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 93 毫秒
1.
Control of Acid Resistance in Escherichia coli   总被引:4,自引:0,他引:4       下载免费PDF全文
Acid resistance (AR) in Escherichia coli is defined as the ability to withstand an acid challenge of pH 2.5 or less and is a trait generally restricted to stationary-phase cells. Earlier reports described three AR systems in E. coli. In the present study, the genetics and control of these three systems have been more clearly defined. Expression of the first AR system (designated the oxidative or glucose-repressed AR system) was previously shown to require the alternative sigma factor RpoS. Consistent with glucose repression, this system also proved to be dependent in many situations on the cyclic AMP receptor protein. The second AR system required the addition of arginine during pH 2.5 acid challenge, the structural gene for arginine decarboxylase (adiA), and the regulator cysB, confirming earlier reports. The third AR system required glutamate for protection at pH 2.5, one of two genes encoding glutamate decarboxylase (gadA or gadB), and the gene encoding the putative glutamate:gamma-aminobutyric acid antiporter (gadC). Only one of the two glutamate decarboxylases was needed for protection at pH 2.5. However, survival at pH 2 required both glutamate decarboxylase isozymes. Stationary phase and acid pH regulation of the gad genes proved separable. Stationary-phase induction of gadA and gadB required the alternative sigma factor sigmaS encoded by rpoS. However, acid induction of these enzymes, which was demonstrated to occur in exponential- and stationary-phase cells, proved to be sigmaS independent. Neither gad gene required the presence of volatile fatty acids for induction. The data also indicate that AR via the amino acid decarboxylase systems requires more than an inducible decarboxylase and antiporter. Another surprising finding was that the sigmaS-dependent oxidative system, originally thought to be acid induced, actually proved to be induced following entry into stationary phase regardless of the pH. However, an inhibitor produced at pH 8 somehow interferes with the activity of this system, giving the illusion of acid induction. The results also revealed that the AR system affording the most effective protection at pH 2 in complex medium (either Luria-Bertani broth or brain heart infusion broth plus 0.4% glucose) is the glutamate-dependent GAD system. Thus, E. coli possesses three overlapping acid survival systems whose various levels of control and differing requirements for activity ensure that at least one system will be available to protect the stationary-phase cell under naturally occurring acidic environments.  相似文献   

2.
Several members of the family Enterobacteriaceae were examined for differences in extreme acid survival strategies. A surprising degree of variety was found between three related genera. The minimum growth pH of Salmonella typhimurium was shown to be significantly lower (pH 4.0) than that of either Escherichia coli (pH 4.4) or Shigella flexneri (pH 4.8), yet E. coli and S. flexneri both survive exposure to lower pH levels (2 to 2.5) than S. typhimurium (pH 3.0) in complex medium. S. typhimurium and E. coli but not S. flexneri expressed low-pH-inducible log-phase and stationary-phase acid tolerance response (ATR) systems that function in minimal or complex medium to protect cells to pH 3.0. All of the organisms also expressed a pH-independent general stress resistance system that contributed to acid survival during stationary phase. E. coli and S. flexneri possessed several acid survival systems (termed acid resistance [AR]) that were not demonstrable in S. typhimurium. These additional AR systems protected cells to pH 2.5 and below but required supplementation of minimal medium for either induction or function. One acid-inducible AR system required oxidative growth in complex medium for expression but successfully protected cells to pH 2.5 in unsupplemented minimal medium, while two other AR systems important for fermentatively grown cells required the addition of either glutamate or arginine during pH 2.5 acid challenge. The arginine AR system was only observed in E. coli and required stationary-phase induction in acidified complex medium. The product of the adi locus, arginine decarboxylase, was responsible for arginine-based acid survival.  相似文献   

3.
Escherichia coli O157:H7 is a highly acid-resistant food-borne pathogen that survives in the bovine and human gastrointestinal tracts and in acidic foods such as apple cider. This property is thought to contribute to the low infectious dose of the organism. Three acid resistance (AR) systems are expressed in stationary-phase cells. AR system 1 is sigma(S) dependent, while AR systems 2 and 3 are glutamate and arginine dependent, respectively. In this study, we sought to determine which AR systems are important for survival in acidic foods and which are required for survival in the bovine intestinal tract. Wild-type and mutant E. coli O157:H7 strains deficient in AR system 1, 2, or 3 were challenged with apple cider and inoculated into calves. Wild-type cells, adapted at pH 5.5 in the absence of glucose (AR system 1 induced), survived well in apple cider. Conversely, the mutant deficient in AR system 1, shown previously to survive poorly in calves, was susceptible to apple cider (pH 3.5), and this sensitivity was shown to be caused by low pH. Interestingly, the AR system 2-deficient mutant survived in apple cider at high levels, but its shedding from calves was significantly decreased compared to that of wild-type cells. AR system 3-deficient cells survived well in both apple cider and calves. Taken together, these results indicate that E. coli O157:H7 utilizes different acid resistance systems based on the type of acidic environment encountered.  相似文献   

4.
Escherichia coli O157:H7 is a highly acid-resistant food-borne pathogen that survives in the bovine and human gastrointestinal tracts and in acidic foods such as apple cider. This property is thought to contribute to the low infectious dose of the organism. Three acid resistance (AR) systems are expressed in stationary-phase cells. AR system 1 is σS dependent, while AR systems 2 and 3 are glutamate and arginine dependent, respectively. In this study, we sought to determine which AR systems are important for survival in acidic foods and which are required for survival in the bovine intestinal tract. Wild-type and mutant E. coli O157:H7 strains deficient in AR system 1, 2, or 3 were challenged with apple cider and inoculated into calves. Wild-type cells, adapted at pH 5.5 in the absence of glucose (AR system 1 induced), survived well in apple cider. Conversely, the mutant deficient in AR system 1, shown previously to survive poorly in calves, was susceptible to apple cider (pH 3.5), and this sensitivity was shown to be caused by low pH. Interestingly, the AR system 2-deficient mutant survived in apple cider at high levels, but its shedding from calves was significantly decreased compared to that of wild-type cells. AR system 3-deficient cells survived well in both apple cider and calves. Taken together, these results indicate that E. coli O157:H7 utilizes different acid resistance systems based on the type of acidic environment encountered.  相似文献   

5.
This study demonstrates that exposure of log-phase Lactococcus lactis subsp. cremoris 712 cells to mildly acid conditions induces resistance to normally lethal intensities of environmental stresses such as acid, heat, NaCl, H2O2, and ethanol. The intracellular pH (pHi) played a major role in the induction of this multistress resistance response. The pHi was dependent on the extracellular pH (pHo) and on the specific acid used to reduce the pHo. When resuspended in fresh medium, cells were able to maintain a pH gradient even at pHo values that resulted in cell death. Induction of an acid tolerance response (ATR) coincided with an increase in the ability of cells to resist change to an unfavorable pHi; nevertheless, a more favorable pHi was not the sole reason for the increased survival at acid pHo. Cells with an induced ATR survived exposure to a lethal pHo much better than did uninduced cells with a pHi identical to that of the induced cells. Survival following lethal acid shock was dependent on the pHi during induction of the ATR, and the highest survival was observed following induction at a pHi of 5.9, which was the lowest pHi at which growth occurred. Increased acid tolerance and the ability to maintain a higher pHi during lethal acid stress were not acquired if protein synthesis was inhibited by chloramphenicol during adaptation.  相似文献   

6.
AIMS: The objective of this study was to investigate the effect of growing conditions on the glutamate-, arginine- and lysine-dependent acid resistance (AR) systems of Escherichia coli O157:H7. METHODS AND RESULTS: Seven E. coli O157:H7 strains were grown in five different media at neutral or acidic pH under aerobic or anaerobic conditions, and the survival rate after acid shocks (pH 2.0, 1 h, 37 degrees C) in the presence of glutamate, arginine and lysine was determined. Six strains induced the glutamate-dependent AR at stationary phase, and maximal survival were observed (> or =10%) when grown in pH 5- Luria-Bertani media with glucose (LBG) and in pH 4.5-anaerobic media. The arginine- and lysine-dependent systems were also present, but were only induced if cells had grown in LBG. For strain ATCC 43895, the minimum glutamate concentration that resulted in at least 10% survival rate was 10 micromol l(-1), but it required at least 10-fold more arginine and lysine. CONCLUSIONS: The lysine-dependent AR system could be as important as the arginine-mediated one, but the contribution of both systems to E. coli O157:H7 overall AR response might be minor compared with the glutamate-dependent system. SIGNIFICANCE AND IMPACT OF THE STUDY: Under typical environmental conditions, the glutamate-dependent AR system might be solely responsible for protecting cells against acidic pH.  相似文献   

7.
Commensal and pathogenic strains of Escherichia coli possess three inducible acid resistance systems that collaboratively protect cells against acid stress to pH 2 or below. The most effective system requires glutamate in the acid challenge media and relies on two glutamate decarboxylases (GadA and B) combined with a putative glutamate:gamma-aminobutyric acid antiporter (GadC). A complex network of regulators mediates induction of this system in response to various media, pH and growth phase signals. We report that the LuxR-like regulator GadE (formerly YhiE) is required for expression of gadA and gadBC regardless of media or growth conditions. This protein binds directly to the 20 bp GAD box sequence found in the control regions of both loci. Two previously identified AraC-like regulators, GadX and GadW, are only needed for gadA/BC expression under some circumstances. Overexpression of GadX or GadW will not overcome a need for GadE. However, overexpression of GadE can supplant a requirement for GadX and W. Data provided also indicate that GadX and GadE can simultaneously bind the area around the GAD box region and probably form a complex. The gadA, gadBC and gadE genes are all induced by low pH in exponential phase cells grown in minimal glucose media. The acid induction of gadA/BC results primarily from the acid induction of gadE. Constitutive expression of GadE removes most pH control over the glutamate decarboxylase and antiporter genes. The small amount of remaining pH control is governed by GadX and W. The finding that gadE mutations also diminish the effectiveness of the other two acid resistance systems suggests that GadE influences the expression of additional acid resistance components. The number of regulatory proteins (five), sigma factors (two) and regulatory feedback loops focused on gadA/BC expression make this one of the most intensively regulated systems in E. coli.  相似文献   

8.
Enterohemorrhagic strains of Escherichia coli must pass through the acidic gastric barrier to cause gastrointestinal disease. Taking into account the apparent low infectious dose of enterohemorrhagic E. coli, 11 O157:H7 strains and 4 commensal strains of E. coli were tested for their abilities to survive extreme acid exposures (pH 3). Three previously characterized acid resistance systems were tested. These included an acid-induced oxidative system, an acid-induced arginine-dependent system, and a glutamate-dependent system. When challenged at pH 2.0, the arginine-dependent system provided more protection in the EHEC strains than in commensal strains. However, the glutamate-dependent system provided better protection than the arginine system and appeared equally effective in all strains. Because E. coli must also endure acid stress imposed by the presence of weak acids in intestinal contents at a pH less acidic than that of the stomach, the ability of specific acid resistance systems to protect against weak acids was examined. The arginine- and glutamate-dependent systems were both effective in protecting E. coli against the bactericidal effects of a variety of weak acids. The acids tested include benzoic acid (20 mM; pH 4.0) and a volatile fatty acid cocktail composed of acetic, propionic, and butyric acids at levels approximating those present in the intestine. The oxidative system was much less effective. Several genetic aspects of E. coli acid resistance were also characterized. The alternate sigma factor RpoS was shown to be required for oxidative acid resistance but was only partially involved with the arginine- and glutamate-dependent acid resistance systems. The arginine decarboxylase system (including adi and its regulators cysB and adiY) was responsible for arginine-dependent acid resistance. The results suggest that several acid resistance systems potentially contribute to the survival of pathogenic E. coli in the different acid stress environments of the stomach (pH 1 to 3) and the intestine (pH 4.5 to 7 with high concentrations of volatile fatty acids). Of particular importance to the food industry was the finding that once induced, the acid resistance systems will remain active for prolonged periods of cold storage at 4 degrees C.  相似文献   

9.
Pathogenic strains of Escherichia coli, such as E. coli O157:H7, have a low infectious dose and an ability to survive in acidic foods. These bacteria have evolved at least three distinct mechanisms of acid resistance (AR), including two amino acid decarboxylase-dependent systems (arginine and glutamate) and a glucose catabolite-repressed system. We quantified the survival rates for each AR mechanism separately in clinical isolates representing three groups of Shiga toxin-producing E. coli (STEC) clones (O157:H7, O26:H11/O111:H8, and O121:H19) and six commensal strains from ECOR group A. Members of the STEC clones were not significantly more acid resistant than the commensal strains when analyzed using any individual AR mechanism. The glutamate system provided the best protection in a highly acidic environment for all groups of isolates (<0.1 log reduction in CFU/ml per hour at pH 2.0). Under these conditions, there was notable variation in survival rates among the 30 O157:H7 strains, which depended in part on Mg(2+) concentration. The arginine system provided better protection at pH 2.5, with a range of 0.03 to 0.41 log reduction per hour, compared to the oxidative system, with a range of 0.13 to 0.64 log reduction per hour. The average survival rate for the O157:H7 clonal group was significantly less than that of the other STEC clones in the glutamate and arginine systems and significantly less than that of the O26/O111 clone in the oxidative system, indicating that this clonal group is not exceptionally acid resistant with these specific mechanisms.  相似文献   

10.
The mechanistic target of rapamycin (mTOR) complex 2 (mTORC2) signaling controls cell metabolism, promotes cell survival, and contributes to tumorigenesis, yet its upstream regulation remains poorly defined. Although considerable evidence supports the prevailing view that amino acids activate mTOR complex 1 but not mTORC2, several studies reported paradoxical activation of mTORC2 signaling by amino acids. We noted that after amino acid starvation of cells in culture, addition of an amino acid solution increased mTORC2 signaling. Interestingly, we found the pH of the amino acid solution to be alkaline, ∼pH 10. These observations led us to discover and demonstrate here that alkaline intracellular pH (pHi) represents a previously unknown activator of mTORC2. Using a fluorescent pH-sensitive dye (cSNARF1-AM) coupled with live-cell imaging, we demonstrate that culturing cells in media at an alkaline pH induces a rapid rise in the pHi, which increases mTORC2 catalytic activity and downstream signaling to the pro-growth and pro-survival kinase Akt. Alkaline pHi also activates AMPK, a canonical sensor of energetic stress. Functionally, alkaline pHi activates AMPK-mTOR signaling, which attenuates apoptosis caused by growth factor withdrawal. Collectively, these findings reveal that alkaline pHi increases mTORC2- and AMPK-mediated signaling to promote cell survival during conditions of growth factor limitation, analogous to the demonstrated ability of energetic stress to activate AMPK–mTORC2 and promote cell survival. As an elevated pHi represents an underappreciated hallmark of cancer cells, we propose that the alkaline pHi stress sensing by AMPK–mTORC2 may contribute to tumorigenesis by enabling cancer cells at the core of a growing tumor to evade apoptosis and survive.  相似文献   

11.
Escherichia coli MG1655 acid-inducible genes were identified by whole-genome expression profiling. Cultures were grown to the mid-logarithmic phase on acidified glucose minimal medium, conditions that induce glutamate-dependent acid resistance (AR), while the other AR systems are either repressed or not induced. A total of 28 genes were induced in at least two of three experiments in which the gene expression profiles of cells grown in acid (pH 5.5 or 4.5) were compared to those of cells grown at pH 7.4. As expected, the genes encoding glutamate decarboxylase, gadA and gadB, were significantly induced. Interestingly, two acid-inducible genes code for small basic proteins with pIs of >10.5, and six code for small acidic proteins with pIs ranging from 5.7 to 4.0; the roles of these small basic and acidic proteins in acid resistance are unknown. The acid-induced genes represented only five functional grouping categories, including eight genes involved in metabolism, nine associated with cell envelope structures or modifications, two encoding chaperones, six regulatory genes, and six unknown genes. It is unlikely that all of these genes are involved in the glutamate-dependent AR. However, nine acid-inducible genes are clustered in the gadA region, including hdeA, which encodes a putative periplasmic chaperone, and four putative regulatory genes. One of these putative regulators, yhiE, was shown to significantly increase acid resistance when overexpressed in cells that had not been preinduced by growth at pH 5.5, and mutation of yhiE decreased acid resistance; yhiE could therefore encode an activator of AR genes. Thus, the acid-inducible genes clustered in the gadA region appear to be involved in glutatmate-dependent acid resistance, although their specific roles remain to be elucidated.  相似文献   

12.
Pathogenic strains of Escherichia coli, such as E. coli O157:H7, have a low infectious dose and an ability to survive in acidic foods. These bacteria have evolved at least three distinct mechanisms of acid resistance (AR), including two amino acid decarboxylase-dependent systems (arginine and glutamate) and a glucose catabolite-repressed system. We quantified the survival rates for each AR mechanism separately in clinical isolates representing three groups of Shiga toxin-producing E. coli (STEC) clones (O157:H7, O26:H11/O111:H8, and O121:H19) and six commensal strains from ECOR group A. Members of the STEC clones were not significantly more acid resistant than the commensal strains when analyzed using any individual AR mechanism. The glutamate system provided the best protection in a highly acidic environment for all groups of isolates (<0.1 log reduction in CFU/ml per hour at pH 2.0). Under these conditions, there was notable variation in survival rates among the 30 O157:H7 strains, which depended in part on Mg2+ concentration. The arginine system provided better protection at pH 2.5, with a range of 0.03 to 0.41 log reduction per hour, compared to the oxidative system, with a range of 0.13 to 0.64 log reduction per hour. The average survival rate for the O157:H7 clonal group was significantly less than that of the other STEC clones in the glutamate and arginine systems and significantly less than that of the O26/O111 clone in the oxidative system, indicating that this clonal group is not exceptionally acid resistant with these specific mechanisms.  相似文献   

13.
Induction of acid tolerance response (ATR) of exponential-phase Escherichia coli K-12 cells grown and adapted at different conditions was examined. The highest level of protection against pH 2.5 challenges was obtained after adaptation at pH 4.5-4.9 for 60 min. To study the genetic systems, which could be involved in the development of log-phase ATR, we investigated the acid response of E. coli acid resistance (AR) mutants. The activity of the glutamate-dependent system was observed in exponential cells grown at pH 7.0 and acid adapted at pH 4.5 in minimal medium. Importantly, log-phase cells exhibited significant AR when grown in minimal medium pH 7.0 and challenged at pH 2.5 for 2 h without adaptation. This AR required the glutamate-dependent AR system. Acid protection was largely dependent on RpoS in unadapted and adapted cells grown in minimal medium. RpoS-dependent oxidative, glutamate and arginine-dependent decarboxylase AR systems were not involved in triggering log-phase ATR in cells grown in rich medium. Cells adapted at pH 4.5 in rich medium showed a higher proton accumulation rate than unadapted cells as determined by proton flux assay. It is clear from our study that highly efficient mechanisms of protection are induced, operate and play the main role during log-phase ATR.  相似文献   

14.
Elevated pCO(2) inhibits cell growth. This growth inhibition is accompanied by a decrease in intracellular pH (pHi), as well as a decrease in glycolysis. Elevated concentrations (mM) of some amino acids have been shown by others to protect cells exposed to two very different environmental stresses: nutrient starvation and hyperosmolality. The fact that many of the amino acids shown to have protective effects against other stresses are transported into the cell through a pHi-sensitive transporter led us to study the possibility of using these amino acids as protective agents under elevated pCO(2). Screening experiments using 5, 15, and 25 mM of each amino acid showed that not all amino acids that protect cells from hyperosmolality protect them from elevated pCO(2). Glycine betaine and glycine were chosen for further characterization in both hybridoma and CHO cells. Asparagine and threonine were also tested in hybridoma and CHO cells, respectively. All amino acids tested under 195 mm Hg pCO(2)/435 mOsm/kg (50% growth inhibition) restored the specific growth rate (mu) in hybridoma cells to that observed under control conditions (40 mm Hg/320 mOsm/kg). Addition of each amino acid resulted in an increase in the consumption rate and intracellular accumulation of that amino acid. In CHO cells, glycine betaine also restored mu to control values, while glycine and threonine partially restored mu. In hybridoma cells, the higher specific antibody productivity obtained at elevated pCO(2) was maintained with the lowest amino acid concentration (5 mM). Productivity decreased toward control values with increasing amino acid concentrations. Elevated pCO(2) decreased the specific tPA productivity in the CHO cell line studied. Only glycine betaine resulted in a 20% increase in productivity at 195 mm Hg/435 mOsm/kg. With the exception of glycine betaine in hybridoma cells, amino acids did not mitigate the associated pHi decrease of at least 0.2 pH units at 195 mm Hg/435 mOsm/kg. pHi in hybridoma cells under elevated pCO(2) in the presence of glycine betaine was about 0.1 pH units below that of control. Amino acids had no effect on the cell size response of hybridoma cells, while they partially offset the increase in CHO cell size at elevated pCO(2). Glycine betaine, asparagine, and glycine increased the specific glucose consumption rate observed at 195 mm Hg/435 mOsm/kg (50% of control) to values greater than 70% of control in hybridoma cells. In CHO cells, only glycine betaine increased q(glc) (by 20%) under elevated pCO(2). All amino acids tested improved the cell yield from glutamine at 195 mm Hg/435 mOsm/kg in both cell lines.  相似文献   

15.
16.
To study the role of intracellular pH (pHi) in catecholamine secretion and the regulation of pHi in bovine chromaffin cells, the pH-sensitive fluorescent indicator [2',7'-bis(carboxyethyl)-5(6)-carboxyfluorescein] was used to monitor the on-line changes in pHi. The pHi of chromaffin cells at resting state is approximately 7.2. The pHi was manipulated first by incubation of the cells with NH4+, and then the solution was replaced with a NH4(+)-free solution to induce acidification of the cytoplasm. The pHi returned toward the basal pH value after acidification within 5-10 min in the presence of Na+ or Li+, but the pHi stayed acidic when Na(+)-free buffers were used or in the presence of amiloride and its analogues. These results suggest that the pH recovery process after an acid load is due to the Na+/H+ exchange activity in the plasma membrane of the chromaffin cells. The catecholamine secretion evoked by carbachol and Na+ removal was enhanced after the cytoplasm had been made more acidic. It appears that acidic pH favors the occurrence of exocytosis.  相似文献   

17.
A glutamate-dependent acid resistance gene in Escherichia coli.   总被引:7,自引:0,他引:7       下载免费PDF全文
Stationary-phase cultures of Escherichia coli can survive several hours or exposure to extreme acid (pH 2 to 3), a level well below the pH range for growth (pH 4.5 to 9). To identify the genes needed for survival in extreme acid, a microliter screening procedure was devised. Colonies from a Tn10 transposon pool in E. coli MC4100 were inoculated into buffered Luria broth, pH 7.0, in microtiter wells, grown overnight, and then diluted in Luria broth, pH 2.5, at 37 degrees C for 2 h. From 3,000 isolates screened, 3 Tet(r) strains were identified as extremely acid sensitive (<0.1% survival at pH 2.5 for 2 h). Flanking sequences of the Tn10 inserts were amplified by inverse PCR. The sequences encoded a hydrophobic partial peptide of 88 residues. A random-primer-generated probe hybridized to Kohara clones 279 and 280 at 32 min (33.7 min on the revised genomic map EcoMap7) near gadB (encoding glutamate decarboxylase). The gene was designated xasA for extreme acid sensitive. xasA::Tn10 strains grown at pH 7 to 8 showed 100-fold-less survival in acid than the parent strain. Growth in mild acid (pH 5 to 6) restored acid resistance; anaerobiosis was not required, as it is for acid resistance in rpoS strains. xasA::Tn10 eliminated enhancement of acid resistance by glutamic acid. xasA was found to be a homolog of gadC recently sequenced in Shigella flexneri, in which it appears to encode a permease for the decarboxylated product of GadB. These results suggest that GadC (XasA) participates in a glutamate decarboxylase alkalinization cycle to protect E. coli from cytoplasmic acidification. The role of the glutamate cycle is particularly important for cultures grown at neutral pH before exposure to extreme acid.  相似文献   

18.
The acid tolerance response (ATR) is an adaptive system triggered at external pH (pHo) values of 5.5 to 6.0 that will protect cells from more severe acid stress (J. Foster and H. Hall, J. Bacteriol. 172:771-778, 1990). Correlations between the internal pH (pHi) of adapted versus unadapted cells at pHo of 3.3 indicate that the ATR system produces an inducible pH-homeostatic function. This function serves to maintain the pHi above 5 to 5.5. Below this range, cells rapidly lose viability. Development of this pH homeostasis mechanism was sensitive to protein synthesis inhibitors and operated only to augment the pHi at pHo values below 4. In contrast, classical constitutive pH homeostasis was insensitive to protein synthesis inhibitors and was efficient only at pHo values above 4. Physiological studies indicated an important role for the Mg(2+)-dependent proton-translocating ATPase in affording ATR-associated survival during exposure to severe acid challenges. Along with being acid intolerant, cells deficient in this ATPase did not exhibit inducible pH homeostasis. We speculate that adaptive acid tolerance is important to Salmonella species in surviving acid encounters in both the environment and the infected host.  相似文献   

19.
Interactions between intracellular pH (pHi) and H+-coupled transmembrane transport of glycine have been studied by means of 31P-NMR, using both aerobic and 'energy starved' cells of the yeast Saccharomyces cerevisiae. The general features of glycine transport in the yeast strain used (NCYC 239) are similar to those already reported for Saccharomyces carlsbergensis and S. cerevisiae, there being two kinetically distinct glycine uptake systems, with pH-independent K1/2 values near 14 and 0.4mM, respectively, but pH-dependent maximal velocities. Glycine transport itself has no measurable effect on pHi in aerobic cells, and only a marginal effect in energy-starved cells, but changes of pHi, imposed by extracellular addition of butyric acid, strongly influence glycine transport. Indeed, the dependence of glycine influx (in energy-starved cells) upon cytoplasmic H+ concentration appears to be third order, showing Hill slopes of 2.7-3.0. A crucial kinetic role for cytoplasmic pH in glycine transport is further indicated by a proportionality between the decline of flux and the decline of pHi produced by various metabolic inhibitors and uncouplers. Extracellular pH (pHo), by contrast, has only a weak effect on glycine influx, showing a Hill slope of 0.5. The major observations can be accommodated by a simple cyclic carrier scheme, in which 2 or more protons are transported along with glycine, but only one extracellular proton binding site dissociates in the testing range, with a pK near 5.5. The model requires a finite membrane potential, which must be somewhat sensitive to both pHi and pHo, and accommodates the discrepancy between measured net proton flux (one per glycine) and the kinetically required proton flux (two or more per glycine) by shunting through other proton-conducting pathways in the yeast membrane.  相似文献   

20.
Cells are more sensitive to heat when they are heated in an acidic environment, and this study confirms (K. G. Hofer and N. F. Mivechi, J. Natl. Cancer Inst., 65, 621, 1980) that intracellular pH (pHi) and not extracellular pH (pHe) is responsible for the sensitization. The relationship between pHe, pHi, and heat survival of cells heated in vitro in various buffers at pHe 6.3-8.0 was investigated. Cells' adaptation to low environmental pH in terms of increases in pHi and heat survival also was investigated. Finally, we studied the relationships among pHe, pHi, and survival from heat for cells heated in sodium-free reconstructed medium. Intracellular pH was measured by the distribution of the weak acid, [2-14C]5,5-dimethyl-2,4-oxazolidinedione. Our results are summarized as follows: (1) CHO cells maintained the same relationship between pHe and pHi in four different media or buffers (McCoy's 5a medium buffered with CO2 and NaHCO3 or 4-(2-hydroxyethyl)-1-piperazineethanesulfonic acid (Hepes) and 2-(N-morpholino)ethanesulfonic acid (Mes), Krebs-Ringer bicarbonate solution, and Krebs-Ringer phosphate solution) with pHi being 0.05 to 0.20 pH units higher than pHe as it varied from 7.0 to 6.4; furthermore, heat sensitization by acid was the same in medium buffered with NaHCO3 or Hepes and Mes. (2) The low pHe adapted cells multiplied with an increased doubling time of 20.7 +/- 0.7 h and appeared morphologically similar to the unadapted cells. However, the pHi of these cells was 0.15-0.30 pH units higher than that of the unadapted cells when pHe was varied between 7.0 and 6.3. (3) After being heated at 43.5 degrees C for 55 min or at 42.5 degrees C for 150 min at pHe 6.3-7.2, the pHi of the adapted cells increased by 0.2-0.1 pH units. However, heat caused no significant change in the unadapted cells. (4) Heat survival plotted versus pHe was 1000-fold higher for the adapted cells than for the unadapted cells at pHe of 6.3. However, heat survival plotted versus pHi was identical for the two cell types. (5) In sodium-free reconstructed McCoy's 5a medium, pHi was 0.25-0.1 pH units lower than that in the sodium-containing counterpart at pHe 6.3-7.2, and heat sensitization increased accordingly; however, heat survival plotted versus pHi was identical for the two types of media.  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