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 distribution of salicylic acid between the intracellular and extracellular phases has been used to estimate the intracellular pH in the Ehrlich cell and Escherichia coli. The validity of the method was established by: (i) comparison of the results obtained with salicylic acid with those obtained with 5,5-dimethyloxazolidine-2,4-dione; (ii) by following changes of the apparent intracellular pH under circumstances in which such changes are predictable, e.g., the addition of weak acids or proton conductors to the incubation medium during incubation at acidic pH; (iii) by comparison of the apparent intracellular pH changes with the uptake of H+ by the cells estimated from the changes of the medium pH. Optimal results are obtained with this indicator when the extracellular pH is below 5.5, because in this case the indicator is to a sufficient extent in its penetrating form, so that its movement can reflect intracellular pH changes occurring in less than 30 s. When the intracellular pH falls below 5.2 measurable binding of salicylic acid to the intracellular material of the Ehrlich cell takes place, but above this pH no binding has been found.The Ehrlich cell and cells of Escherichia coli behaved similarly under various experimental circumstances tested, but striking differences were found in the inherent permeability of the membrane to H+ and in the changes in this parameter by lowering the temperature to 2°C.  相似文献   

2.
Influence of two types of freezing, at -196 and -50 degrees C with following thawing of Escherichia coli cells at 37 degrees C on the value of intracellular pH has been studied by means of 31P NMR spectroscopy. All the cycles of freeze-thawing have been shown to result in acidification of intracellular medium on 1.0 unit pH apart from the freezing type. The extracellular medium (pHex) was acidified too, but the degree of pHex changes after freeze-thawing depended on the freezing depth.  相似文献   

3.
Escherichia coli became more acid tolerant following incubation for 60 min in a medium containing L-glutamate at pH 7.0, 7.5 or 8.5. Several agents, including cAMP, NaCl, sucrose, SDS and DOC, prevented tolerance appearing if present with L-glutamate. Lesions in cysB, hns, fur, himA and relA, which frequently affect pH responses, failed to prevent L-glutamate-induced acid tolerance but a lesion in L-glutamate decarboxylase abolished the response. Induction of acid tolerance by L-glutamate was associated with the accumulation in the growth medium of a protein (or proteins) which was able to convert pH 7.0-grown cultures to acid tolerance, and the original L-glutamate-induced tolerance response was dependent on this component(s). Acid tolerance was also induced by L-aspartate at pH 7.0 and induction of such tolerance was dependent on an extracellular protein (or proteins). The L-glutamate and L-aspartate acid tolerance induction processes are further examples of a number of stress tolerance responses which differ from most inductions in that extracellular components, including extracellular sensors, are required.  相似文献   

4.
During prolonged incubation in stationary phase Escherichia coli undergoes starvation-induced differentiation, resulting in highly resistant cells. In rich medium with high amino acid content further incubation of cultures at high cell density leads to the generation of a population of cells no longer able to form colonies. The viability loss is due to some component of spent medium, active at high pH and high cell density, and can be prevented either by keeping the pH close to neutrality, by washing off the nonsalt components of the medium, or by keeping the saturating cell density low. Exposure to short-chain n-alcohols within a specific time window in stationary phase also prevents viability loss, in an rpoS-dependent fashion. The development of stress resistance, a hallmark of stationary-phase cells, is affected following alcohol treatment, as is the response to extracellular factors in spent medium. Alcohols seem to block cells in an early phase of starvation-induced differentiation, most likely by interfering with processes important for regulation of sigma(s) such as cell density signals and sensing the nutrient content of the medium.  相似文献   

5.
Studies with Ehrlich ascites tumor cells showed that small decreases in the pH of the incubation medium from 7.4 increase the magnitude of incorporation of free fatty acid (FFA) into the cells from an albumin solution. A similar effect occurred when rabbit erythrocytes, rat heart slices, or rat liver slices were incubated with FFA-bovine albumin solutions and when tumor cells were incubated with FFA in media containing human albumin, -lactoglobulin, or rat plasma. The effect was not seen when the medium contained no protein. When the pH of the albumin-containing medium was lowered from 7.4 to 6.6, oxidation of FFA to CO(2) by the tumor cells increased, esterification of the FFA (mostly into phospholipids and triglycerides) increased, and less esterified radio-active fatty acid was depleted from the cells. Hence, more fatty acid accumulated in the cells in more acid media. These findings suggest that small changes in extracellular pH might regulate FFA utilization and lipid accumulation in mammalian tissues.  相似文献   

6.
An extracellular induction component (EIC), needed for acid tolerance induction at pH 5.0 in Escherichia coli, arises from an extracellular precursor which senses acid stress and is activated (forming the EIC) by such stress. The precursor, which is a heat-stable protein, was formed by cells which had not been subjected to acid stress, being present in culture media after growth at pH values from 7.0 to 9.0. This stress-sensing molecule was activated to the EIC at pH values from 4.5 to 6.0 but not at pH 6.5 and did not form EIC on incubation at an extremely acidic pH e.g. 2.0. The precursor was not inactivated at pH 2.0. Precursor activation might be reversible, as the EIC lost its ability to induce acid tolerance after incubation at pH 9.0, but regained it if subsequently incubated at pH 5.0. Whereas the sensor formed at pH 7.0 can only be activated at pH 5.0 to 6.0, that synthesized at pH 9.0 can be activated at pH 5.0 to 7.5. Accordingly, this work shows that the acid stress sensor is extracellular, and it is proposed that its presence in the medium rather than in the cells, allows more sensitive and rapid responses to acid stress.  相似文献   

7.
2,3-Butanediol (2,3-BDO) is an organic compound with a wide range of industrial applications. Although Escherichia coli is often used for the production of organic compounds, the wild-type E. coli does not contain two essential genes in the 2,3-BDO biosynthesis pathway, and cannot ferment 2,3-BDO. Therefore, a 2,3-BDO biosynthesis mutant strain of Escherichia coli was constructed and cultured. To determine the optimum culture factors for 2,3-BDO production, experiments were conducted under different culture environments ranging from strongly acidic to neutral pH. The extracellular metabolite profiles were obtained using high-performance liquid chromatography (HPLC), and the intracellular metabolite profiles were analyzed by ultra-performance liquid chromatography and quadruple time-of-flight mass spectrometry (UPLC/ Q-TOF-MS). Metabolic flux analysis (MFA) was used to integrate these profiles. The metabolite profiles showed that 2,3-BDO production favors an acidic environment (pH 5), whereas cell mass favors a neutral environment. Furthermore, when the pH of the culture fell below 5, both the cell growth and 2,3-BDO production were inhibited.  相似文献   

8.
The effect of variable extracellular pH on intracellular pH, cell energy status, and thermal sensitivity was evaluated in CHO cells over the extracellular pH range of 6.0 to 8.6. Extracellular pH was adjusted with either lactic acid, HCl, or NaOH. Regardless of the method of pH adjustment, the results obtained were similar. The relationship between extracellular and intracellular pH was dependent upon the pH range examined. Intracellular pH was relatively resistant to a change in extracellular pH over the pHe range of 6.8 to 7.8 (i.e., delta pHi congruent to delta pHe X 0.33). Above and below this range, delta pHi congruent to delta pHe X 1.08 or X 0.76, respectively. Cellular survival after a 30-min heat treatment at 44 degrees C remained constant over the extracellular pH range of 7.0 to 8.4, but varied substantially over a similar intracellular pH range. The cellular concentration of the high energy phosphate reservoir, phosphocreatine, decreased with decreasing pH. However, the cellular concentrations of ATP, ADP, and AMP remained constant over the entire pH range examined. It is concluded that increased thermal sensitivity resulting from a change in extracellular pH is not due to cellular energy depletion. Furthermore, intracellular pH is a more accurate indicator of thermal sensitivity than is extracellular pH.  相似文献   

9.
The effects of extracellular ATP and/or the phorbol ester 12-O-tetradecanoylphorbol-13-acetate (TPA) on the intracellular pH of Ehrlich ascites tumor cells were measured using both distribution of [14C]5,5-dimethyloxazolidine-2,4-dione, and the fluorescent indicator 5(6)-carboxyfluorescein. Micromolar concentrations of extracellular ATP induce a biphasic change in the intracellular pH characterized by a rapid acidification of 0.04 pH units followed by an alkalinization of 0.11 pH units. Concurrently with the alkalinization, an increase in the total cellular [Na+] from 37.5 to 45.0 mM is observed. The pH change is half-maximally activated by 0.5-2.5 microM extracellular ATP. The intracellular alkalinization, but not the initial acidification, phase requires extracellular Na+, with half-maximal alkalinization in the presence of 24-32 mM Na+, and is inhibited by amiloride. Exposure of Ehrlich ascites tumor cells to TPA alone produces a slight alkalinization of approximately 0.04 pH units. Conversely, preincubation of the cells with TPA partially inhibits the ATP-induced changes in intracellular pH. Under identical conditions TPA also inhibits the ATP-induced increase in the cytosolic [Ca2+]. The half-maximal dose for both effects is produced by 3-10 nM TPA. These data indicate that extracellular ATP triggers the activation of Na+/H+ exchange. Furthermore, activation of protein kinase C mediates at least part of the Na+/H+ exchange, although a second mechanism may also exist.  相似文献   

10.
Woodside, E. E. (U.S. Army Medical Research Laboratory, Fort Knox, Ky.), and W. Kocholaty. Carbohydrate and lipid content of radiation-resistant and -sensitive strains of Escherichia coli. J. Bacteriol. 87:1140-1146. 1964.-Total lipid contents of acetate minimal medium cultures of Escherichia coli, strains B, B/r, and B(s), were not significantly different when identical pretreatment and extraction procedures were compared. Wide variations in intracellular hexose and pentose derivatives of E. coli B, B/r, and B(s) were induced by changes in carbon and nitrogen sources and by changes in the growth phases. The three strains produced more intracellular carbohydrate when grown in nutrient broth-glucose medium than when grown in unsupplemented nutrient broth. Acetate minimal medium cultures of the radiation-sensitive mutant, E. coli B(s), contained the least, and the radiation-resistant mutant, E. coli B/r the largest, amounts of intracellular hexoses. Environmental conditions which increased the radiation resistance of E. coli B/r were similar to the environmental conditions which favored increased intracellular hexose accumulation. After X ray of E. coli B/r, considerable amounts of hexoses and pentoses were released into the growth medium. Alterations in hexose distribution patterns of X-rayed E. coli B/r preceded alterations in pentose distribution patterns. Prolonged postirradiation incubation resulted in a net synthesis of extracellular hexose, with concomitant loss of intracellular hexose accumulation.  相似文献   

11.
Changes in the membrane potential, pH gradient, proton motive force, and intracellular pH of Escherichia coli were followed during the chemotactic responses to a variety of potentially membrane-active compounds. Lipophilic weak acids, decreases in extracellular pH, and nigericin each caused a repellent response. Lipophilic weak bases, increases in extracellular pH, and valinomycin in the presence of K+ each caused an attractant response. Changes in membrane potential, pH gradient, and proton motive force did not correlate with the behavioral responses to these treatments, but changes in intracellular pH did correlate. Furthermore, the strength of the response to a weak acid was correlated with the magnitude of the change of the intracellular pH, and many compounds which could alter the intracellular pH were found to be chemotactically active. Apparently these attractants and repellents are not detected by specific chemoreceptors but rather are detected via the ability of cells to sense and respond to changes in intracellular pH. The pathway of sensory transduction which proceeds through methyl-accepting chemotaxis protein I was found to be involved in the response to a change in intracellular pH.  相似文献   

12.
The ability of glucose to reverse the effects of dinitrophenol on amino acid uptake in Ehrlich cells is a function of pH. At pH 6.0, the presence of glucose does not reverse the inhibitory action of the uncoupler. Nearly complete restoration occurs with glucose at pH 7.4. At pH 8, the presence of glucose may cause a modest increase in amino acid uptake in presence of dinitrophenol. At all pH values, glucose restores ATP and cellular K+ to the control levels at the same pH. Although the cytoplasmic pH changes with changes in the external pH, the cell interior is more alkaline than the medium near pH 6.0 and more acid than the medium at pH 7.8 even after 45 min incubation at 37°C. With dinitrophenol and in presence of glucose the difference in pH between the medium and the cell is minimal at both pH 6.0 and 7.8.  相似文献   

13.
The ability of glucose to reverse the effects of dinitrophenol on amino acid uptake in Ehrlich cells is a function of pH. At pH 6.0, the presence of glucose does not reverse the inhibitory action of the uncoupler. Nearly complete restoration occurs with glucose at pH 7.4. At pH 8, the presence of glucose may cause a modest increase in amino acid uptake in presence of dinitrophenol. At all pH values, glucose restores ATP and cellular K+ to the control levels at the same pH. Although the cytoplasmic pH changes with changes in the external pH, the cell interior is more alkaline than the medium near pH 6.0 and more acid than the medium at pH 7.8 even after 45 min incubation at 37 degrees C. With dinitrophenol and in presence of glucose the difference in pH between the medium and the cell is minimal at both pH 6.0 and 7.8.  相似文献   

14.
The endo-beta-1,3-1,4-glucanase enzyme of Bacillus subtilis C120, when synthesized in Escherichia coli, is located mainly in the cytoplasm, but enzyme activity is also detected in the periplasmic space and in the extracellular medium. The proportion recovered in the extracellular medium is not altered by changes in the levels of synthesis of the enzyme. Lysis of E. coli cells is ruled out as the cause of the secretion by the normal localization of beta-galactosidase, an intracellular protein. However, beta-lactamase, which is normally found in the periplasmic space, is detected in the extracellular medium of E. coli transformants containing beta-glucanase plasmids, suggesting that the presence of beta-glucanase in the cell alters the permeability of the outer membrane. The beta-glucanase proteins found in the extracellular medium, the periplasmic space and the cytoplasm have the same electrophoretic mobilities as the secreted enzyme of B. subtilis. Amino-terminal sequencing has shown that the beta-glucanase enzyme in the intracellular fraction of E. coli is processed at a site two amino acids distant from the processing site used in B. subtilis.  相似文献   

15.
The Rhodopseudomonas palustris KUGB306 hemA gene codes for 5-aminolevulinic acid (ALA) synthase. This enzyme catalyzes the condensation of glycine and succinyl-CoA to yield ALA in the presence of the cofactor pyridoxal 5'- phosphate. The R. palustris KUGB306 hemA gene in the pGEX-KG vector system was transformed into Escherichia coli BL21. The effects of physiological factors on the extracellular production of ALA by the recombinant E. coli were studied. Terrific Broth (TB) medium resulted in significantly higher cell growth and ALA production than did Luria-Bertani (LB) medium. ALA production was significantly enhanced by the addition of succinate together with glycine in the medium. Maximal ALA production (2.5 g/l) was observed upon the addition of D-glucose as an ALA dehydratase inhibitor in the late-log culture phase. Based on the results obtained from the shake-flask cultures, fermentation was carried out using the recombinant E. coli in TB medium, with the initial addition of 90 mM glycine and 120 mM succinate, and the addition of 45 mM D-glucose in the late-log phase. The extracellular production of ALA was also influenced by the pH of the culture broth. We maintained a pH of 6.5 in the fermenter throughout the culture process, achieving the maximal levels of extracellular ALA production (5.15 g/l, 39.3 mM).  相似文献   

16.
The relationship between the steady-state sodium gradient (delta pNa) and the protonmotive force developed by endogenously respiring Escherichia coli cells has been studied quantitatively, using 23Na NMR for measurement of intracellular and extracellular sodium concentrations, 31P NMR for measurement of intracellular and extracellular pH, and tetraphenylphosphonium distribution for measurement of membrane potential. At constant protonmotive force, the sodium concentration gradient was independent of extracellular concentrations over the measured range of 4-285 mM, indicating that intracellular sodium concentration is not regulated. The magnitude of delta pNa was measured as a function of the composition and magnitude of the protonmotive force. At external pH values below 7.2, delta pNa was parallel to delta pH but showed no simple relationship to the membrane potential; above pH 7.2 the parallel relationship began to diverge, with delta pH continuing to decrease but delta pNa starting to level off or increase. Although plots of delta pNa versus delta pH had slopes of close to 1, the value of delta pNa consistently exceeded that of delta pH by approximately 0.4 units, indicating a partially electrogenic character to the putative H+/Na+ antiport. The apparent stoichiometry was 1.13 +/- 0.01 at external pH below 7.2. The possible significance of this nonintegral stoichiometry is discussed according to a model in which two distinct integral stoichiometries (possibly 1H+/1Na+ and 2H+/1Na+) are available with some relative probability; the model predicts futile cycling of sodium ions and a dissipative proton current. In the course of this study, we discovered that the magnitude of the pH gradient developed by the cells was osmolarity-dependent, yielding steady-state intracellular pH values that varied from 7.1 at 100 mosm to 7.7 at 800 mosm.  相似文献   

17.
G R Finch  M E Stiles    D W Smith 《Applied microbiology》1987,53(12):2894-2896
Selective and nonselective growth media were evaluated at two incubation temperatures, 35 and 44.5 degrees C, for the recovery of a nalidixic acid-resistant marker strain of Escherichia coli ATCC 11775 by membrane filtration from ozonated 0.05 M phosphate buffer (pH 6.9). There were significantly fewer bacteria recovered with the standard m-FC agar when compared with the same growth medium prepared without bile salts and rosolic acid. This effect was particularly noticeable at the elevated incubation temperature of 44.5 degrees C. These findings are contrary to previous work which concluded that the standard American Public Health Association membrane filtration procedure is suitable for recovery of fecal coliform indicator bacteria from ozonated wastewater.  相似文献   

18.
In vitro cholic acid (CA) transformation by mixed fecal culture was investigated. Concentrations of glucose, peptone, and yeast extract in the medium and the initial pH of the medium markedly affected the CA transformation. Yeast extract enhanced the transformation, whereas high concentrations of glucose and peptone inhibited it. When the initial pH of the medium was below 6.5, CA was converted to 7-keto-deoxycholic acid (7KD), and formation of deoxycholic acid (DC) was not observed. In contrast, with an initial pH of 7.0, about 60% of the CA was converted to 7KD after 3 days of incubation, and then DC gradually formed after 4 days of incubation, following the disappearance of 7KD. The formation of DC in the cultured samples was paralleled in each case by disappearance of 7KD. In pure culture systems, Escherichia coli and some strains of Bacteroides formed 7KD from CA. No DC formation was observed in pure cultures of any of the strains examined.  相似文献   

19.
Escherichia coli grown at pH 5·0 became acid-tolerant (acid-habituated) but, in addition, neutralized medium filtrates from cultures of E. coli grown to log-phase or stationary-phase at pH 5·0 (pH 5·0 filtrates) induced acid tolerance when added to log-phase E. coli growing at pH 7·0. In contrast, filtrates from pH 7·0-grown cultures were ineffective. The pH 5·0 filtrates were inactivated by heating in a boiling water-bath but there was less activity loss at 75 °C. Protease also inactivated such filtrates, which suggested that a heat-resistant protein (or proteins) in the filtrates was essential for the induction of acid tolerance. Filtrates from cells grown at pH 5·0 plus phosphate or adenosine 3':5'-cyclic monophosphate (cAMP) were much less effective in inducing acid tolerance, while the conversion of pH 7·0-grown log-phase cells to acid tolerance by pH 5·0 filtrates was inhibited by cAMP and bicarbonate. It seems likely that the acid tolerance response (acid habituation) involved the functioning of the extracellular protein(s) as protease reduces tolerance induction if added during acid habituation. Most inducible responses are believed to involve the functioning of only intracellular reactions and components ; the present results suggest that this is not the case for acid habituation, as an extracellular protein (or proteins) is needed for induction.  相似文献   

20.
The study of glutathione status in aerobically grown Escherichia coli cultures showed that the total intracellular glutathione (GSHin + GSSGin) level falls by 63% in response to a rapid downshift in the extracellular pH from 6.5 to 5.5. The incubation of E. coli cells in the presence of 50 mM acetate or 10 micrograms/ml gramicidin S decreased the total intracellular glutathione level by 50 and 25%, respectively. The fall in the total intracellular glutathione level was accompanied by a significant decrease in the (GSHin:GSSGin) ratio. The most profound effect on the extracellular glutathione level was exerted by gramicidin S, which augmented the total glutathione level by 1.8 times and the (GSHout:GSSGout) ratio by 2.1 times. The gramicidin S treatment and acetate stress inhibited the growth of mutant E. coli cells defective in glutathione synthesis 5 and 2 times more severely than the growth of the parent cells. The pH downshift and the exposure of E. coli cells to gramicidin S and 50 mM acetate enhanced the expression of the sodA gene coding for superoxide dismutase SodA.  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