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 31 毫秒
1.
Induction of acid resistance (habituation) in Escherichia coli at pH 5.0 took ca 5 min in broth at 37 degrees C and 30-60 min in minimal medium. Induction occurred at a range of pH values from 4.0 to 6.0; it was dependent on continuing protein and RNA synthesis but substantial acid resistance appeared in the presence of nalidixic acid. Acid resistance was long-lasting; organisms grown at pH 5.0 retained most of their resistance after 2 h growth at pH 7.0. Organisms grown at pH 5.0 showed increased synthesis of a number of cytoplasmic proteins compared with the level in cells grown at pH 7.0. DNA repair-deficient strains carrying recA, uvrA or polA1 mutations were more acid-sensitive than the repair-proficient parents but were able to habituate at pH 5.0. Organisms grown at pH 5.0 transferred the ColV plasmid much more effectively at acid pH than did those grown at pH 7.0 and habituated recipients appeared better able to repair incoming acid-damaged plasmid DNA than did those that were non-habituated. Induction of acid resistance at pH 5.0 may be significant for the survival of organisms exposed to periodic discharges of acid effluent in the aquatic environment and habituation may also allow plasmid transfer and repair of acid-damaged plasmid DNA during or after such exposure.  相似文献   

2.
Adaptive acid tolerance response of Streptococcus sobrinus   总被引:1,自引:0,他引:1  
Streptococcus mutans and Streptococcus sobrinus are the bacteria most commonly associated with human dental caries. A major virulence attribute of these and other cariogenic bacteria is acid tolerance. The acid tolerance mechanisms of S. mutans have begun to be investigated in detail, including the adaptive acid tolerance response (ATR), but this is not the case for S. sobrinus. An analysis of the ATR of two S. sobrinus strains was conducted with cells grown to steady state in continuous chemostat cultures. Compared with cells grown at neutral pH, S. sobrinus cells grown at pH 5.0 showed an increased resistance to acid killing and were able to drive down the pH through glycolysis to lower values. Unlike what is found for S. mutans, the enhanced acid tolerance and glycolytic capacities of acid-adapted S. sobrinus were not due to increased F-ATPase activities. Interestingly though, S. sobrinus cells grown at pH 5.0 had twofold more glucose phosphoenolpyruvate:sugar phosphotransferase system (PTS) activity than cells grown at pH 7.0. In contrast, glucose PTS activity was actually higher in S. mutans grown at pH 7.0 than in cells grown at pH 5.0. Silver staining of two-dimensional gels of whole-cell lysates of S. sobrinus 6715 revealed that at least 9 proteins were up-regulated and 22 proteins were down-regulated in pH 5.0-grown cells compared with cells grown at pH 7.0. Our results demonstrate that S. sobrinus is capable of mounting an ATR but that there are critical differences between the mechanisms of acid adaptation used by S. sobrinus and S. mutans.  相似文献   

3.
Growth of Escherichia coli 1829 ColV, I-K94 at pH 5.0 led to an increase in u.v. resistance compared with cells grown at pH 7.0. This was due to a phenotypic change, since organisms grown at pH 7.0 showed increased resistance after only 2.5–5.0 min incubation at the mildly acid pH. Other E. coli K12 derivatives became more u.v.-resistant at pH 5.0 including uvrA, recA and polA1 mutants. Organisms grown at pH 5.0 also showed increased Weigle reactivation of u.v.-irradiated Λ phage and this applied to the repair-deficient mutants as well as the parent strains. Both the increased u.v. resistance of acid-habituated cells and their increased ability to bring about Weigle reactivation appear to involve RecA-independent processes and are presumably, therefore, independent of the SOS response.  相似文献   

4.
Growth of Escherichia coli 1829 ColV, I-K94 at pH 5.0 led to an increase in u.v. resistance compared with cells grown at pH 7.0. This was due to a phenotypic change, since organisms grown at pH 7.0 showed increased resistance after only 2.5-5.0 min incubation at the mildly acid pH. Other E. coli K12 derivatives became more u.v.-resistant at pH 5.0 including uvrA, recA and polA1 mutants. Organisms grown at pH 5.0 also showed increased Weigle reactivation of u.v.-irradiated lambda phage and this applied to the repair-deficient mutants as well as the parent strains. Both the increased u.v. resistance of acid-habituated cells and their increased ability to bring about Weigle reactivation appear to involve RecA-independent processes and are presumably, therefore, independent of the SOS response.  相似文献   

5.
Exposure of Saccharomyces cerevisiae to sorbic acid strongly induces two plasma membrane proteins, one of which is identified in this study as the ATP-binding cassette (ABC) transporter Pdr12. In the absence of weak acid stress, yeast cells grown at pH 7.0 express extremely low Pdr12 levels. However, sorbate treatment causes a dramatic induction of Pdr12 in the plasma membrane. Pdr12 is essential for the adaptation of yeast to growth under weak acid stress, since Deltapdr12 mutants are hypersensitive at low pH to the food preservatives sorbic, benzoic and propionic acids, as well as high acetate levels. Moreover, active benzoate efflux is severely impaired in Deltapdr12 cells. Hence, Pdr12 confers weak acid resistance by mediating energy-dependent extrusion of water-soluble carboxylate anions. The normal physiological function of Pdr12 is perhaps to protect against the potential toxicity of weak organic acids secreted by competitor organisms, acids that will accumulate to inhibitory levels in cells at low pH. This is the first demonstration that regulated expression of a eukaryotic ABC transporter mediates weak organic acid resistance development, the cause of widespread food spoilage by yeasts. The data also have important biotechnological implications, as they suggest that the inhibition of this transporter could be a strategy for preventing food spoilage.  相似文献   

6.
红壤中镉在有机酸作用下的解吸行为   总被引:1,自引:1,他引:0  
采用平衡批处理法,研究了3种有机酸及其两两混合液在序列pH值梯度下(pH 3.0~7.0)对华南山地红壤Cd解吸行为的影响.结果表明,草酸与苹果酸不利于Cd的解吸,反而促进了吸附,其中草酸只是在较高浓度(20 mmol·L-1)且土壤溶液pH>5.0时促进解吸.随着pH值升高,草s酸、苹果酸以及不含有机酸的对照溶液对红壤中Cd的解吸率都快速下降.柠檬酸在pH<5.0时不利于Cd解吸;在pH>5.0时显著促进Cd解吸,但两种浓度柠檬酸解吸特征有所不同,在低浓度(2 mmol·L-1)下对镉的解吸率呈降低-升高-降低变化,在高浓度(20 mmol·L-1)下呈降低-升高变化.在低pH条件下(pH 3.0、4.0),苹果酸最有利于Cd的解吸,但3种酸对Cd解吸率差别不大,在较高pH条件下(pH 5.0~7.0),柠檬酸最有利于解吸,且解吸率大大高于草酸与苹果酸.有机酸混合没有明显的交互作用,对Cd的解吸率介于相应单独有机酸之间.  相似文献   

7.
The efficacy of different organic acids in decreasing the heat resistance of Paenibacillus polymyxa spores was assessed. The relationship between concentration of the undissociated form of different organic acids and decrease in heat resistance was also investigated. The heat resistance of P. polymyxa spores was tested in distilled water at 85, 90 and 95 degrees C, at pH4 and in the presence of 50, 100 and 200 mmol l(-1) of the undissociated form of lactic, citric or acetic acid and sodium citrate or acetate. The undissociated form of organic acids was responsible for increasing the heat sensitivity of spores. The most effective acid was lactic acid. The D values of the spores decreased rapidly (between 74 and 43%) in the presence of 50 mmol l(-1) of the undissociated form of organic acid, and increasing concentrations of these forms affected the heat resistance of spores less than proportionally. The heat resistance of the spores in milk was approximately threefold lower than in distilled water. This work has shown that the undissociated fraction of organic acids increases, albeit non-linearly, the sensitivity of spores to heat, even in complex substrates such as milk. By knowing the amount of organic acids added to a given substrate, their dissociation constants and the final pH, it could be possible to estimate the concentration of undissociated forms and the corresponding increase in lethality of heat treatments. This would help the food industry to maximize the lethality achieved by heat processes and/or safely reduce the heat treatments already in use.  相似文献   

8.
Several acid tolerance responses of Escherichia coli were associated with secretion into the growth media of components (frequently proteins) which altered acid tolerance of other cultures. First, medium filtrates from cultures induced to acid tolerance by several conditions converted pH 7.0-grown organisms to tolerance and, for most such responses, filtrate proteins were needed for full induction. Secondly, filtrates from cultures induced to acid sensitivity at alkaline pH produced sensitisation of resistant cultures. Thirdly, filtrates from inherently tolerant or sensitive strains altered tolerance or sensitivity of normal strains. In many cases, filtrate components were essential for the original response e.g. acid habituation at pH 5.O. Extracellular components may function as intermediates only in stress tolerance responses, but other adaptive responses must be tested as such components may function in other inducible processes.  相似文献   

9.
The effect of fatty acids on Mycobacterium smegmatis was examined in vitro at pH 5.0 to 7.0 to determine the role of fatty acids in the intracellular killing of mycobacteria. Unsaturated fatty acids showed strong bactericidal activity in low concentrations (0.005 to 0.02 mM), whereas saturated fatty acids, except for lauric and myristic acids, were not very effective even at a concentration of 0.2 mM. Addition of a saturated fatty acid (palmitic or stearic acid) to an unsaturated fatty acid (oleic or linoleic acid) did not strongly interfere with the bactericidal effect of the unsaturated fatty acid at pH 5.0 and 6.0. Ca2+ (3.0 mM), Mg2+ (1.0 mM), and gamma-globulin (0.4%) showed weak reversal effects on the bactericidal activity of unsaturated fatty acids at pH 5.0 and 6.0. Serum albumin and serum showed strong reversal effects. The concentrations of each fatty acid in a mixture (molar ratio, 1:1:1:1) of oleic, linoleic, palmitic, and stearic acids required for the killing of M. smegmatis in the presence of 2% serum (bovine, rabbit, or human) were 0.05 to 0.10 mM at pH 5.0 and 6.0 and 0.05 to 0.20 mM at pH 7.0, depending on the serum used. The susceptibilities of M. kansasii, M. bovis strain BCG, and M. tuberculosis to the mixture of the four fatty acids in the presence of 2% bovine serum were similar to that of M. smegmatis, although M. fortuitum was more resistant.  相似文献   

10.

Background

The TolC outer membrane channel is a key component of several multidrug resistance (MDR) efflux pumps driven by H+ transport in Escherichia coli. While tolC expression is under the regulation of the EvgA-Gad acid resistance regulon, the role of TolC in growth at low pH and extreme-acid survival is unknown.

Methods and Principal Findings

TolC was required for extreme-acid survival (pH 2) of strain W3110 grown aerobically to stationary phase. A tolC deletion decreased extreme-acid survival (acid resistance) of aerated pH 7.0-grown cells by 105-fold and of pH 5.5-grown cells by 10-fold. The requirement was specific for acid resistance since a tolC defect had no effect on aerobic survival in extreme base (pH 10). TolC was required for expression of glutamate decarboxylase (GadA, GadB), a key component of glutamate-dependent acid resistance (Gad). TolC was also required for maximal exponential growth of E. coli K-12 W3110, in LBK medium buffered at pH 4.5–6.0, but not at pH 6.5–8.5. The TolC growth requirement in moderate acid was independent of Gad. TolC-associated pump components EmrB and MdtB contributed to survival in extreme acid (pH 2), but were not required for growth at pH 5. A mutant lacking the known TolC-associated efflux pumps (acrB, acrD, emrB, emrY, macB, mdtC, mdtF, acrEF) showed no growth defect at acidic pH and a relatively small decrease in extreme-acid survival when pre-grown at pH 5.5.

Conclusions

TolC and proton-driven MDR efflux pump components EmrB and MdtB contribute to E. coli survival in extreme acid and TolC is required for maximal growth rates below pH 6.5. The TolC enhancement of extreme-acid survival includes Gad induction, but TolC-dependent growth rates below pH 6.5 do not involve Gad. That MDR resistance can enhance growth and survival in acid is an important consideration for enteric organisms passing through the acidic stomach.  相似文献   

11.
Abstract Streptococcus mutans GS-5 synthesizes an exo-β-d-fructosidase, FruA, capable of degrading levans, inulins, sucrose and raffinose, with the greatest activity on levans. A previous analysis of the deduced amino acid sequence of the FruA protein revealed the presence of a C-terminus with an LPXTGX membrane sorting sequence and membrane spanning domain, characteristic of many Gram-positive cocci surface proteins. Here it is demonstrated that FruA, which had been previously shown to exist almost exclusively as an extracellular enzyme, can be detected in significant proportions at the surface of S. mutans cells. Moreover, growth of S. mutans GS-5 at steady state in continuous culture at pH values of 7.0, 6.0, or 5.0 revealed that the amount of cell-associated enzyme increased with decreasing pH values, such that roughly 50% of the total fructanase activity of pH 5.0-grown organisms was cell-associated. This result was confirmed using anti-recombinant-FruA antisera in Western blotting of culture supernate and cell-associated enzyme preparations from chemostat-grown cells. Incubation of S. mutans at pH values of 5.0, 6.0 or 7.0 in buffered media yielded results similar to those observed in the chemostat experiments. The release of FruA from S. mutans was also shown to be inhibitable by copper, which is known to interfere with the release of the surface adhesin, P1, from intact cells and protoplasts of S. mutans . These data provide evidence for a unique post-translational mechanism for the regulation of the catabolism of polysaccharides by bacteria. The control of degradation of plaque fructans by modulation of the release of the fructanase enzyme from S. mutans may play a critical role in the temporal and spatial separation of the synthesis and degradation of dental plaque fructans.  相似文献   

12.
The antibacterial activity of p -aminobenzoic acid against Listeria monocytogenes, Salmonella enteritidis and Escherichia coli was compared with the activity of commonly used acidulants: formic, propionic, acetic, lactic and citric acids. Viable count evaluations and MIC determinations indicated that p -aminobenzoic acid caused greater inhibitory effects than the other organic acids. The activity of p -aminobenzoic acid on the growth of the test organisms at selected pH values indicated that p -aminobenzoic acid was more active at low pH than at high pH. Uptake studies showed that the uptake of p -aminobenzoic acid by E. coli was markedly decreased as the pH values increased. Electron micrographs of E. coli cells grown in the presence of p -aminobenzoic acid indicate that p -aminobenzoic acid caused marked damage to the cell envelope. It is suggested that p -aminobenzoic acid has at least two mechanisms of action: one mechanism in common with other organic acids and the other mechanism by interfering with the synthesis of the peptidoglycan layer by an action on the dihydrofolate reductase enzyme.  相似文献   

13.
Summary A bench scale continuously stirred reactor was used to study the acidogenic phase of the anaerobic fermentation of stillage. The residence time of the effluent in the reactor ranged from 15.7 to 8.2 hours, pH was around 5.0 and temperature was maintained at 35°C. The results indicate that the residence time had no appreciable effect on the production or composition of the organic acids. The main acid products found in the reactor effluent were acetic, propionic and butiric acids.  相似文献   

14.
The membrane lipids from two obligately and two facultatively alkalophilic strains of Bacillus spp. were characterized in a comparative study that included B. subtilis. Preparations of membrane lipids were made from pH 10.5-grown cells of all of the alkalophiles and from pH 7.5- or 7.0-grown cells of the two facultative strains and B. subtilis. The two obligate alkalophiles contained high ratios of membrane lipid to membrane protein, and the lipid fraction contained a high proportion of neutral lipid. These characteristics are probably not prerequisites for growth at very high pH since one or another of the facultative strains failed to show these properties at high pH. All of the alkalophiles contained appreciable amounts of squalene and C40 isoprenoids. Among the polar lipids, the alkalophiles all contained high concentrations of anionic phospholipids, including phosphatidylglycerol and especially large amounts of cardiolipin; phosphatidylethanolamine was the other major phospholipid. Small amounts of bis(monoacylglycero)phosphate were found in most, but not all, of the alkalophile preparations. Glycolipids and phosphoglycolipids were absent. The fatty acid composition of the total phospholipid and individual fractions revealed two features that distinguished between the obligate and facultative strains. Membranes from the obligately alkalophilic species contained a high concentration of branched-chain fatty acids, comparable to that in membranes from B. subtilis, as well as a relatively high content of unsaturated fatty acids. By contrast, the facultatively alkalophilic strains contained almost no unsaturated fatty acids and a lower concentration of branched-chain fatty acids than either the obligate alkalophiles or B. subtilis.  相似文献   

15.
16.
Escherichia coli K12 transferred from pH 7.0 to pH 9.0 gains alkylhydroperoxide (AHP) tolerance. The aim here was to establish whether extracellular components (ECs) are needed for such induction. Therefore, the effects of removing ECs during incubation at pH 9.0 were tested and the abilities of culture filtrates to induce tolerance were examined. First, AHP tolerance did not appear, at pH 9.0, if cultures were subjected to continuous filtration or dialysis, against the same medium, suggesting that an EC might be needed. Second, neutralized filtrates from pH 9.0-grown cultures induced tolerance at pH 7.0, and these filtrates were inactivated by dialysis, filtration or heating but not by protease. Thus, pH 9.0 filtrates have a small non-protein extracellular induction component (EIC), which acts as an alarmone, 'warning' cells of stress and preparing them to resist it. Filtrates from pH 7.0-grown cultures did not induce AHP tolerance at pH 7.0 but if incubated at pH 9.0 without organisms, gained such ability. It is proposed that pH 7.0 filtrates have an EIC precursor (termed an extracellular sensing component, ESC), which senses alkaline pH, and is converted by it to the EIC. The ESC in pH 6.0 filtrates was distinct from that in pH 7.0 filtrates; there may be several oligomeric (or conformational) forms of this ESC. As the EIC is small, it can diffuse away from the alkalinized region and induce tolerance in unstressed organisms.  相似文献   

17.
Escherichia coli has an ability, rare among the Enterobacteriaceae, to survive extreme acid stress under various host (e.g., human stomach) and nonhost (e.g., apple cider) conditions. Previous microarray studies have exposed a cluster of 12 genes at 79 centisomes collectively called an acid fitness island (AFI). Four AFI genes, gadA, gadX, gadW, and gadE, were already known to be involved in an acid resistance system that consumes an intracellular proton through the decarboxylation of glutamic acid. However, roles for the other eight AFI gene products were either unknown or subject to conflicting findings. Two new aspects of acid resistance are described that require participation of five of the remaining eight AFI genes. YhiF (a putative regulatory protein), lipoprotein Slp, and the periplasmic chaperone HdeA protected E. coli from organic acid metabolites produced during fermentation once the external pH was reduced to pH 2.5. HdeA appears to handle protein damage caused when protonated organic acids diffuse into the cell and dissociate, thereby decreasing internal pH. In contrast, YhiF- and Slp-dependent systems appear to counter the effects of the organic acids themselves, specifically succinate, lactate, and formate, but not acetate. A second phenomenon was defined by two other AFI genes, yhiD and hdeD, encoding putative membrane proteins. These proteins participate in an acid resistance mechanism exhibited only at high cell densities (>10(8) CFU per ml). Density-dependent acid resistance does not require any demonstrable secreted factor and may involve cell contact-dependent activation. These findings further define the complex physiology of E. coli acid resistance.  相似文献   

18.
In this study, we determined the internal cellular pH response of Leuconostoc mesenteroides and Lactobacillus plantarum to the external pH created by the microorganisms themselves or by lactic or acetic acids and their salts added to the growth medium. Growth of Leuconostoc mesenteroides stopped when its internal pH reached 5.4 to 5.7, and growth of L. plantarum stopped when its internal pH reached 4.6 to 4.8. Variation in growth medium composition or pH did not alter the growth-limiting internal pH reached by these microorganisms. L. plantarum maintained its pH gradient in the presence of either 160 mM sodium acetate or sodium lactate down to an external pH of 3.0 with either acid. In contrast, the DeltapH of Leuconostoc mesenteroides was zero at pH 4.0 with acetate and 5.0 with lactate. No differences were found between d-(-)- and l-(+)-lactic acid for the limiting internal pH for growth of either microorganism. The comparatively low growth-limiting internal pH and ability to maintain a pH gradient at high organic acid concentration may contribute to the ability of L. plantarum to terminate vegetable fermentations.  相似文献   

19.
Summary Clostridum propionicum is a chemical autotroph that metabolizes alanine to propionic acid (reduction product) and acetic acid (oxidation product). The ratio of propionate/acetate predicted by the electron balance is 2:1. This study reports the effect of pH on growth and organic acid production by this organism when grown in both test tube cultures initially buffered from pH 7.0 to 5.0, and in fermentors maintained at pH 7.0 and 6.5. Highest growth and organic acid production was found at pH 7.0 in both cases. HPLC analysis showed that at pH 7.0, the ratios of propionate to acetate were 0.45:1 (stationary tube, 24 h). The highest ratio observed was 1.8:1 (stationary tube, pH 6.0, 24h). This tube produced 8.5% of the acids produced in the pH 7.0 culture tube. The identify of the major portion of the reduction products of the organism remains unknown.  相似文献   

20.
The classic biochemical pH-stat model of cytosolic pH regulation in plant cells presupposes a pH-dependent biosynthesis and degradation of organic acids, specifically malic acid, in the cytosol. This model has been used to explain the higher tissue accumulation of organic acids in nitrate (NO3)-grown, relative to ammonium (NH4+)-grown, plants, the result of proposed cytosolic alkalinization by NO3 metabolism, and acidification by NH4+ metabolism. Here, a critical examination of the model shows that its key assumptions are fundamentally problematic, particularly in the context of the effects on cellular pH of nitrogen source differences. Specifically, the model fails to account for proton transport accompanying inorganic nitrogen transport, which, if considered, renders the H+ production of combined transport and assimilation (although not the accumulation) to be equal for NO3 and NH4+ as externally provided N sources. We show that the model's evidentiary basis in total-tissue mineral ion and organic acid analysis is not directly relevant to subcellular (cytosolic) pH homeostasis, while the analysis of the ionic components of the cytosol is relevant to this process. A literature analysis further shows that the assumed greater activity of the enzyme phosphoenolpyruvate (PEP) carboxylase under nitrate nutrition, which is a key characteristic of the biochemical pH-stat model as it applies to nitrogen source, is not borne out in numerous instances. We conclude that this model is not tenable in its current state, and propose an alternative model that reaffirms the anaplerotic role of PEP carboxylase within the context of N nutrition, in the production of carbon skeletons for amino acid synthesis.  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