic DNA with oligomers (described above) flanking fur, generating a single 3,500-bp product (3; data not shown). Control PCR amplifications were performed with genomic DNA from IN1 and EDL933 or plasmid DNA (pfur::aphA442) isolated from E. coli S17␭pir, which generated 2,400- and 3,500 bp DNA products, respectively (data not shown). Similar to results in an E. coli MC3100⌬fur background (Fig. the IN1 fur::aphA mutant produced significantly less urease activity than did the isogenic wild-type strain. The urease specific activities for IN1 and IN1 fur::aphA were 199.2 ⫾ 7.8 and 111.9 ⫾ 8.7 nmol of NH3/min/mg of protein, respectively (P ⬍ 0.001 by the Student t test). These strains were also transformed with pchuA-phoA. The alkaline phosphatase activities for IN1 and IN1fur::aphA were 14.5 ⫾ 1.6 and 79.1 ⫾ 15.6 arbitrary units (5), respectively. (P was 0.001). EDLfur::aphA did not differ in ure expression from EDL933 because neither strain produced detectable urease activity in stationary-phase cultures (data not shown). Altogether, these results suggest that Fur may enhance the basal level of ure expression in O157:H7 isolates IN1 and

1030

NOTES

INFECT. IMMUN.

FIG. 3. Urease activity of cloned ure gene cluster of EHEC strains EDL933, MO28, and IN1 in the E. coli MC3100⌬fur strain. Cloned EHEC ure clusters were transformed into E. coli MC3100 and an isogenic ⌬fur strain. French-pressed lyates were prepared from overnight cultures grown in M9 medium modified with 1.0 ␮M NiCl2. Lysates were standardized for protein concentration and assayed for urease activity, reported as nanomoles of NH3 per minute per milligram of protein (30, 48) (B). Each bar represents the average of triplicate measurements of three separate experiments. Student t test results are reported for paired strains. As a positive control, a recombinant EDL933 chuA-phoA fusion (45) was introduced in trans into MC3100 and MC3100⌬fur. Subsequent assays of alkaline phosphatase activity are reported as arbitrary units based on the method of Brickman and Beckwith (5) (A).

MO28 during stationary-phase growth. Preliminary characterizations of ureD transcriptional fusions suggest that both IN1 fur::aphA and EDL fur::aphA produce less transcript than do the isogenic wild-type strains (data not shown). Thus, Furdependent regulation is probably occurring in EDL933, similar to IN1, but secondary to another mechanism of regulation. Screening for conditions that induce O157:H7 isolate EDL933 ure expression. To ascertain whether the native EDL933 could be induced to express urease like its cloned ure determinants, EDL933 was analyzed for regulation signals known to affect other bacterial ureases. Bacterial cultures were grown overnight in 5.0 ml of M9 minimal medium containing 1 ␮M NiCl2 at 37°C with aeration. Several variations in culture conditions were assessed for effects on urease expression, including (i) M9 minimal medium in which NH4Cl was separately replaced with 0.2% (wt/vol) proline, 0.2% (wt/vol) histidine, and 0.2% (wt/vol) glutamine (4, 9); (ii) M9 minimal medium in which NH4Cl was replaced with 0, 100, 250, and 500 mM urea (34); and (iii) M9 minimal medium in which sodium phosphate was omitted and subsequently buffered with 2-(N-morpholino)ethanesulfonic acid (MES) at pH 5.0, 6.0, and 7.5 (6, 41). Under all conditions tested, no urease activity was detected in O157:H7 strain EDL933. Likewise, polyclonal sera against UreC and UreD failed to react with proteins in whole-lysate preparations (data not shown). In contrast, IN1 had 1.9-foldgreater enzymatic activity in L broth at pH 5.0 than at pH 7.0 (148 ⫾ 10.1 versus 77.4 ⫾ 3.6 nmol of NH3/min/mg of protein) (P ⫽ 0.037), suggesting that enhanced urease expression may be correlated with exposure to acidic environments.

Introduction of a gene encoding a high-affinity Ni2⫹ transporter (nixA) of H. pylori also did not affect urease detection in EDL933 under any condition tested; thus, Ni2⫹ availability was not limiting for the metalloenzyme (18). As the frequency and severity of E. coli O157:H7 outbreaks continue to increase, there is an accompanying need to understand the pathogenicity of this organism. Undoubtedly during the course of infection, EHEC encounters both gastric acidity and permeable, weak acids in the human intestinal tract (7). It is not surprising that EHEC has evolved a resilient mechanism to regulate acid resistance, offering protection from transient and prolonged exposure to various acids. Overall, our findings suggest that O157:H7 strains encode functional urease genes, which may be regulated both by an unindentified trans-acting factor and, more conclusively, by Fur. Like the glutamate- and arginine-dependent systems described elsewhere for E. coli (26, 28, 29), urease expression could modify internal and/or surrounding anion concentrations, enabling EHEC to survive acidic conditions and perhaps contributing to its low infectious dose. To our knowledge, this report is the first description of a Fur-regulated urease. The O157:H7 isolates IN1 (FDA2-51) and MO28 (FDA2-25) are SOR⫺ MUG⫹ URE⫹ strains obtained from the Minnesota Department of Public Health. E. coli strains MC3100 and MC3100⌬fur (24) were generously supplied by P. F. Sparling at the University of North Carolina at Chapel Hill and may otherwise be known as W3100⌬lac and W3100 fur⌬lac (14). The chuA-phoA fusion was the gift of Alfredo Torres at the CVD, University of Maryland, Baltimore, and was constructed as pSHU251 as described in the work of Mills and Payne (31). This work was supported in part by Public Health Service grant AI23328 from the National Institutes of Health.

VOL. 70, 2002

NOTES

ADDENDUM IN PROOF In a recent survey of 55 diarrheagenic strains of Escherichia coli, 22 strains (exclusively enterohemorrhagic E. coli [EHEC]) hybridized with an internal ureC probe. However, only one of the 22 EHEC strains was phenotypically urease positive (M. Nanko, T. Hayashi, and T. Honda, J. Clin. Microbiol. 39:4541– 4543, 2001). REFERENCES 1. 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. 2. Arnold, K. W., and C. W. Kaspar. 1995. Starvation- and stationary-phaseinduced acid tolerance in Escherichia coli O157:H7. Appl. Environ. Microbiol. 61:2037–2039. 3. Ausubel, F. M., R. Brent, R. E. Kingston, D. D. Moore, J. A. Smith, J. G. Seidman, and K. Struhl. 1987. Current protocols in molecular biology. Greene Publishing Associates and Wiley-Interscience, New York, N.Y. 4. Bender, R. A., K. A. Janssen, A. D. Resnick, M. Blumenberg, F. Foor, and B. Magasanik. 1977. Biochemical parameters of glutamine synthetase from Klebsiella aerogenes. J. Bacteriol. 129:1001–1009. 5. Brickman, E., and J. Beckwith. 1975. Analysis of the regulation of Escherichia coli alkaline phosphatase synthesis using deletions and ␾80 transducing phages. J. Mol. Biol. 96:307–316. 6. Chen, Y. M., C. A. Weaver, D. R. Mendelsohn, and R. A. Burne. 1998. Transcriptional regulation of the Streptococcus salivarius 57.1 urease operon. J. Bacteriol. 180:5769–5775. 7. Cherrington, C. A., M. Hinton, G. C. Mead, and I. Chopra. 1991. Organic acids: chemistry, antibacterial activity and practical applications. Adv. Microb. Physiol. 32:87–108. 8. Christensen, W. B. 1946. Urea decomposition as a means of differentiating Proteus and paracolon cultures from each other and from Salmonella and Shigella types. J. Bacteriol. 52:461–466. 9. Collins, C. M., D. M. Gutman, and H. Laman. 1993. Identification of a nitrogen-regulated promoter controlling expression of Klebsiella pneumoniae urease genes. J. Bacteriol. 8:187–198. 10. Colpas, G. J., and R. P. Hausinger. 2000. In vivo and in vitro kinetics of metal transfer by the Klebsiella aerogenes urease nickel metallochaperone, UreE. J. Biol. Chem. 275:10731–10737. 11. Conner, D. E., and J. S. Kotrola. 1995. Growth and survival of Escherichia coli O157:H7 under acidic conditions. Appl. Environ. Microbiol. 61:382–385. 12. Cummings, J. H., E. W. Pomare, W. J. Branch, C. P. E. Naylor, and G. T. MacFarlane. 1987. Short chain fatty acids in human large intestine, portal, hepatic, and venous blood. Gut 28:1822–1824. 13. De Koning-Ward, T. F., and R. M. Robins-Browne. 1995. Contribution of urease to acid tolerance in Yersinia enterocolitica. Infect. Immun. 63:3790–3795. 14. DeLorenzo, V., M. Herrero, F. Giovannini, and J. B. Neilands. 1988. Fur (ferric uptake regulation) protein and CAP (catabolite-activator protein) modulate transcription of fur gene in Escherichia coli. Eur. J. Biochem. 173:537–546. 15. Donnenberg, M. D., and J. B. Kaper. 1991. Construction of an eae deletion mutant of enteropathogenic Escherichia coli by using a positive-selection suicide vector. Infect. Immun. 59:4310–4317. 16. Escolar, L., J. Pérez-Martín, and V. de Lorenzo. 1999. Opening the iron box: transcriptional metalloregulation by the Fur protein. J. Bacteriol. 181:6223– 6229. 17. Foster, J. W. 1991. Salmonella acid shock proteins are required for the adaptive acid tolerance response. J. Bacteriol. 173:6896–6902. 18. Fulkerson, J. F., Jr., R. M. Garner, and H. L. T. Mobley. 1998. Conserved residues and motifs in the NixA protein of Helicobacter pylori are critical for the high affinity transport of nickel ions. J. Biol. Chem. 273:235–241. 19. Griffin, P. M., S. M. Ostroff, R. V. Tauxe, K. D. Greene, J. G. Wells, J. H. Lewis, and P. A. Blake. 1988. Illnessess associated with Escherichia coli O157:H7 infections. Ann. Intern. Med. 109:705–712. 20. Guyer, D. M., I. R. Henderson, J. P. Nataro, and H. L. T. Mobley. 2000. Identification of Sat, an autotransporter toxin produced by uropathogenic Escherichia coli. Mol. Microbiol. 38:53–66. 21. Hall, H. K., and J. W. Foster. 1996. The role of Fur in the acid tolerance response of Salmonella typhimurium is physiologically and genetically separable from its role in iron acquisition. J. Bacteriol. 178:5683–5691. 22. Heimer, S. R., and H. L. T. Mobley. 2001. Interaction of Proteus mirabilis urease apoenzyme and accessory proteins identified with yeast two-hybrid technology. J. Bacteriol. 183:1423–1433.

Editor: V. J. DiRita

1031

23. Island, M. D., and H. L. T. Mobley. 1995. Proteus mirabilis urease: linkerinsertion mutagenesis of the positive activator and accessory genes. J. Bacteriol. 177:5653–5660. 24. Jones, B. D., and H. L. T. Mobley. 1988. Proteus mirabilis urease: genetic organization, regulation, and expression of structural gene. J. Bacteriol. 170:3342–3349. 25. Kaper, J. 1998. Enterohemorrhagic Escherichia coli. Curr. Opin. Microbiol. 1:103–108. 26. 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. 27. Lee, M. H., S. B. Mulrooney, M. J. Renner, Y. Markowicz, and R. P. Hausinger. 1992. Klebsiella aerogenes urease gene cluster: sequence of ureD and demonstration that four accessory genes (ureD, ureE, ureF, and ureG) are involved in nickel metallocenter biosynthesis. J. Bacteriol. 174:4324–4330. 28. Lin, J., I. S. Lee, J. Frey, J. L. Slonczewski, and J. W. Foster. 1995. Comparative analysis of extreme acid survival in Salmonella typhimurium, Shigella flexneri, and Escherichia coli. J. Bacteriol. 177:4097–4104. 29. Lin, J., M. P. Smith, K. C. Chapin, H. S. Baik, G. N. Bennett, and J. W. Foster. 1996. Mechanisms of acid resistance in enterohemorrhagic Escherichia coli. Appl. Environ. Microbiol. 62:3094–3100. 30. McGee, D. J., C. A. May, R. M. Garner, J. M. Himpsl, and H. L. T. Mobley. 1999. Isolation of Helicobacteri pylori genes that modulate urease activity. J. Bacteriol. 181:2477–2484. 31. Mills, M., and S. M. Payne. 1997. Identification of shuA, the gene encoding the heme receptor of Shigella dysenteriae, and analysis of invasion and intracellular multiplication of a shuA mutant. Infect. Immun. 65:5358–5363. 32. Mobley, H. L. T., M. D. Island, and R. P. Hausinger. 1995. Molecular biology of microbial ureases. Microbiol. Rev. 59:451–480. 33. Mulrooney, S. B., and R. P. Hausinger. 1990. Sequence of the Klebsiella aerogenes urease genes and evidence for accessory proteins facilitating nickel incorporation. J. Bacteriol. 172:5837–5843. 34. Nicholson, E. B., E. A. Concaugh, and H. L. T. Mobley. 1991. Proteus mirabilis urease: use of a UreA-LacZ fusion demonstrates that induction is highly specific for urea. Infect. Immun. 59:3360–3365. 35. Perna, N. T., G. Pluckett III, V. Burland, B. Mau, J. D. Glasner, D. J. Rose, G. F. Mayhew, P. S. Evans, J. Gregor, H. A. Kirkpatrick, G. Pósfai, J. Hackett, S. Klink, A. Boutin, Y. Shao, L. Miller, E. J. Grotbeck, N. W. Davis, A. Lim, E. Dimalanta, K. Potamousis, J. Apodaca, T. S. Anantharaman, J. Lin, G. Yen, D. C. Schwartz, R. A. Welch, and F. R. Blattner. 2001. Genomic sequence of enterohemorrhagic Escherichia coli O157:H7. Nature 409:529–533. 36. Portillo, F. G.-D., J. W. Foster, and B. B. Finlay. 1993. The role of acid tolerance response genes in Salmonella typhimurium virulence. Infect. Immun. 61:4489–4492. 37. Pósfai, G., M. D. Koob, H. A. Kirkpatrick, and F. R. Blattner. 1997. Versatile insertion plasmids for targeted genome manipulations in bacteria: isolation, deletion, and rescue of the pathogenicity island LEE of the Escherichia coli O157:H7 genome. J. Bacteriol. 179:4426–4428. 38. Riley, L. W., R. S. Remis, S. D. Helgerson, H. B. McGee, J. G. Wells, B. R. Davis, R. J. Herbert, E. S. Olcott, L. M. Johnson, N. T. Hargrett, P. A. Blake, and M. L. Cohen. 1983. Hemorrhagic colitis associated with a rare Escherichia coli serotype. N. Engl. J. Med. 308:681–685. 39. Sambrook, J., E. F. Fritsch, and T. Maniatis. 1989. Molecular cloning: a laboratory manual, 2nd ed. Cold Spring Harbor Laboratory Press, Cold Spring Harbor, N.Y. 40. Sanger, F., S. Nicklen, and A. R. Coulson. 1977. DNA sequencing with chain-terminating inhibitors. Proc. Natl. Acad. Sci. USA 74:5463–5467. 41. Sissons, C. H., H. E. R. Perinpanayagam, E. M. Hancock, and T. W. Cutress. 1990. pH regulation of urease levels in Streptococcus salivarius. J. Dent. Res. 69:1131–1137. 42. Slonczewski, J. L., D. J. McGee, J. Phillips, C. Kirkpatrick, and H. L. T. Mobley. 2000. pH-dependent protein profiles of Helicobacter pylori analyzed by two-dimensional gels. Helicobacter 5:240–247. 43. Soriano, A., and R. P. Hausinger. 1999. GTP-dependent activation of urease apoprotein in complex with the UreD, UreF, and UreG accessory proteins. Proc. Natl. Acad. Sci. USA 96:11140–11144. 44. Thomas, C. E., and P. F. Sparling. 1994. Identification and cloning of a fur homologue from Neisseria meningitidis. Mol. Microbiol. 11:725–737. 45. Torres, A. G., and S. M. Payne. 1997. Haem iron-transport system in enterohaemorrhagic Escherichia coli O157:H7. Mol. Microbiol. 23:825–833. 46. Verweyen, H. M., H. Karch, M. Brandis, and L. B. Zimmerhackl. 2000. Enterohemorrhagic Escherichia coli infections: following transmission routes. Pediatr. Nephrol. 14:73–83. 47. Vieira, J., and J. Messing. 1982. The pUC plasmids, an M13mp7-derived system for insertion mutagenesis and sequencing with synthetic universal primers. Gene 19:259–268. 48. Weatherburn, M. W. 1967. Phenol-hypochlorite reaction for determination of ammonia. Anal. Chem. 39:971–974.