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 46 毫秒
1.
Proton transport across charged membrane and pH oscillations.   总被引:1,自引:0,他引:1       下载免费PDF全文
Based on Eyring's multibarrier activation process, a mathematical model and equation is developed to account for proton diffusion through an immobilized protein and enzyme membrane perfused with an electrolyte, substrate, and a buffer. With this model we find that, in the presence of a buffer, our solution approaches the continuum case very rapidly. We apply our model to membranes composed of papain and bovine serum albumin and find that our theory closely stimulates the experimental observations on the effect of salt and buffer on proton diffusion. Our theory shows that the pH oscillations observed in the diffusion controlled papain-benzoyl-L-arginine ethyl ester (BAEE) reaction may be the result of CO2 dissolved in the bath at high pH. In our theory, under certain conditions and in agreement with experimental observation, the buffer penetration depth oscillates near the boundary of a papain membrane in a solution containing BAEE and borate. We also find that at low ionic strength small ions as well as a buffer are seen to oscillate if a membrane is highly charged.  相似文献   

2.
Y Pocker  N Janji? 《Biochemistry》1988,27(11):4114-4120
In our earlier paper we showed that the rates of CO2 hydration and HCO3- dehydration catalyzed by the high-activity form of mammalian erythrocyte carbonic anhydrase (CA II) were dependent on solution viscosity increase and that the effect was linked to some kind of proton-transfer-related event [Pocker, Y., & Janji?, N. (1987) Biochemistry 26, 2597-2606]. In order to further elucidate the source of the observed viscosity effect, the dependence of kcat and Km for CA II catalyzed HCO3- dehydration at pH 5.90 on sucrose-induced viscosity increase was investigated at several concentrations of 2-(N-morpholino)ethanesulfonic acid (MES) buffer, including the very low buffer concentration region (less than 10 mM) where the proton transfer between the shuttle group on the enzyme and buffer becomes rate limiting. In all examined cases, kcat steadily decreased with added sucrose while Km remained independent of the viscosity increase. The extent to which this reaction was dependent on viscosity was found to be constant, within experimental error, over the entire range of MES buffer concentrations studied (1-20 mM). Furthermore, the viscosity effect was qualitatively and quantitatively the same when an exceptionally large buffer (i.e., bovine serum albumin) was used instead of the more commonly used biological buffer (i.e., MES).(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

3.
R K Airas 《Biochemistry》1978,17(23):4932-4938
A kinetic study was performed on the reaction mechanism of pantothenase (EC 3.5.1.22) catalyzed hydrolysis of the pantothenic acid. A nonlinear progress curve is derived if the reaction occurs at low buffer concentrations. The nonlinearity is due to partial reversibility of the reaction; an acylenzyme (pantoyl-enzyme) is formed during the reaction, and beta-alanine, the other end product, is able to react with the acyl-enzyme and return back to pantothenate. The dependence of the beta-alanine return reaction on buffer concentration and on pH suggests a general acid catalysis during the reaction. A reaction mechanism is suggested, in which the -NH3+ form of beta-alanine participates in the return reaction, and the deacylation of the acyl-enzyme is acid catalyzed.  相似文献   

4.
Buffer power and intracellular pH of frog sartorius muscle.   总被引:2,自引:1,他引:1       下载免费PDF全文
Intracellular pH (pHi) and buffer power of frog muscle were measured using pH-sensitive microelectrodes under conditions used previously in energy balance experiments because pH strongly influences the molar enthalpy change for phosphocreatine splitting, the major net reaction during brief contractions. The extracellular pH (pHe) of HEPES buffered Ringer's solution influenced pHi, but change in pHi developed slowly. Addition or removal of CO2 or NH3 from the extracellular solution caused a rapid change in pHi. The mean buffer power measured with CO2 was 38.4 mmol.l-1.pH unit-1 (+/- SEM 2.1, n = 49) and with NH3 was 36.2 (+/- SEM 5.5, n = 4) at 20-22 degrees C. At 5 degrees C, in experiments with CO2 the mean buffer power was 40.3 (+/- SEM 2.6, n = 3). For pHi values above approximately 7.0, the observed buffer power was greater than that expected from the values in the literature for the histidine content of intracellular proteins, carnosine and inorganic phosphate in the sarcoplasm. The measured pHi values were similar to those assumed in energy balance calculations, but the high measured buffer power suggests that other buffering reactions occur in addition to those included in energy balance calculations.  相似文献   

5.
At pH 7.4, CO2, rather than HCO3-, markedly enhances the oxidation of diverse substrates by SOD1 plus H2O2. Since the concentration of CO2 would fall with rising pH in HCO3- buffers, it was of interest to explore the effects of pH on the peroxidase activity of SOD1 in the presence and in the absence of HCO3-. The rate of NADPH peroxidation in the HCO3- buffer was minimally affected by pH in the range of 8-10.5; in a pyrophosphate buffer, the rate increased markedly, such that at pH 10.5 the rates in the two buffers were nearly identical. Similar results were obtained when urate was used as the peroxidizeable substrate. These results are explicable on the basis of an increase in the rate with pH due to the ionization of H2O2 to the effective HO2- coupled with a decrease in [CO2] due to the ionizations of H2CO3, which displaces the hydration equilibrium to the right. These two opposing effects counteract in the HCO3(-)-buffered reaction mixtures; in the pyrophosphate buffer, only the effect of increasing [H02-] was seen.  相似文献   

6.
pH dependence of carbon monoxide binding to ferrous horseradish peroxidase   总被引:1,自引:0,他引:1  
The kinetic parameters of the reaction of horseradish peroxidase with CO have been determined at pH values between 10 and 3. At pH 7.0 the CO binding equilibrium constant L was measured using submicromolar concentrations of horseradish peroxidase; the value obtained corresponds to the ratio of the association and dissociation kinetic constants as expected for a simple binding mechanism to a monomeric hemeprotein. The CO association rate constant is pH-independent below pH 7, whereas in going from pH 7 to pH 11 a 2-fold increase can be detected, as previously reported (Kertesz, D., Antonini, E., Brunori, M., Wyman, J., and Zito, R. (1965) Biochemistry 4, 2672-2676). On the other hand, CO dissociation displays a peculiar pH rate profile characterized by a progressive decrease from pH 10 to pH 5 and by a very marked increase as the pH is further lowered to pH congruent to 3. Furthermore, the rate of CO dissociation is markedly enhanced in peroxidase reconstituted with protoheme dimethyl ester, suggesting a role of the propionates in the regulation of this process.  相似文献   

7.
The steady-state kinetics of CO2 hydration catalyzed by human carbonic anhydrase I (carbonate hydro-lyase, EC 4.2.1.1) has been investigated at three pH values corresponding to different parts of the pH-rate profile. Two buffer systems with similar pKa values were used at each pH. The results show that the catalyzed rates depend on the buffer concentration but also on the chemical nature of the buffer. For example, at pH 8.8 the buffer 1,2-dimethylimidazole behaves formally as a second substrate in a 'ping-pong' mechanism yielding a maximal kcat value of 2.2 x 10(5) s-1, whereas much lower rates were obtained with Taps buffers. Similarly, at pH 7.3 1-methylimidazole yields higher rates than Mops and at pH 6.3 3,5-lutidine is more efficient than Mes. Non-Michaelis-Menten kinetics were observed with all buffers except 1,2-dimethylimidazole. In addition, while the apparent buffer activation by 1,2-dimethylimidazole can be described by a single Km value of 26 mM, the Mes concentration dependence is consistent with the presence of two components of similar magnitudes with Km values of 45 mM and 0.15 mM. These results are interpreted within the framework of the 'zinc-hydroxide' mechanism in terms of multiple pathways for the rate-contributing transfer of a proton from the zinc-bound water molecule, formed during CO2/HCO3- interconversion, to the reaction medium, thus, regenerating zinc-bound OH-.  相似文献   

8.
ABSTRACT: BACKGROUND: Buffering to achieve pH control is crucial for successful trichloroethene (TCE) anaerobic bioremediation. Bicarbonate (HCO3-) is the natural buffer in groundwater and the buffer of choice in the laboratory and at contaminated sites undergoing biological treatment with organohalide respiring microorganisms. However, HCO3- also serves as the electron acceptor for hydrogenotrophic methanogens and hydrogenotrophic homoacetogens, two microbial groups competing with organohalide respirers for hydrogen (H2). We studied the effect of HCO3- as a buffering agent and the effect of HCO3--consuming reactions in a range of concentrations (2.5-30 mM) with an initial pH of 7.5 in H2-fed TCE reductively dechlorinating communities containing Dehalococcoides, hydrogenotrophic methanogens, and hydrogenotrophic homoacetogens. RESULTS: Rate differences in TCE dechlorination were observed as a result of added varying HCO3- concentrations due to H2-fed electrons channeled towards methanogenesis and homoacetogenesis and pH increases (up to 8.7) from biological HCO3- consumption. Significantly faster dechlorination rates were noted at all HCO3- concentrations tested when the pH buffering was improved by providing 4-(2-hydroxyethyl)-1-piperazineethanesulfonic acid (HEPES) as an additional buffer. Electron balances and quantitative PCR revealed that methanogenesis was the main electron sink when the initial HCO3- concentrations were 2.5 and 5 mM, while homoacetogenesis was the dominant process and sink when 10 and 30 mM HCO3- were provided initially. CONCLUSIONS: Our study reveals that HCO3- is an important variable for bioremediation of chloroethenes as it has a prominent role as an electron acceptor for methanogenesis and homoacetogenesis. It also illustrates the changes in rates and extent of reductive dechlorination resulting from the combined effect of electron donor competition stimulated by HCO3- and the changes in pH exerted by methanogens and homoacetogens.  相似文献   

9.
A previously reported method for iodination of the tyrosine moiety of oxidation-sensitive biomolecules was found to cause unacceptable damage to biomolecules containing thiols and thioether groups. This was due to the oxidation of the sulfur-containing residues by molecular iodine (I(2)). To selectively iodinate the tyrosine moiety with minimum oxidation to the sulfur functionality, studies of the kinetics of the reactions between I-(3) and various amino acids and small peptides at various pH values in phosphate buffer were undertaken. Within the pH range studied (5.5-8.2), the results showed that the iodination reaction is strongly catalyzed by hydroxide ions, whereas the oxidation of the sulfur group was insensitive to pH. The results also showed that both reactions are strongly catalyzed by HPO-(4) ion. In a complex molecule, such as methionine-enkephalin, oxidation of the methionine residue (undesirable reaction) proceeds in parallel with iodination of the tyrosine residue (desirable reaction). If such a molecule was iodinated in 0.01 M phosphate buffer at pH values above 7.5, the iodination reaction would proceed much more rapidly than the oxidation reaction, resulting in a high yield of iodinated substrate with little oxidative damage.  相似文献   

10.
1. The steady-state kinetics of the interconversion of CO2 and HCO3 catalyzed by human carbonic anhydrase C was studied using 1H2O and 2H2O as solvents. The pH-independent parts of the parameters k(cat) and Km are 3-4 times larger in 1H2O than in 2H2O for both directions of the reaction, while the ratios k(cat)/Km show much smaller isotope effects. With either CO2 or HCO3 as substrate the major pH dependence is observed in k(cat), while Km appears independent of pH. The pKa value characterizing the pH-rate profiles is approximately 0.5 unit larger in 2H2O than in 1H2O. 2. The hydrolysis of p-nitrophenyl acetate catalyzed by human carbonic anhudrase C is approximately 35% faster in 2H2O than in 1H2O. In both solvents the pKa values of the pH-rate profiles are similar to those observed for the CO2-HCO3 interconversion. 3. It is tentatively proposed that the rate-limiting step at saturating concentrations of CO2 or HCO3 is an intramolecular proton transfer between two ionizing groups in the active site. It cannot be decided whether the transformation between enzyme-bound CO2 and HCO3 involves a proton trnasfer or not.  相似文献   

11.
Amoxicillin can be produced in an enzymatic suspension-to-suspension reaction in which the substrate(s) and product(s) are mainly present as solid particles, while the reaction takes place in the liquid phase. During these suspension-to-suspension reactions different subprocesses take place, such as dissolution/crystallization of substrates and products, enzymatic synthesis of the product(s), and undesired enzymatic hydrolysis of substrates and/or products. All these subprocesses are influenced by pH and also influence the pH because the reactants are weak electrolytes. This paper describes a quantitative model for predicting pH and concentrations of reactants during suspension-to-suspension reactions. The model is based on mass and charge balances, pH-dependent solubilities of the reactants, and enzyme kinetics. For the validation of this model, the kinetically controlled synthesis of amoxicillin from 6-aminopenicillanic acid and D-(p)hydroxyphenylglycine methyl ester was studied. The pH and the dissolved concentrations took a very different course at different initial substrate amounts. This was described quite reasonably by the model. Therefore, the model can be used as a tool to optimize suspension-to-suspension reactions of weak electrolytes.  相似文献   

12.
The regulation of intracellular pH (pHi) is important in controlling muscle contraction. In these experiments, a spectrophotometric method of determining pHi was developed, and the method was then used to study muscle pHi regulation during CO2-induced changes in extracellular pH (pHb). Studies were performed in vitro on 27 diaphragm muscle strips obtained from adult hamsters. pHi was measured from the ratio of the absorbances of the acid (lambda = 530 nm) and alkaline (lambda = 460 nm) forms of a vital dye, neutral red, using the unstained diaphragm spectrum as a reference blank. A standard neutral red calibration curve constructed from eight diaphragm muscle homogenates indicated that the absorbance ratio was highly linear, with pH over the range 6.00-8.00. In intact muscle strips gassed with 95% O2-5% CO2, pHb was 7.45 +/- 0.03 (SE) and pHi was 7.00 +/- 0.01 (SE). When the muscle was aerated with CO2 concentrations from 3 to 30%, pHb and pHi changed rapidly and reached a steady state in 10-15 min. However, when pHb ranged from 6.80-7.80, pHi changed little from the value observed when pHb was 7.40. When pHb was less than 6.80 or greater than 7.80, changes in pHi and pHb were quantitatively similar. The results suggest that, in the isolated diaphragm, overall pHi is stable and effectively buffered over a wide range of CO2-induced changes in buffer solution pH.  相似文献   

13.
To test the hypothesis that histidine 64 in the active site of human carbonic anhydrase II functions as a proton-transfer group in the catalysis of CO2 hydration, we have studied a site-specific mutant having histidine 64 replaced by alanine, which cannot transfer protons. The steady-state kinetics of CO2 hydration has been measured as well as the exchange of 18O between CO2 and water at chemical equilibrium. The results show that the rate of exchange between CO2 and HCO3- at chemical equilibrium is essentially unaffected by the amino acid substitution at pH greater than 7.0 and slightly decreased in the mutant at pH less than 7.0 (by a factor of 2 at pH 6.0). However, in the absence of buffer the rate of release from the active site of water bearing substrate oxygen is smaller by as much as 20-fold for the mutant as compared to unmodified enzyme. Furthermore, in the unmodified enzyme water release is inhibited by micromolar concentrations of Cu2+ ions, but no such inhibition is observed with the alanine 64 variant. These results suggest that the mutation has specifically affected the rate of proton transfer between the active site and the reaction medium. This kinetic defect in the mutant can be overcome by increasing the concentration of certain buffers, such as imidazole and 1-methylimidazole, but not by others buffers, such as MOPS or HEPES. Similarly, the maximal rate of CO2 hydration at steady state catalyzed by the alanine 64 variant is very low in the presence of MOPS or TAPS buffers but considerably higher in the presence of imidazole derivatives.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

14.
In isolated perfused rat liver, urea synthesis from ammonium ions was dependent on extracellular HCO3- and CO2 concentrations when the HCO3-/CO2 ratio in the influent perfusate was constant (pH 7.4). Urea synthesis was half-maximal at HCO3- = 4 mM, CO2 = 0.19 mM and was maximal at HCO3- and CO2 concentrations above 20 mM and 0.96 mM, respectively. At physiological HCO3- (25 mM) and CO2 (1.2 mM) concentrations in the influent perfusate, acetazolamide, the inhibitor of carbonic anhydrase, inhibited urea synthesis from ammonium ions (1 mM) by 50-60% and led to a 70% decrease in citrulline tissue levels. Acetazolamide concentrations required for maximal inhibition of urea synthesis were 0.01-0.1 mM. At subphysiological HCO3- and CO2 concentrations, inhibition of urea synthesis by acetazolamide was increased up to 90%. Inhibition of urea synthesis by acetazolamide was fully overcome in the presence of unphysiologically high HCO3- and CO2 concentrations, indicating that the inhibitory effect of acetazolamide is due to an inhibition of carbonic-anhydrase-catalyzed HCO3- supply for carbamoyl-phosphate synthetase, which can be bypassed when the uncatalyzed intramitochondrial HCO3- formation from portal CO2 is stimulated in the presence of high portal CO2 concentrations. With respect to HCO3- supply of mitochondrial carbamoyl-phosphate synthetase, urea synthesis can be separated into a carbonic-anhydrase-dependent (sensitive to acetazolamide at 0.5 mM) and a carbonic-anhydrase-independent (insensitive to acetazolamide) portion. Carbonic-anhydrase-independent urea synthesis linearly increased with the portal 'total CO2 addition' (which was experimentally determined to be CO2 addition plus 0.036 HCO3- addition) and was independent of the perfusate pH. At a constant 'total CO2 addition', carbonic-anhydrase-dependent urea synthesis was strongly affected by perfusate pH and increased about threefold when the perfusate pH was raised from 6.9 to 7.8. It is concluded that the pH dependent regulation of urea synthesis is predominantly due to mitochondrial carbonic anhydrase-catalyzed HCO3- supply for carbamoyl phosphate synthesis, whereas there is no control of urea synthesis by pH at the level of the five enzymes of the urea cycle. Because HCO3- provision for carbamoyl phosphate synthetase increases with increasing portal CO2 concentrations even in the absence of carbonic anhydrase activity, susceptibility of ureogenesis to pH decreases with increasing portal CO2 concentrations. This may explain the different response of urea synthesis to chronic metabolic and chronic respiratory acidosis in vivo.  相似文献   

15.
16.
Transmembrane acid–base fluxes affect the intracellular pH and unstirred layer pH around a superfused biological preparation. In this paper the factors influencing the unstirred layer pH and its gradient are studied. An analytical expression of the unstirred layer pH gradient in steady state is derived as a function of simultaneous transmembrane fluxes of (weak) acids and bases with the dehydration reaction of carbonic acid in equilibrium. Also a multicompartment computer model is described consisting of the extracellular bulk compartment, different unstirred layer compartments and the intracellular compartment. With this model also transient changes and the influence of carbonic anhydrase (CA) can be studied. The analytical expression and simulations with the multicompartment model demonstrate that in steady state the unstirred layer pH and its gradient are influenced by the size and type of transmembrane flux of acids and bases, their dissociation constant and diffusion coefficient, the concentration, diffusion coefficient and type of mobile buffers and the activity and location of CA. Similar principles contribute to the amplitude of the unstirred layer pH transients. According to these models an immobile buffer does not influence the steady-state pH, but reduces the amplitude of pH transients especially when these are fast. The unstirred layer pH provides useful information about transmembrane acid–base fluxes. This paper gives more insight how the unstirred layer pH and its transients can be interpreted. Methodological issues are discussed.  相似文献   

17.
It is often assumed that pH(i) is spatially uniform within cells. A double-barreled microperfusion system was used to apply solutions of weak acid (acetic acid, CO(2)) or base (ammonia) to localized regions of an isolated ventricular myocyte (guinea pig). A stable, longitudinal pH(i) gradient (up to 1 pH(i) unit) was observed (using confocal imaging of SNARF-1 fluorescence). Changing the fractional exposure of the cell to weak acid/base altered the gradient, as did changing the concentration and type of weak acid/base applied. A diffusion-reaction computational model accurately simulated this behavior of pH(i). The model assumes that H(i)(+) movement occurs via diffusive shuttling on mobile buffers, with little free H(+) diffusion. The average diffusion constant for mobile buffer was estimated as 33 x 10(-7) cm(2)/s, consistent with an apparent H(i)(+) diffusion coefficient, D(H)(app), of 14.4 x 10(-7) cm(2)/s (at pH(i) 7.07), a value two orders of magnitude lower than for H(+) ions in water but similar to that estimated recently from local acid injection via a cell-attached glass micropipette. We conclude that, because H(i)(+) mobility is so low, an extracellular concentration gradient of permeant weak acid readily induces pH(i) nonuniformity. Similar concentration gradients for weak acid (e.g., CO(2)) occur across border zones during regional myocardial ischemia, raising the possibility of steep pH(i) gradients within the heart under some pathophysiological conditions.  相似文献   

18.
Fluctuations in pH and head-space pressure in a fermentor introduce temporary changes in off-gas CO(2) concentrations. These changes are quantified using a simple model based on kinetics of CO(2) hydration and gas-liquid mass transfer. The model is verified experimentally. An eigenvalue analysis of the model indicates that mass transfer is the parameter which controls the dynamics of CO(2) equilibration.  相似文献   

19.
Stationary phase cultures of a clonal line of rat astrocytes (C6) were maintained at pH values ranging from 6.0 to 8.4 using media buffered with various combinations of organic buffers or graded concentrations of bicarbonate ion at a constant CO2 tension. The accumulation of a soluble acidic protein unique to the nervous system (S-100) in media buffered with organic buffers was optimal in the pH range 6.4 to 6.8, significantly more acid than that optimal for cell growth (pH 7.0 to 7.8). Cells maintained in CO2-bicarbonate-buffered media exhibited a higher and less marked pH optimum for S-100 protein accumulation and a lower efficiency of accumulation of the protein. These data suggest that the organic buffer ions themselves, apart from their function as buffers, are influencing the accumulation of S-100. The specific activity (assayed at the enzymatic pH optimum) of a membrane-bound enzyme enriched in glial cells and myelin, 2',3'-cyclic nucleotide 3'-phosphohydrolase, was markedly pH-dependent. The optimal pH range was 6.4 to 6.7 in organic buffer controlled media. In CO2-bicarbonate controlled media the optimal pH range was only slightly higher (pH 6.6 to 7.0), but the specific activities were reduced relative to organic buffer-grown cells. The structural relationship of some of the aminoethanesulfonic acid buffers used in these experiments to certain compounds of neurochemical interest (such as taurine and alpha-flupenthixol) is noted.  相似文献   

20.
Dicarboxylic acids that are produced from renewable resources are becoming attractive building blocks for the polymers industry. In this respect, fumaric acid is very interesting. Its low aqueous solubility facilitates product recovery. To avoid excessive waste salt production during downstream processing, a low pH for fumaric acid fermentation will be beneficial. Studying the influence of pH, working volume and shaking frequency on cell cultivation helped us to identify the best conditions to obtain appropriate pellet morphologies of a wild type strain of Rhizopus oryzae. Using these pellets, the effects of pH and CO(2) addition were studied to determine the best conditions to produce fumaric acid in batch fermentations under nitrogen-limited conditions with glucose as carbon source. Decreasing either the fermentation pH below 5 or increasing the CO(2) content of the inlet air above 10% was unfavourable for the cell-specific productivity, fumaric acid yield, and fumaric acid titer. However, switching off the pH control late in the batch phase did not affect these performance parameters and allowed achieving pH of 3.6. A concentration of 20 gL(-1) of fumaric acid was obtained at pH 3.6 while the average cell mass specific productivity and fumaric acid yield were the same as at pH 5.0. Consequently, relatively modest amounts of inorganic base were required for pH control, while recovery of the acid should be relatively easy at pH 3.6.  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