首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 15 毫秒
1.
The roles that accessory gene products play in activating the Helicobacter pylori urease apoprotein were examined. The activity of the urease apoprotein increased in the following order when it was expressed with the accessory genes: ureG相似文献   

2.
ureI encodes an inner membrane protein of Helicobacter pylori. The role of the bacterial inner membrane and UreI in acid protection and regulation of cytoplasmic urease activity in the gastric microorganism was studied. The irreversible inhibition of urease when the organism was exposed to a protonophore (3,3',4', 5-tetrachlorsalicylanide; TCS) at acidic pH showed that the inner membrane protected urease from acid. Isogenic ureI knockout mutants of several H. pylori strains were constructed by replacing the ureI gene of the urease gene cluster with a promoterless kanamycin resistance marker gene (kanR). Mutants carrying the modified ureAB-kanR-EFGH operon all showed wild-type levels of urease activity at neutral pH in vitro. The mutants resisted media of pH > 4.0 but not of pH < 4.0. Whereas wild-type bacteria showed high levels of urease activity below pH 4.0, this ability was not retained in the ureI mutants, resulting in inhibition of metabolism and cell death. Gene complementation experiments with plasmid-derived H. pylori ureI restored wild-type properties. The activation of urease activity found in structurally intact but permeabilized bacteria treated with 0.01% detergent (polyoxy-ethylene-8-laurylether; C12E8), suggested a membrane-limited access of urea to internal urease at neutral pH. Measurement of 14C-urea uptake into Xenopus oocytes injected with ureI cRNA showed acid activation of uptake only in injected oocytes. Acceleration of urea uptake by UreI therefore mediates the increase of intracellular urease activity seen under acidic conditions. This increase of urea permeability is essential for H. pylori survival in environments below pH 4.0. ureI-independent urease activity may be sufficient for maintenance of bacterial viability above pH 4.0.  相似文献   

3.
4.
The Helicobacter pylori ureE gene product was previously shown to be required for urease expression, but its characteristics and role have not been determined. The UreE protein has now been overexpressed in Escherichia coli, purified, and characterized, and three altered versions were expressed to address a nickel-sequestering role of UreE. Purified UreE formed a dimer in solution and was capable of binding one nickel ion per dimer. Introduction of an extra copy of ureE into the chromosome of mutants carrying mutations in the Ni maturation proteins HypA and HypB resulted in partial restoration of urease activity (up to 24% of the wild-type levels). Fusion proteins of UreE with increased ability to bind nickel were constructed by adding histidine-rich sequences (His-6 or His-10 to the C terminus and His-10 as a sandwich fusion) to the UreE protein. Each fusion protein was overexpressed in E. coli and purified, and its nickel-binding capacity and affinity were determined. Each construct was also expressed in wild-type H. pylori and in hypA and hypB mutant strains for determining in vivo urease activities. The urease activity was increased by introduction of all the engineered versions, with the greatest Ni-sequestering version (the His-6 version) also conferring the greatest urease activity on both the hypA and hypB mutants. The differences in urease activities were not due to differences in the amounts of urease peptides. Addition of His-6 to another expressed protein (triose phosphate isomerase) did not result in stimulation of urease, so urease activation is not related to the level of nonspecific protein-bound nickel. The results indicate a correlation between H. pylori urease activity and the nickel-sequestering ability of the UreE accessory protein.  相似文献   

5.
Moderate levels of urease activity (ca. 300 mU mg(-1)) were detected in Rhizobium leguminosarum bv. viciae UPM791 vegetative cells. This activity did not require urea for induction and was partially repressed by the addition of ammonium into the medium. Lower levels of urease activity (ca. 100 mU mg(-1)) were detected also in pea bacteroids. A DNA region of ca. 9 kb containing the urease structural genes ( ureA, ureB and ureC), accessory genes ( ureD, ureE, ureF, and ureG), and five additional ORFs ( orf83, orf135, orf207, orf223, and orf287) encoding proteins of unknown function was sequenced. Three of these ORFs ( orf83, orf135 and orf207) have a homologous counterpart in a gene cluster from Sinorhizobium meliloti, reported to be involved in urease and hydrogenase activities. R. leguminosarum mutant strains carrying Tn 5 insertions within this region exhibited a urease-negative phenotype, but induced wild-type levels of hydrogenase and nitrogenase activities in bacteroids. orf287 encodes a potential transmembrane protein with a C-terminal GGDEF domain. A mutant affected in orf287 exhibited normal levels of urease activity in culture cells. Experiments aimed at cross-complementing Ni-binding proteins required for urease and hydrogenase synthesis (UreE and HypB, respectively) indicated that these two proteins are not functionally interchangeable in R. leguminosarum.  相似文献   

6.
7.
Ureaplasma urealyticum is a pathogenic ureolytic mollicute which colonizes the urogenital tracts of humans. A genetic polymorphism between the two biotypes of U. urealyticum at the level of the urease genes was found. The urease gene cluster from a biotype 1 representative of U. urealyticum (serotype I) was cloned and sequenced. Seven genes were found, with ureA, ureB, and ureC encoding the structural subunits and ureE, ureF, ureG, and a truncated ureI) gene encoding accessory proteins. Urease expression was not obtained when the plasmid containing these genes was incorporated into an opal suppressor strain of Escherichia coli, although this enzymatic activity was found in the same E. coli strain transformed with pC6b, a plasmid with previously cloned urease genes from the U. urealyticum T960 strain of biotype 2 (serotype 8). Although there are 12 TGA triplets encoding tryptophan within urease genes, the level of expression obtained was comparable to the levels reported for other bacterial genes expressed in E. coli. Nested deletion experiments allowed us to demonstrate that ureD is necessary for urease activity whereas another open reading frame located downstream is not. The promoter for ureA and possibly other urease genes was identified for both serotypes.  相似文献   

8.
9.
10.
11.
12.
13.
The enzymatic activity of Helicobacter pylori's urease neutralises stomach acidity, thereby promoting infection by this pathogen. Urease protein has also been found to interact with host cells in vitro, although this property's possible functional importance has not been studied in vivo. To test for a role of the urease surface in the host/pathogen interaction, surface exposed loops that display high thermal mobility were targeted for inframe insertion mutagenesis. H. pylori expressing urease with insertions at four of eight sites tested retained urease activity, which in three cases was at least as stable as was wild-type urease at pH 3. Bacteria expressing one of these four mutant ureases, however, failed to colonise mice for even two weeks, and a second had reduced bacterial titres after longer term (3 to 6 months) colonisation. These results indicate that a discrete surface of the urease complex is important for H. pylori persistence during gastric colonisation. We propose that this surface interacts directly with host components important for the host-pathogen interaction, immune modulation or other actions that underlie H. pylori persistence in its special gastric mucosal niche.  相似文献   

14.
15.
16.
17.
18.
The urease accessory protein encoded by ureE from Klebsiella aerogenes is proposed to deliver Ni(II) to the urease apoprotein during enzyme activation. Native UreE possesses a histidine-rich region at its carboxyl terminus that binds several equivalents of Ni(2+); however, a truncated form of this protein (H144*UreE) binds only 2 Ni(2+) per dimer and is functionally active (Brayman, T. G., and Hausinger, R. P. (1996) J. Bacteriol. 178, 5410-5416). The urease activation kinetics were studied in vivo by monitoring the development of urease activity upon adding Ni(2+) to spectinomycin-treated Escherichia coli cells that expressed the complete K. aerogenes urease gene cluster with altered forms of ureE. Site-specific alterations of H144*UreE decrease the rate of in vivo urease activation, with the most dramatic changes observed for the H96A, H110A, D111A, and H112A substitutions. Notably, urease activity in cells producing H96A/H144*UreE was lower than cells containing a ureE deletion. Prior studies had shown that H110A and H112A variants each bound a single Ni(2+) per dimer with elevated K(d) values compared with control H144*UreE, whereas the H96A and D111A variants bound 2 Ni(2+) per dimer with unperturbed K(d) values (Colpas, G. J., Brayman, T. G., Ming, L.-J., and Hausinger, R. P. (1999) Biochemistry 38, 4078-4088). To understand why cells containing the latter two proteins showed reduced rates of urease activation, we characterized their metal binding/dissociation kinetics and compared the results to those obtained for H144*UreE. The truncated protein was shown to sequentially bind two Ni(2+) with k(1) approximately 18 and k(2) approximately 100 M(-1) s(-1), and with dissociation rates k(-1) approximately 3 x 10(-3) and k(-2) approximately 10(-4) s(-1). Similar apparent rates of binding and dissociation were noted for the two mutant proteins, suggesting that altered H144*UreE interactions with Ni(2+) do not account for the changes in cellular urease activation. These conclusions are further supported by in vitro experiments demonstrating that addition of H144*UreE to urease apoprotein activation mixtures inhibited the rate and extent of urease formation. Our results highlight the importance of other urease accessory proteins in assisting UreE-dependent urease maturation.  相似文献   

19.
Arginase of the Helicobacter pylori urea cycle hydrolyzes L-arginine to L-ornithine and urea. H. pylori urease hydrolyzes urea to carbon dioxide and ammonium, which neutralizes acid. Both enzymes are involved in H. pylori nitrogen metabolism. The roles of arginase in the physiology of H. pylori were investigated in vitro and in vivo, since arginase in H. pylori is metabolically upstream of urease and urease is known to be required for colonization of animal models by the bacterium. The H. pylori gene hp1399, which is orthologous to the Bacillus subtilis rocF gene encoding arginase, was cloned, and isogenic allelic exchange mutants of three H. pylori strains were made by using two different constructs: 236-2 and rocF::aphA3. In contrast to wild-type (WT) strains, all rocF mutants were devoid of arginase activity and had diminished serine dehydratase activity, an enzyme activity which generates ammonium. Compared with WT strain 26695 of H. pylori, the rocF::aphA3 mutant was approximately 1, 000-fold more sensitive to acid exposure. The acid sensitivity of the rocF::aphA3 mutant was not reversed by the addition of L-arginine, in contrast to the WT, and yielded a approximately 10, 000-fold difference in viability. Urease activity was similar in both strains and both survived acid exposure equally well when exogenous urea was added, indicating that rocF is not required for urease activity in vitro. Finally, H. pylori mouse-adapted strain SS1 and the 236-2 rocF isogenic mutant colonized mice equally well: 8 of 9 versus 9 of 11 mice, respectively. However, the rocF::aphA3 mutant of strain SS1 had moderately reduced colonization (4 of 10 mice). The geometric mean levels of H. pylori recovered from these mice (in log(10) CFU) were 6.1, 5.5, and 4.1, respectively. Thus, H. pylori rocF is required for arginase activity and is crucial for acid protection in vitro but is not essential for in vivo colonization of mice or for urease activity.  相似文献   

20.
Inducible NO synthase (iNOS) expression and production of NO are both up-regulated with Helicobacter pylori infection in vivo and in vitro. We determined whether major pathogenicity proteins released by H. pylori activate iNOS by coculturing macrophages with wild-type or mutant strains deficient in VacA, CagA, picB product, or urease (ureA(-)). When filters were used to separate H. pylori from macrophages, there was a selective and significant decrease in stimulated iNOS mRNA, protein, and NO(2)(-) production with the ureA(-) strain compared with wild-type and other mutants. Similarly, macrophage NO(2)(-) generation was increased by H. pylori protein water extracts of all strains except ureA(-). Recombinant urease stimulated significant increases in macrophage iNOS expression and NO(2)(-) production. Taken together, these findings indicate a new role for the essential H. pylori survival factor, urease, implicating it in NO-dependent mucosal damage and carcinogenesis.  相似文献   

设为首页 | 免责声明 | 关于勤云 | 加入收藏

Copyright©北京勤云科技发展有限公司  京ICP备09084417号