Bookshelf » Helicobacter pylori » Physiology and Molecular Biology » Regulation of Urease for Acid Habitation
 
hp
Helicobacter pylori
Physiology and Genetics
MobleyHarry L. T.
MendzGeorge L.
HazellStuart L.
2001
Microbiology
part

 Chapter 25:  Regulation of Urease for Acid Habitation

George Sachs,1 David R. Scott,1 David L. Weeks,1 Marina Rektorscheck,2 and Klaus Melchers2
1University of California, Los Angeles, Los Angeles, CA, 90073
2Byk Gulden, Konstanz, Germany
A2543

*

*

*

*

The discovery of Helicobacter pylori as a gastroinfective organism responsible for peptic ulcer disease engendered a revolution in therapeutic approaches to treatment of this disease and may eventually modify treatment of gastroesophageal reflux disease (3, 22). Infection of the body of the stomach by the organism enhances the risk of gastric cancer (12), making its eradication from this location a priority. Occupancy of the body of the stomach is enhanced by chronic treatment with acid-inhibitory drugs such as proton pump inhibitors (PPIs). The infection is extracorporeal and therefore requires high-efficacy bactericidal agents for eradication since there is no synergy with macrophages as occurs with systemic infections. The stomach is also a difficult environment for targeting drugs, which lowers the efficiency of systemic antibiotics. These considerations suggest that understanding the unique ability of H. pylori to colonize the normal human stomach would help the discovery of effective drugs against this organism.

Almost all prokaryotes depend on the presence of a proton motive force (PMF) across their cytoplasmic membrane for generation of ATP and for solute import and export (10). The PMF is the electrochemical gradient of protons across the cytoplasmic membrane and consists of a proton gradient and an electrical potential difference. The majority of species are neutralophiles that live within a predominantly neutral pH range. Some, however, inhabit acidic or alkaline environments, attesting to the adaptability of this early life form. The values of the two components of the PMF, the relative pH and potential gradient across the cytoplasmic membrane, reflect the pH of the environment. At neutral pH, the larger component is the electrical potential; at acidic pH, the larger component is the pH gradient. Hence, there is a reciprocal relationship between environmental pH and cytoplasmic membrane potential in these organisms.

It is well established that H. pylori grows best at neutral pH and fails to survive at a pH below 4.0 or above 8.2 in the absence of chemicals such as urea, thus, behaving as a neutralophile. Hence, it can be expected that its PMF will be present over the pH range at which neutralophiles can survive, but not outside this range in the absence of urea.

An external file that holds a picture, illustration, etc., usually as some form of binary object. The name of referred object is ch25f1.jpg.

Figure 1.

Relationship between membrane potential and external pH in H. pylori. The potential across the inner membrane of H. pylori as a function of fixed medium pH using a fluorescent membrane potential dye, as detailed in reference 17. There is a reciprocal relationship between medium acidity and membrane potential as predicted from the chemiosmotic theory.

Experimental determination of pH and electrical gradients in microorganisms relies on the distribution of lipid-permeable ions and weak bases or acids between the intracellular and extracellular environments (19). In the case of H. pylori, it was discovered that fluorescent probes of cytoplasmic pH and membrane potential were applicable to the measurement of its PMF. It is possible to measure the membrane potential with the carbocyanine dye, DiSC3 (5), and the cytoplasmic pH with the dye BCECF-AM. The expected reciprocal relationship between pH and electrical gradient was found between pH 4.0 and 8.2; irreversible loss of PMF-generating ability was found below a pH of 4.0 and above a pH of 8.2 (Fig. 1) (9). These findings indicate that the organism is a neutralophile. Given the relative acidity of the environment of the stomach, H. pylori must have mechanisms for achieving acid resistance, required both for infection and for colonization.

Acid Resistance

Many microorganisms can survive gastric acidity during their transit to other sites of infection. A straightforward means of surviving the high acidity of the stomach would be to restrict the proton permeability of the inner membrane, so that passage of H+ across the F1F0 ATPase or other H+-coupled reactions would occur only at very acidic pH. Alternatively, Na+ gradients could be used to drive ATP synthesis and solute import/export as is found in Propionum modestum (5). A means employed by acidophiles is the presence of a proton conductance reversing the potential across the inner membrane in the face of a large inward pH gradient (8). Either relative proton impermeability or high proton conductance would make it difficult for the organism to survive at neutral pH, unless a series of neutral pH responses were present. These would require large changes in the amino acid composition of membrane proteins, for example, by inserting sufficient charged amino acids to elevate or reduce the isoelectric point of the membrane domain of transport proteins; this is an unlikely scenario. However, some modification of the isoelectric point of proton-conducting transporters could modify acid resistance. An example is the Fc subunit of H. pylori that has a significantly higher isoelectric point than the corresponding protein of Escherichia coli or Bacillus subtilis.

Yersinia enterocolitica has a cytoplasmic urease with a pH optimum of 5.0, and the enzyme is inactive at neutral external pH. At acidic pH, with entry of protons into the organism, the cytoplasm acidifies to a pH of 5.0, the urease becomes active, and its activity buffers the cytoplasm by the generation of NH3 (24). Since protein synthesis is absent in Y. enterocolitica at a cytoplasmic pH of 5.0, this mechanism allows gastric survival but not gastric habitation.

H. pylori synthesizes large quantities of a dimeric urease with optimum activity at neutral pH, which accounts for 10 to 15% of its total protein. This protein synthesis is constitutive (11). In vitro and in vivo, urease, detected by either immunostaining or activity, is found both inside and outside the organism (14). The presence of external urease is likely due to lysis of the bacterium with adhesion of the urease to the outside surface (13). In principle, the activity of either urease or both could account for gastric survival and colonization. Their relative contributions are determined by their respective amounts and the pH of the medium. At acidic pH of 4.5 or less, the outer urease is inactive; therefore, it does not contribute to acid survival.

Urease activity is essential for infection of animal models by H. pylori (1, 6, 20). The experiments carried out to date with urease-negative organisms either show no infection or weak infection when acid secretion is inhibited with omeprazole, a PPI. No experiments have been described wherein infection by this class of mutant was induced in the presence of a PPI, and then disappeared when the PPI treatment was stopped after infection. This type of experiment would show that urease activity was essential not only for infection but also for colonization.

H. pylori urease has a neutral pH optimum between 7.5 and 8.5 but essentially no activity below a pH of 4.5, and activity is lost irreversibly at this pH. Hence, the urease component exposed to the medium on the surface of the organism cannot be responsible for acid resistance at a pH of 4.5 or below, but this urease component could contribute to elevating the pH of the gastric environment of the organism if the pH remained above 4.5. The result of pH elevation due to outer urease activity would be that pH would rise to greater than 8.2 in the absence of acid or buffer, and then this neutralophile would not survive. Thus, urease activity on the surface could be lethal at neutral pH.

Figure 2
Figure 2 pH profiles of H. pylori urease activity in soluble (more...)
An external file that holds a picture, illustration, etc., usually as some form of binary object. The name of referred object is ch25f2.jpg.

Figure 2.

pH profiles of H. pylori urease activity in soluble urease (A) and intact bacteria (B) and profile of urea uptake in UreI-expressing oocytes (C). The activity of soluble or surface urease as a function of medium pH is shown. In contrast to the pH profile of free urease as shown in panel A, in intact bacteria (B) there is little activity at neutral pH, and a large increase at pH below 6.5 is interpreted as acid activation of intrabacterial urease. In panel C, urea uptake in UreI-expressing oocytes as a function of medium pH shows strong similarity to activation of urease in intact bacteria shown in panel B.

Intrabacterial urease shows a marked increase of activity with medium acidification, whether urease activity is measured directly by CO2 release or by pH elevation using a microphysiometer in contrast to the decrease in activity of surface or soluble urease (Fig. 2A and B) (15, 17). The increase of activity is between 10- and 20-fold when external urease activity is present but can rise much more if surface urease is first washed off or selectively inactivated.

Activation begins at a medium pH of 6.5 and reaches its maximum value at pH 5.5. Urease activity remains constant down to a pH between 2.5 and 3.0 and is detectable even at a pH of 2.0. This behavior corresponds to that of an acid resistance mechanism, wherein the neutralizing capacity is maintained over a broad range of acidic pH and is lost at pH levels where ammonia generation would generate a toxic environment for the organism. Since what is vital for the organism is the PMF (pH/potential gradient) across the inner membrane, uncontrolled internal alkalinization would be unsafe, and the objective of creating a safe environment would be achieved by neutralization of the periplasm. This goal is also economical because the volume of the periplasm, a fraction of a femtoliter, is small compared to the volume of the "microenvironment'' surrounding the organism. A membrane potential of about 100 mV is measured in the presence of urea at medium pH from 3.0 to 6.0, and the calculated or measured periplasmic pH is 6.2 (15, 17). These values of the components of the PMF are compatible with protein synthesis and growth.

The mechanism underlying this pH activation of urease is of vital interest in considering possible selective targets for monotherapy. A clue to understanding this mechanism was gleaned from measurements of the Km,app for urea in solubilized enzyme and suspensions of intact bacteria at neutral and at acidic pH. At acidic pH the Km,app for intact bacteria was ~1 mM, similar to the Km,app for soluble enzyme at neutral pH. However, at neutral pH the Km,app for urease in intact bacteria is ~200 mM. The apparent absence of saturation at neutral pH in intact organisms suggested that at this pH, the rate-limiting step was no longer urease activity but access of urea to intrabacterial urease. Hence, the observed activation step of urease appears to involve urea transport (17).

Passive urea permeability across phospholipid bi-layers is 4 × 10−6 cm s−1. Owing to the small volume (~8 fl) and large surface-to-volume ratio of H. pylori, it is difficult to measure urea uptake by bacteria because equilibration is very rapid. To overcome this problem the Xenopus oocyte expression system was employed, since these oocytes have a volume of ~0.5 μl and a much smaller surface-to-volume ratio. Direct evidence for acid-activated urea transport was obtained by oocyte expression of one of the genes of the urease gene cluster, UreI.

There are seven genes in the urease gene cluster; ureA and ureB encode the structural subunits of the enzyme, and ureE, F, G, and H code for proteins involved in urease assembly, although only the gene products of the last three are absolutely essential for generation of activity. Protein-protein interactions between UreE and UreG and UreF and UreH were observed by using a yeast two-hybrid system, and interactions of UreB with UreE and UreH were demonstrated by the same methodology. It appears that two pairs of proteins are formed that bind to UreB, and then assembly and nickel insertion into the UreA/B complex occurs (21).

Figure 3
Figure 3 pH profile of H. pylori urease activity in UreI-negative (more...)
An external file that holds a picture, illustration, etc., usually as some form of binary object. The name of referred object is ch25f3.jpg.

Figure 3.

pH profile of H. pylori urease activity in UreI-negative mutants. The activity of urease in UreI-negative mutants as a function of pH (●) shows that the enzyme has lost its acid activation in the mutants. Full urease activity is restored by lysis or the addition of 0.01% C12E8, owing to the disruption of the bacterial inner membrane and this also shows the nonpolar nature of the mutant strain ([open circle]).

The open reading frame ureI is located between ureB and ureE and preceded by a promoter region. This gene encodes a product with a sequence predicted to be an inner membrane protein (7). Nonpolar deletion of ureI results in loss of acid activation of intrabacterial urease (16), and acid activation is restored when the mutant is complemented by a vector containing ureI. It is important to note that in the ureI-negative mutant, urease is fully activated when a low concentration of the nonionic detergent C12E8 is added to permeabilize the bacterial inner membrane (14). The loss of activation in the ureI-negative mutant measured by CO2 production and the restoration of urease activity with the addition of detergent are shown in Fig. 3. The data obtained with pH measurements are in accordance with the radioactivity assays. In general, nongastric helicobacters, although they have varying amounts of urease activity, do not show acid activation or express UreI (16). Thus, UreI expression appears to be a specialized adaptation within the urease gene cluster for gastric habitation.

Figure 4
Figure 4 Confocal image of H. pylori. A confocal image of (more...)
An external file that holds a picture, illustration, etc., usually as some form of binary object. The name of referred object is ch25f4.jpg.

Figure 4.

Confocal image of H. pylori. A confocal image of H. pylori incubated in the presence of BCECF-free acid at a pH of 5.5 immediately following the addition of urea. This dye does not penetrate the inner membrane and increases fluorescence with an increase of pH. The halo of increased fluorescence at the periphery of the organism defines the site of elevation of pH as being initially in the periplasm.

The elevation of pH with the addition of urea was visualized with confocal microscopy of bacteria cocultured with AGS cells and BCECF-free acid as a fluorescent pH probe of either the medium or the periplasm. This dye does not penetrate cell or inner bacterial membranes. Addition of urea at a pH of 5.5 to weakly buffered medium results in immediate elevation of pH in the bacterial periplasm, as shown in Fig. 4, followed by elevation of the medium pH in wild-type cells but not in ureI deletion mutants. In contrast, addition of detergent results in initial alkalinization of the cytoplasm in both wild-type and mutant organisms (2).

The Mechanism of UreI

Interesting properties are revealed when the cRNA for UreI is injected into oocytes. First and foremost, the oocytes acquire an acid-activated urea uptake, which is similar to that found for the urease activity in intact bacteria shown in Fig. 2B and C. Urea uptake by oocytes is essentially nonsaturable and highly selective; for example, thiourea is not taken up. Acid-stimulated urea uptake is temperature independent (23). The protein contains six predicted membrane-inserted segments with N and C termini in the periplasm. There are two periplasmic loops containing five or six histidine residues. Removal of any of the last three histidyls, two found in the second periplasmic loop and one in the C terminus, results in inactivation of the protein. These data suggest that UreI is a proton-gated urea channel and that protonation of periplasmic histidines, at least in the second periplasmic loop, is involved in the gating reaction. Interestingly, a homologous gene in Streptococcus salivarius also encodes a protein that transports urea in Xenopus oocytes but does so at neutral and acidic pH. This gene lacks most of the region encoding the periplasmic loops. Cross-linking and sodium dodecyl sulfate gel profiles suggest that the transport protein is an oligomer of at least two monomers, but it is not known whether function requires oligomerization (Weeks, unpublished observations).

An external file that holds a picture, illustration, etc., usually as some form of binary object. The name of referred object is ch25f5.jpg.

Figure 5.

A model illustrating the function of UreI. UreI is shown as a proton-gated urea channel providing urea access to intrabacterial urease under acidic conditions due to protonation of His-123. Two additional histidines and several carboxylic acids in the periplasmic loops may also participate in the gating response.

These data suggest that urease-dependent adaptation to acidity in vitro depends on the transport function of UreI. UreI deletion mutants are unable to colonize mice, supporting the role of UreI in gastric acid adaptation (18). In a gerbil model, infection and colonization with UreI deletion mutants were achieved when infection was carried out in the presence of inhibitors of the gastric H,K-ATPase, and colonization disappeared when the inhibitor was no longer administered. These results suggest that urease activity is required for both infection and colonization (G. Hanauer and K. Melchers, unpublished observations). A model for the function of UreI is shown in Fig. 5.

Conclusions

H. pylori shows a remarkable acid-resistance response to the gastric environment by a specialized means of utilizing an intracellular, neutral pH optimum urease. Within the urease gene cluster it has a gene encoding an acid-activated urea channel, UreI. Instead of an acidic pH optimum urease suitable only for passage through the stomach, it has a neutral pH optimum urease that allows the cytoplasm to remain relatively neutral while the organism protects itself against acid by enhancing urea access to intrabacterial urease via Urel and achieving buffering of the periplasm. This maintains an adequate PMF for survival and for growth in an otherwise damaging environment.

There are undoubtedly other subtleties involved in the regulation of urease activity inside the organism as well as the NH3 efflux across the inner membrane to buffer the periplasm. For example, at low rates of intrabacterial urease activity, the NH3 produced will not result in significant cytoplasmic alkalinization. This is the situation in the absence of acid activation of Urel. However, at a pH where urea transport is fully activated, NH3 production intracellularly probably increases by more than two orders of magnitude. This rapid production of NH3 could alkalinize the cytoplasm even in the presence of NH3 efflux through a bilayer with permeability of 10−1 cm s−1. However, if NH3 efflux were enabled through the urea channel, the combination of NH3 efflux across the bilayer and through UreI could maintain intrabacterial pH near neutrality in the face of high intrabacterial urease activity. Further, as soon as the periplasmic pH reaches 6.5, urea entry is switched off, thus inactivating intra-cellular urease.

It would seem possible that this mechanism of H. pylori adaptation to life in the human stomach may form the basis for methods designed to eradicate the organism. Presumably, drugs could be developed targeting the periplasmic domain of UreI and preventing acid activation of urea transport. If the bacterium is exposed, even for a short time, to a pH of less than 4.0, interference with urea uptake at these acidic pH values could prove lethal to the organism.

References

1.
Andrutis K. A., Fox J. G., Schauer D. B., Marini R. P., Murphy J. C., Yan L., Solnick J. V. Inability of an isogenic urease-negative mutant stain of Helicobacter mustelae to colonize the ferret stomach. Infect. Immun. 1995; 63: 37223725. [PubMed] [Free Full Text in PMC icon.Free Full text in PMC]
2.
Athmann C., Zeng N., Kang T., Marcus E. A., Scott D. R., Rektorschek M., Buhmann A., Melchers K., Sachs G. Sites of pH elevation due to NH3 generation by the intra-bacterial urease of Helicobacter pylori co-cultured with gastric cells. J. Clin. Invest. 2000; 106: 339347. [PubMed] [Free Full Text in PMC icon.Free Full text in PMC]
3.
Blaser M. J. Hypothesis: the changing relationships of Helicobacter pylori and humans: implications for health and disease. J. Infect. Dis. 1999; 179: 15231530. [PubMed]
4.
Bode G., Malfertheimer P., Lenhardt G., Nilius M., Dischuneit H. Ultrastructural localization of urease of Helicobacter pylori. Med. Microbiol. Immun. 1993; 182: 223242. [PubMed]
5.
Dimroth P. Bacterial sodium ion-coupled energetics. Antonie Van Leeuwenhoek. 1994; 65: 381395. [PubMed]
6.
Eaton K. A., Krakowka S. Effect of gastric pH on urease-dependent colonization of gnotobiotic piglets by Helicobacter pylori. Infect. Immun. 1994; 62: 36043607. [PubMed] [Free Full Text in PMC icon.Free Full text in PMC]
7.
Labigne A., Cussac V., Courcoux P. Shuttle cloning and nucleotide sequences of Helicobacter pylori genes responsible for urease activity. J. Bacteriol. 1991; 173: 19201931. [PubMed] [Free Full Text in PMC icon.Free Full text in PMC]
8.
Matin A. Bioenergetic parameters and transport in obligate acidophiles. Biochim. Biophys. Acta. 1990; 108: 267270.
9.
Meyer-Rosberg K., Scott D. R., Rex D., Melchers K., Sachs G. The effect of environmental pH on the proton motive force of Helicobacter pylori. Gastroenterology. 1996; 111: 886900. [PubMed]
10.
Mitchell P. Chemiosmotic coupling in oxidative and photosynthetic phosphorylation. Biol. Rev. 1966; 41: 445502. [PubMed]
11.
Mobley H. L. T., Island M. D., Hausinger R. P. Molecular biology of microbial ureases. Microbiol. Rev. 1995; 59: 451480. [PubMed] [Free Full Text in PMC icon.Free Full text in PMC]
12.
Parsonnet J., Friedman G. D., Vandersteed D. P., Chang Y., Vogelman J. H., Sibley R. K. H. pylori infection and the risk of gastric cancer. New Engl. J. Med. 1991; 325: 11311134. [PubMed]
13.
Phadnis S. H., Parlow M. H., Levy M., Ilver D., Caulkins C. M., Connors J. B., Dunn B. E. Surface localization of Helicobacter pylori urease and a heat shock protein homolog requires bacterial lysis. Infect. Immun. 1996; 64: 905912. [PubMed] [Free Full Text in PMC icon.Free Full text in PMC]
14.
Rektorscheck M., Buhmann A., Weeks D., Schwan D., Bensch K. W., Eskandari S., Sachs G., Melchers K. Acid resistance of Helicobacter pylori depends on the UreI membrane protein and an inner membrane proton barrier. Mol. Microbiol. 2000; 36: 141152. [PubMed]
15.
Rektorscheck M., Weeks D., Sachs G., Melchers K. Influence of pH on metabolism and urease activity of Helicobacter pylori. Gastroenterology. 1998; 115: 628641. [PubMed]
16.
Scott D. R., Marcus E. A., Weeks D. L., Lee A., Melchers K., Sachs G. Expression of the Helicobacter pylori ureI gene is required for acidic pH activation of aytoplasmic urease. Infect. Immun. 2000; 68: 470477. [PubMed] [Free Full Text in PMC icon.Free Full text in PMC]
17.
Scott D. R., Weeks D., Hong C., Postius S., Melchers K., Sachs G. The role of internal urease in acid resistance of Helicobacter pylori. Gastroenterology. 1998; 114: 5870. [PubMed]
18.
Skouloubris S., Thiberge J., Labigne A., de Reuse H. The Helicobacter pylori UreI protein is not involved in urease activity but is essential for bacterial survival in vivo. Infect. Immun. 1998; 66: 45174521. [PubMed] [Free Full Text in PMC icon.Free Full text in PMC]
19.
Skulachev V. P. Chemisomotic concept of membrane bioenergetics: what is already clear and what is still waiting for elucidation. J. Bioenerg. Biomembr. 1994; 26: 589598. [PubMed]
20.
Tsuda M., Karita M., Morshed M. G., Okita K., Nakazawa T. A urease-negative mutant of Helicobacter pylori constructed by allelic exchange mutagenesis lacks the ability to colonize the nude mouse stomach. Infect. Immun. 1994; 62: 35863589. [PubMed] [Free Full Text in PMC icon.Free Full text in PMC]
21.
Voland P., Sachs G. Identification of the interaction of UreG/UreE and UreF/UreH in H. pylori using a yeast two-hybrid screen. Gastroenterology. 2000; 118: A-736.
22.
Warren J. R., Marshall B. Unidentified curved bacilli on gastric epithelium in active chronic gastritis. Lancet. 1983; i: 12731275. [PubMed]
23.
Weeks D. L., Eskandari S., Scott D. R., Sachs G. A H+ gated urea channel: the link between Helicobacter pylori urease and gastric colonization. Science. 2000; 287: 482485. [PubMed]
24.
Young G. M., Amid D., Miller V. L. A bifunctional urease enhances survival of pathogenic Yersinia enterocolitica and Morganella morganii at low pH. J. Bacteriol. 1996; 178: 64876495. [PubMed] [Free Full Text in PMC icon.Free Full text in PMC]

 

Next
Helicobacter pylori2001
(navigation arrows) Go to previous chapter Go to next chapter Go to top of this page Go to bottom of this page Go to Table of Contents