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.
Direct microcalorimetric measurements were made of the reaction between acetylcholine chloride and acetylcholinesterase (EC 3.1.1.7) that was extracted from electric eel (Electrophorus electricus) and purified by affinity chromatography. Tris-HCl, sodium phosphate and potassium phosphate were used as buffers and sources of ions for the reaction. At pH 7.2 and in 0.1-0.2 M phosphate buffer, the delta H for acetylcholine hydrolysis was found to be -0.107 kcal/mol (under buffered conditions) and -0.931 kcal/mol under unbuffered conditions (water). At pH 8.0 in 0.1 M Tris-HCl buffer, values greater than -2.5 kcal/mol were obtained, with the highest value of -9.2 kcal/mol being seen with bovine erythrocyte acetylcholinesterase. Tris-HCl buffer at 4 X 10(-2) M enhanced the reaction velocity by 51.2% over that of 4 X 10(-3) M buffer. Enzyme purity, pH and ionic milieu of reaction mixture, and substrate concentration affected the measured delta H value.  相似文献   

2.
The metabolic flux through the alpha-ketoglutarate dehydrogenase reaction in perfused livers was monitored by measuring the rate of 14CO2 production from [1-14C]alpha-ketoglutarate. The rates of 14CO2 production and glucose production from [1-14C]alpha-ketoglutarate were increased with increasing perfusate alpha-ketoglutarate concentrations. Vasopressin, angiotensin II, and the alpha 1-adrenergic agonist phenylephrine stimulated transiently by 2.5-fold the metabolic flux through the alpha-ketoglutarate dehydrogenase reaction in the presence and absence of Ca2+ in the perfusion medium. High concentrations of glucagon (1 x 10(-8) M) and 8-p-chlorophenylthio-cAMP (100 microM) (data not shown) also stimulated transiently the metabolic flux through the alpha-ketoglutarate dehydrogenase reaction. However, lower glucagon concentrations (1 x 10(-9) M) stimulated the rate of 14CO2 production from [1-14C]alpha-ketoglutarate only under conditions optimized to fix the cellular oxidation-reduction state at an intermediate level, when glucagon (1 x 10(-9) M)-mediated elevation of cAMP content was greater than that observed under highly oxidizing and reducing conditions. These data indicate that agonists which increase cytosolic free Ca2+ levels stimulate the metabolic flux through the alpha-ketoglutarate dehydrogenase complex. Furthermore, the data presented here demonstrate for the first time that physiological glucagon concentrations stimulate the metabolic flux through the alpha-ketoglutarate dehydrogenase reaction only under conditions known to be optimal for glucagon-mediated Ca2+ mobilization in the isolated perfused rat liver.  相似文献   

3.
Studies of the alpha-ketoglutarate dehydrogenase complex have demonstrated that inorganic phosphate ions cause a decrease in the Km value for alpha-ketoglutarate without changing the maximum reaction rate. In the absence of phosphate (tris-HCl buffer) at low concentrations of alpha-ketoglutarate there are some indications of enzyme-substrate cooperative interactions (the Hill coefficient is 1,6). The cooperativity is removed by ADP, which increases the apparent affinity of the enzyme for alpha-ketoglutarate. Upon divalent cations binding to EDTA in the presence of high (20 mM) concentrations of alpha-ketoglutarate the reaction rate is decreased only by 20%, while the value of Km for the given substrate shows a sharp rise. The nature of Mg2+, Ca2+, Ba2+ and Mn2+ effects on the alpha-ketoglutarate dehydrogenase complex activity depends on their concentration.  相似文献   

4.
W J Ray  J M Puvathingal 《Biochemistry》1990,29(11):2790-2801
The inhibitor complex produced by the binding of alpha-D-glucose 1-phosphate 6-vanadate to the dephospho form of muscle phosphoglucomutase exhibits an unusually small dissociation constant: about 15 fM for the Mg2+ enzyme at pH 7.4, when calculated in terms of the tetraanion. Such tight binding suggests that the enzyme/vanadate/glucose phosphate complex mimics a state that at least approaches the transition state for (PO3-) transfer in the normal enzymic reaction. This hypothesis also is supported by the observation that replacement of Mg2+, the normal metal ion activator, by Li+, a poor activator, substantially reduces the binding constant for the glucose phosphate/vanadate mixed diester. Other indicators that support this hypothesis are described. One is the derived equilibrium constant for replacement of a PO4(2-) group in bound glucose bisphosphate by VO4(2-): 3 x 10(6) when the replaced group is the phosphate at the (PO3-) transfer site of the Mg2+ enzyme--in contrast to about 10 for the same replacement (of PO4(2-) by VO4(2-)) in an aqueous solution of a phosphate ester. Another is the greatly decreased rate at which Mg2+ dissociates from the glucose phosphate/vanadate complex of the enzyme, relative to the rate at which it dissociates from the corresponding bisphosphate complex (rate ratio less than or equal to 3 x 10(-4)), presumably because Mg2+ binds more tightly to the glucose phosphate/vanadate complex than to the corresponding bisphosphate complex. This apparent increase in Mg2+ binding occurs in spite of what appears to be a reduced charge density at the bound vanadate grouping, relative to the bound phosphate grouping, and in spite of the somewhat weaker binding of Mg2+ by dianionic vanadate than by the phosphate dianion. Although a direct assessment of the binding constant for Mg2+ was not possible, the equilibrium constant for Mg2+/Li+ exchange could be evaluated for the complexes of dephospho enzyme with glucose bisphosphate or glucose 1-phosphate 6-vanadate. The results suggest that the glucose phosphate/vanadate complex of the Mg2+ enzyme mimics a state about halfway between the ground state and the transition state for (PO3-) transfer. This estimate also is in accord with the binding of glucose phosphate/vanadate relative to that expected for transition-state binding of glucose bisphosphate. A possible scenario for the (PO3-) transfer catalyzed by the Mg2+ form of phosphoglucomutase is discussed, on the basis of these observations, together with possible reasons why the bound vanadate group appears to mimic an intermediate state for (PO3-) transfer rather than the ground state for phosphate binding.  相似文献   

5.
Thymidilate synthetase (methylenetetrahydrofolate:dUMP C-methyltransferase) in crude extract from Diplococcus pneumoniae exhibits a partial but variable requirement for Mg-2+ depending upon the buffer. Optimum Mg-2+ concentration is between 0.014 and 0.02 M. The optimum pH for activity in a variety of buffers occurred as a broad peak between 7.0 and 7.7. In Tris/acetate buffer, but not in potassium phosphate buffer, the pH optimum was different in the presence and absence of Mg-2+. Methylation of uridylate, cytidylate and deoxycytidylate could not be demonstrated over a pH range of 5.0-8.0. The enzyme exhibited an apparent Km for deoxyuridylate of 3.08 - 10-5 M and an apparent Km for L-(+)(minus)-5,10-methylene tetrahydrofolate of 2.66 - 10-4 M. During molecular-sieve chromatography and sucrose density-gradient centrifugation, the enzyme was detectable only as a single catalytically active form of Mr 34 000-38 000. 2,4-Diamino quinazoline antifolates were better competitive inhibitors (Ki = 3-8 -10-6 M) of thymidylate synthetase than 2,4-diamino pteridines (Ki = 3- 10-5 M). 2-Amino-4-hydroxy-quinazolines were the best inhibitors (Ki = 1.3-2.9 - 10-6 M). All of the 2,4-diamino quinazolines and pteridines inhibited dihydrofolate reductase from D. pneumoniae in a nearly stoichiometric fashion (Ki = less than 10-10 M). The 2-amino-4-hydroxy-quinazolines were poor inhibitors of this enzyme (Ki = 10=5 M).  相似文献   

6.
The second order rate constant (k(cat)/K(m)) for decarboxylation of orotidine by yeast OMP decarboxylase (ODCase), measured by trapping (14)CO(2) released during the reaction, is 2 x 10(-4)M(-1)s(-1). This very low activity may be compared with a value of 3 x 10(7)M(-1)s(-1) for the action of yeast OMP decarboxylase on the normal substrate OMP. Both activities are strongly inhibited by 6-hydroxy UMP (BMP), and abrogated by mutation of Asp-96 to alanine. These results, in conjunction with the binding affinity of inorganic phosphate as a competitive inhibitor (K(i)=7 x 10(-4)M), imply an effective concentration of 1.1 x 10(9)M for the substrate phosphoryl group in stabilizing the transition state for enzymatic decarboxylation of OMP. The observed difference in rate (1.5 x 10(11)-fold) is the largest effect of a simple substituent that appears to have been reported for an enzyme reaction.  相似文献   

7.
The binding of pig skeletal muscle lactate dehydrogenase by F-actin has been studied using the sedimentation method in 10 mM Tris-acetate buffer, pH 6.0 at 20 degrees C. Adsorption capacity of F-actin is equal to (1 +/- 0.1) . 10(-5) moles of lactate dehydrogenase per 1 g of actin. NADH decreases the affinity of F-actin with respect to lactate dehydrogenase. The binding of lactate dehydrogenase by F-actin in diminishing the rate of enzymatic reduction of alpha-ketoglutarate. The microscopic dissociation constant for the complex of the enzyme with F-actin which is estimated from the dependence of the enzymatic reaction rate of F-actin concentration at saturating NADH concentrations is equal (3.0 +2- 0.5) . 10(-7) M. It has been shown that the bound enzyme is characterized by the greater value of Km and the lower value of Vmax in comparison to the free enzyme.  相似文献   

8.
Quantitative nutrient requirements for unrestricted autotrophic growth of Alcaligenes eutrophus were determined. Minimum saturating concentrations of Mg2+, SO42-, PO43-, Fe3+, and Na2+ for an optical density increase of 2 were 10(-4) M 8 X10(-5) M, 5 X 10(-4) to 6 X 10(-4) M, 10(-5) M, and 10(-7) to 2 X 10(-7) M, respectively. Trace metal requirements for cobalt, chromium, and copper were also demonstrated, but minimum concentrations could not be determined because other reagents contributed a high background of these metals. Under certain conditions an apparent response to zinc was observed, although other experiments suggest the zinc salt contained another metal that was required for growth. Poly-beta-hydroxybutyrate biosynthesis was shown to be initiated by a magnesium or sulfate deficiency as well as by a nitrogen or phosphate deficiency.  相似文献   

9.
M E Marsh 《Biochemistry》1989,28(1):346-352
The concomitant binding of calcium and inorganic phosphate ions by the highly phosphorylated rat dentin phosphophoryn (HP) was measured in the pH range of 7.4-8.5 by an ultrafiltration procedure. HP binds almost exclusively the triply charged PO4(3-) ion, and for each PO4(3-) ion bound, the protein binds about 1.5 additional Ca2+ ions. Therefore, the protein-mineral ion complex can be described as a protein with two different ligands, Ca2+ ions and calcium phosphate clusters having a stoichiometry of about Ca1.5PO4. Empirically the binding of calcium and phosphate can best be described as a function of a neutral ion activity product in which 2.5-10% of the phosphate is HPO4(2-). The stoichiometry of the bound clusters is similar to that of amorphous calcium phosphate, and it is clear that the protein does not sequester crystal embryos of octacalcium phosphate or hydroxyapatite. The protein-mineral ion complex is amorphous by electron diffraction analysis and does not catalyze the formation of a crystalline phase when aged in contact with its solution. About 15% of the bound phosphate is buried in protected domains, and it is stable with respect to dissociation for extended periods in phosphate-free calcium buffers. The buried mineral maintains the protein in an aggregated state even at calcium ion concentrations which are too low for the aggregation of unmineralized HP. In vivo HP should be ineffective in the nucleation of a crystalline mineral phase, if it is secreted in a mineralized aggregated state similar to casein and the bivalve phosphoprotein.  相似文献   

10.
Freeze-concentrated aqueous phosphate-buffered (pH 6.8) solutions of carbonyl-hemoglobin (HbCO) and -myoglobin (MbCO) were investigated by Fourier-transform infrared spectroscopy for the effect of alkali cation on the population of conformers. When using sodium phosphates as buffer components, HbCO was transformed from conformer III (at approximately 1951 cm-1) which is the dominant form at ambient temperatures, into conformer IV (at buffer concentration at a given temperature. The conformational changes started slightly below the temperature where ice began to crystallize and the remaining solution became freeze-concentrated, and they were reversible for HbCO. For MbCO in 0.5 M sodium phosphate buffer solution, however, they were irreversible and MbCO denatured completely. When potassium phosphate salts were used for preparing the buffer at the same pH of 6.8, little or no transformation of conformer III into conformer IV was observed. The conformational changes induced by sodium salts are attributed to a decrease in pH and it is shown by infrared spectroscopy that during freeze concentration drastic changes in composition of the two buffer components H2PO4-/HPO(2)4- occur, the acid component increasing strongly relative to the base component. Supersaturation is also important because change from conformer III to IV requires a minimum concentration of sodium salts: whereas 0.1 M sodium phosphate buffer concentration shows a strong effect, 0.03 M concentration does not and therefore behaves like a potassium phosphate buffer.  相似文献   

11.
The rate of the 4 to 5 S estrogen-binding protein (EBP) in vitro transformation was measured by sucrose gradient centrifugation analysis. The temperature-activated 4 to 5 S EBP transformation is found to be highly reproducible without loss of [3H]estradiol-binding activity in a buffer containing an excess of [3H]estradiol, 40 mM Tris, 1 mM dithiothreitol, and 1 M urea at pH 7.4. The presence of [3H]estradiol is necessary for the 4 to 5 EBP transformation. A kinetic analysis of the 4 to 5 EBP transformation shows that it is a bimolecular reaction, the dimerization of the 4 S EBP with a second (similar or dissimilar) monomer or subunit. In buffers containing 0.4 M KCl the apparent second order rate constant is 2.3 plus or minus 0-2 times 10-7 M minus 1 min minus 1 at 28 degrees. The reaction is independent of the initial receptor concentration, suggesting that the 4 S EBP is dissociated into monomeric units in buffers of high ionic strength. In buffers without KCl or with 0.1 M KCl the apparent second order rate constant of receptor transformation increases with decreasing receptor concentration. This suggests that the 4 S EBP is associated weakly with another macromolecule (or macromolecules) in buffers of low ionic strength. The rate of 4 to 5 S EBP transformation shows a 200-fold increase between 0 and 35 degrees. The Arrhenius energy of activation is 21.3 kcal mol minus 1 in buffer without KCl and 19.1 kcal mol minus 1 in buffer with 0.4 M KCl. Following the temperature-activated dimerization, the avidity of binding between the 4 S EBP and its complementary subunit is increased, 0.4 M KCl can no longer cause dissociation, and the 5 S EBP dimer appears. This kinetic analysis indicates that the avidity of binding between the subunits of the estrogen receptor is modulated by estradiol, temperature, and ionic strength. We propose that these interactions of the estrogen receptor's subunits reflect conformational changes involved in receptor activation.  相似文献   

12.
The solubility of palmitoyl-CoA is strongly affected by Mg2+ concentrations commonly used in acyltransferase reactions. In 0.10 M Tris-HCl buffer at pH 7.4 or 8.5, all of the palmitoyl-CoA in 10 microM solutions and 90% of the palmitoyl-CoA in 100 microM solutions are precipitated by 1 mM Mg2+. In 0.05 M phosphate at pH 7.4, and in 0.10 M Tris-HCl containing 0.4 M KCl, the substrate remains soluble at Mg2+ concentrations below 4-5 mM. Above 5 mM Mg2+, palmitoyl-CoA is insoluble in all of these buffers. Substrate solubility could therefore be a limiting factor when free Mg2+ and fatty acyl-CoAs are present together during acyltransferase assays.  相似文献   

13.
Seven strains of viridans streptococci of the species Streptococcus sanguis, S. mutans and S. mitis were investigated for autolysis. The effect of pH, salt concentration and temperature on the autolytic process was studied in Na2HPO4/NaH2PO4 buffer. Whole cells and walls of all strains autolysed most rapidly at pH values above 7. Autolysis of whole cells of S. sanguis and one strain of S. mitis (ATCC15909) was maximal in 0-05 TO 0-2 M buffer, while the two S. mutans strains and S. mitis ATCC15912 showed maximal autolysis in 0-5 and 1-0 M buffers. Cultures harvested in the stationary phase of growth possessed only slightly decreased autolytic activity compared with those from the exponential phase. Whole cells autolysed more rapidly at 37 degrees C Than at 45 degrees C and 10 degrees C. Autolysis of isolated walls of three strains of S. mitis (ATCC903, ATCC15909 and ATCC15912) was maximal at pH 7-0 AND 7-5 and in 1-0 M buffers. Streptococcus mitis ATCC15909 also showed maximal lysis in 0-01 M and 0-5 M buffers. An endopeptidase action of the autolytic system of S. mitis ATCC15912 was indicated by the progressive release of soluble amino groups during autolysis of the walls. No release of reducing groups was observed. Several free amino acids were released during autolysis of these walls, alanine, lysine and glutamic acid being in greatest quanitity.  相似文献   

14.
Wang W  Seah SY 《Biochemistry》2005,44(27):9447-9455
HpaI, a class II pyruvate-specific aldolase involved in the catabolic pathway of hydroxyphenylacetate, is overexpressed and purified. A previous suggestion that phosphate is involved in proton transfer of pyruvate, based on the crystal structure of the homologous 2-dehydro-3-deoxygalactarate aldolase, is not substantiated from biochemical studies with HpaI. Thus, specific activities of the enzyme for the substrate 4-hydroxy-2-ketopentanoate in sodium HEPES and Tris-acetate buffers are higher than in sodium phosphate buffer. The enzyme also catalyzed the partial reaction of pyruvate proton exchange with an initial rate of 0.77 mmol min(-)(1) mg(-)(1) in phosphate-free buffer, as monitored by nuclear magnetic resonance. Steady-state kinetic analysis shows that the enzyme is also able to catalyze the aldol cleavage of 4-hydroxy-2-ketohexanoate and 3-deoxy-d-manno-oct-2-ulosonic acid (KDO). The enzyme exhibits significant oxaloacetate decarboxylase activity, with a k(cat) value 2.4-fold higher than the corresponding value for the aldol cleavage of 4-hydroxy-2-ketopentanoate. Sodium oxalate, an analogue of the enolate intermediate of the enzyme-catalyzed reaction, is a competitive inhibitor of the enzyme, with a K(i) value of 5.5 microM. Replacement of an active site arginine residue (R70) with alanine by site-specific mutagenesis resulted in an enzyme that lacks both aldolase and decarboxylase activities. The mutant enzyme is also unable to catalyze pyruvate proton exchange. The dissociation constant for pyruvate in the R70A mutant, determined by fluorescence titration, is similar to that of the wild-type enzyme, indicating that pyruvate binding is not affected by this mutation. Together, the results show that R70 influences catalysis in HpaI, particularly at the pyruvate proton exchange step.  相似文献   

15.
V B Lawlis  T E Roche 《Biochemistry》1981,20(9):2519-2524
Micromolar Ca2+ markedly reduces NADH inhibition of bovine kidney alpha-ketoglutarate dehydrogenase complex [Lawlis, V. B., & Roche, T. E. (1980) Mol. Cell. Biochem. 32, 147-152]. Product inhibition patterns from initial velocity studies conducted at less than 10(-9) M or at 1.5 X 10(-5) M Ca2+ with NAD+, CoA, or alpha-ketoglutarate as the variable substrate showed that NADH was a noncompetitive inhibitor with respect to each of these substrates, except at high NAD+ concentrations, where reciprocal plots were nonlinear and the inhibition pattern for NADH vs. NAD+ changed from a noncompetitive to a competitive pattern. From slope and intercept replots, 2-fold to 12-fold higher inhibition constants were estimated for inhibition by NADH vs. the various substrates in the presence of 1.5 X 10(-5) M Ca2+ than for inhibition at less than 10(-9) M Ca2+. These inhibition patterns and the lack of an effect of Ca2+ on the inhibition of the dihydrolipoyl dehydrogenase component suggested that Ca2+-modulated NADH inhibition occurs at an allosteric site with competitive binding at the site by high levels of NAD+. Decarboxylation of alpha-keto[1-14C]glutarate by the resolved alpha-ketoglutarate dehydrogenase component was investigated in the presence of 5.0 mM glyoxylate which served as an efficient acceptor. NADH (0.2 mM) or 1.0 mM ATP inhibited the partial reaction whereas 15 muM Ca2+, 1.0 mM ADP, or 10 mM NAD+ stimulated the partial reaction and reduced NADH inhibition of this reaction. Thus these effectors alter the activity of the alpha-ketoglutarate dehydrogenase complex by binding at allosteric sites on the alpha-ketoglutarate dehydrogenase component. Inhibition by NADH over a wide range of NADH/NAD+ ratios was measured under conditions in which the level of alpha-ketoglutarate was adjusted to give matching control activities at less than 10(-9) M Ca2+ or 1.5 X 10(-5) M Ca2+ in either the presence or the absence of 1.6 mM ADP. These studies establish that both Ca2+ and ADP decreased NADH inhibition under conditions compensating for the effects of Ca2+ and ADP on S0.5 for alpha-ketoglutarate. ADP was particularly effective in reducing NADH inhibition; further studies are required to determine whether this occurs through binding of NADH and ADP at the same, overlapping, or interacting sites.  相似文献   

16.
Four proteoglycans and hyaluronan synthesized by cultured human bone cells were isolated using a two-step high-performance liquid chromatography system involving desalting and buffer exchange with a TSK-GEL HW 40(S) column followed by ion-exchange separation on a Nucleogen 4000-10 DEAE column. The desalting of 4 M guanidinium HCl extracts by a TSK-GEL HW 40(S) column equilibrated in a formamide:KH2PO4 buffer produces greater than 95% recoveries, enables quantitation of label incorporation and requires only 40 min to complete. The Nucleogen 4000-10 DEAE column utilizes the same buffer system and requires only 100 min for the resolution of four distinct types of proteoglycans. The formamide:KH2PO4 buffer system is compatible with a previously developed polyacrylamide gel system for the electrophoretic profiling of proteoglycans. After separation by charge density, proteoglycans were further resolved by size distribution using a calibrated TSK-GEL HW 75(F) column which also enabled the estimation of the apparent Mr of hyaluronan produced by the bone cells. The same TSK-GEL HW 40(S) resin is used to exchange pooled proteoglycans into buffers for analyzing enzyme digests of glycosaminoglycan chains and core proteins. The technique has been applied to the analysis of biosynthetically labeled proteoglycans produced in culture by fetal and adult human bone cells. A distinct pattern of proteoglycan size and secretion for both cell types could be shown using this method. The method of analysis is useful for high yield and rapid screening of various cell types for both biosynthetic rate studies and analysis of patterns of proteoglycan synthesis.  相似文献   

17.
1. The thermal denaturation of DNA from rat liver was studied spectrophotometrically. In sodium phosphate buffers denaturation led to a single-stranded form having, at 25 degrees , about 25% of the hypochromism of the intact double helix. 2. The hypochromism of the denatured form was the same in 1mm- as in 10mm-sodium phosphate buffer and was scarcely affected by reaction with formaldehyde. The hypochromism was decreased by about 40% in the presence of 8m-urea. 3. The hypochromism of denatured DNA at low ionic strengths was about the same as that of fragments of reticulocyte ribosomal RNA that were too short to form double-helical secondary structure and about the same as that of RNA after reaction with formaldehyde. 4. The spectrum of DNA was slightly affected by the presence of 8m-urea or 4m-guanidinium chloride. The differences in the spectrum of the native and denatured forms of DNA in 0.1m-sodium phosphate buffer, in 8m-urea-10mm-sodium phosphate buffer and in 4m-guanidinium chloride-10mm-sodium phosphate buffer, pH7.6, were similar but not identical. 5. Denatured rat liver DNA appears to have no double-helical character at 25 degrees in 10mm-sodium phosphate buffer, pH7.6; increasing the buffer concentration to 0.1m leads to a more compact form in which about 40% of the residues form base pairs.  相似文献   

18.
The rate of oxidative phosphorylation was studied in rat liver mitochondria incubated with free Ca2+ concentrations that range from 10(-9) to 5 X 10(-6) M. The highest rate was observed between 0.5-1.0 microM Ca2+. ATP synthesis was measured by polarographic and spectrophotometric techniques and by uptake of radioactive inorganic phosphate. The concentration of Ca2+ at which maximal rates of ATP synthesis take place is modified by Mg2+ and phosphate. The dependence of oxidative phosphorylation on Ca2+ was observed with alpha-ketoglutarate, glutamate + malate, and succinate, but not with beta-hydroxybutyrate. At 10(-9) M Ca2+ there is a continuous exit of endogenous Ca2+, while with 10(-6) M Ca2+, intramitochondrial Ca2+ levels remained constant throughout time. Apparently the control of the level of internal Ca2+ by external Ca2+ modulates the rate of oxidative phosphorylation. Uncoupler-stimulated respiration also depends on Ca2+ concentration, even though at 10(-9) to 10(-6) M Ca2+ the rate of oxidative phosphorylation is lower than the rate of uncoupled respiration. The contribution of the ADP/ATP carrier and the ATP synthase to the kinetic regulation of ATP synthesis at 10(-9) and 10(-6) M Ca2+ was evaluated by titrations with carboxyatractyloside and oligomycin, respectively. The contribution of the carrier and the synthase to the regulation of the final rate of ATP synthesis was different at the two concentrations of Ca2+; therefore, the concentration of extramitochondrial Ca2+ influences the overall kinetics of oxidative phosphorylation.  相似文献   

19.
The effect of insulin on rat adipocytes was studied in isotonic buffers (pH 7.4) containing NaCl, CaCl2, MgSO4, KCl, and bovine serum albumin but no phosphate or bicarbonate anions. In phosphate- and bicarbonate-free buffers the dose-response curve to insulin is shifted to the right, the effects of the hormone on hexose uptake, glucose metabolism, and inhibition of lipolysis being observed at much higher (nearly 2 orders of magnitude) concentrations of insulin. The insulin binding capacity of the cells is only slightly changed. The dose-response curve for isoproterenol which stimulates lipolysis in the same cell type is almost the same in both Krebs-Ringer bicarbonate buffer and phosphate- and bicarbonate-free buffers. The dose-response curves for agents that mimic the action of insulin such as wheat germ agglutinin or vanadate ions are also shifted to the right. The dose-response curve to insulin can be returned to "normal" by readdition of either bicarbonate or phosphate. Almost complete recovery is obtained at either 10 mM bicarbonate or 24 mM phosphate, respectively. External Ca2+ ions which are not required for the proper action of insulin in fat cells maintained in Krebs-Ringer bicarbonate buffer, become essential for insulin action in bicarbonate-free buffer. The study indicates that depletion of bicarbonate and, to a lesser extent, phosphate anions, interferes with an essential insulin-dependent post-binding event. Also, in bicarbonate-free medium, external Ca2+ ions are essential for insulin-mediated processes. The implications of this study to the mode of action of insulin, and to physiological and clinical states of insulin desensitization are discussed.  相似文献   

20.
Net rate constants that define the steady-state rate through a sequence of steps and the corresponding effective energy barriers for two (PO3-)-transfer steps in the phosphoglucomutase reaction were compared as a function of metal ion, M, where M = Mg2+ and Cd2+. These steps involve the reaction of either the 1-phosphate or the 6-phosphate of glucose 1,6-bisphosphate (Glc-P2) bound to the dephosphoenzyme (ED) to produce the phosphoenzyme (EP) and the free monophosphates, glucose 1-phosphate (Glc-1-P) or glucose 6-phosphate (Glc-6-P): EP.M + Glc-1-P----ED.M.Glc-P2----EP.M.Glc-6-P6. Before this comparison was made, net rate constants for the Cd2+ enzyme, obtained at high enzyme concentration via 31P NMR saturation-transfer studies [Post, C. B., Ray, W. J., Jr., & Gorenstein, D. G. (1989) Biochemistry (preceding paper in this issue)], were appropriately scaled by using the observed constants to calculate both the expected isotope-transfer rate at equilibrium and the steady-state rate under initial velocity conditions and comparing the calculated values with those measured in dilute solution. For the Mg2+ enzyme, narrow limits on possible values of the corresponding net rate constants were imposed on the basis of initial velocity rate constants for the forward and reverse directions plus values for the equilibrium distribution of central complexes, since direct measurement is not feasible. The effective energy barriers for both the Mg2+ and Cd2+ enzymes, calculated from the respective net rate constants, together with previously values for the equilibrium distribution of complexes in both enzymic systems [Ray, W. J., Jr., & Long, J. W. (1976) Biochemistry 15, 4018-4025], show that the 100-fold decrease in the kappa cat for the Cd2+ relative to the Mg2+ enzyme is caused by two factors: the increased stability of the intermediate bisphosphate complex and the decreased ability to cope with the phosphate ester involving the 1-hydroxyl group of the glucose ring. In fact, it is unlikely that the efficiency of (PO3-) transfer to the 6-hydroxyl group of bound Glc-1-P (thermodynamically favorable direction) is reduced by more than an order of magnitude in the Cd2+ enzyme. By contrast, the efficiency of the Li+ enzyme in the same (PO3-)-transfer step is less than 4 x 10(-8) that of the Mg2+ enzyme.(ABSTRACT TRUNCATED AT 400 WORDS)  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