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.
W J Ray  J W Burgner  C B Post 《Biochemistry》1990,29(11):2770-2778
Near ultraviolet spectral studies were conducted on two inhibitor complexes obtained by treating the dephospho form of the phosphoglucomutase.Mg2+ complex with inorganic vanadate in the presence of either glucose 1-phosphate [cf. Percival, M. D., Doherty, K., & Gresser, M. J. (1990) Biochemistry (first of four papers in this issue)] or glucose 6-phosphate. Part of the spectral differences between the two inhibitor complexes arises because the glucose phosphate moiety in the complex derived from glucose 1-phosphate binds to the enzyme in a different way from the glucose phosphate moiety in the complex derived from glucose 6-phosphate and because these alternative binding modes produce different environmental effects on the aromatic chromophores of the dephospho enzyme. These spectral differences are strikingly similar to those induced by the binding of glucose 1-phosphate and glucose 6-phosphate to the phospho enzyme--which shows that the glucose 1-phosphate and glucose 6-phosphate moieties occupy positions in the inhibitor complexes closely related to those that they occupy in their respective catalytically competent complexes. This binding congruity indicates that in the inhibitor complexes the oxyvanadium grouping is bound at the site where (PO3-) transfer normally occurs. 31P NMR studies of the phosphate group in these complexes also provide support for this binding pattern. A number of other systems based on compounds with altered structures, such as the deoxysugar phosphates, or systems with different compositions, as in the case of the metal-free enzyme or of the glucose phosphates plus nitrate, also were examined for evidence that complexes analogous to the inhibitor complexes were formed, but none was found.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

2.
Phosphoglucomutase is inhibited by a complex formed from alpha-D-glucose 1-phosphate (Glc-1-P) and inorganic vanadate (Vi). Both the inhibition at steady state and the rate of approach to steady state are dependent on the concentrations of both Glc-1-P and Vi. Inhibition is competitive versus alpha-D-glucose 1,6-bisphosphate (Glc-P2) and is ascribed to binding of the 6-vanadate ester of Glc-1-P (V-6-Glc-1-P) to the dephospho form of phosphoglucomutase (E). The inhibition constant for V-6-Glc-1-P at pH 7.4 was determined from steady-state kinetic measurements to be 2 x 10(-12) M. The first-order rate constant for approach to steady state increases hyperbolically with inhibitor concentration. The results are consistent with rapid equilibrium binding of V-6-Glc-1-P to E, with dissociation constant 1 x 10(-9) M, followed by rate-limiting conversion of the E.V-6-Glc-1-P complex to another species, E*.V-6-Glc-1-P, with first-order rate constant 4 x 10(-2)s-1. The rate constant determined for the reverse reaction, conversion of E*.V-6-Glc-1-P to E.V-6-Glc-1-P, is 2.5 x 10(-4)s-1. Formation of E*.V-6-Glc-1-P can also occur via binding of glucose 6-vanadate to the phospho form of phosphoglucomutase (E-P) followed by phosphoryl transfer and rearrangement of the enzyme-product complex.  相似文献   

3.
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   

4.
Inhibition of yeast plasma membrane ATPase by vanadate occurs only if either Mg2+ or MgATP2- is bound to the enzyme. The dissociation constant of the complex of vanadate and inhibitory sites is 0.14-0.20 microM in the presence of optimal concentrations of Mg2+ and of the order of 1 microM if the enzyme is saturated with MgATP2-. The dissociation constants of Mg2+ and MgATP2- for the sites involved are 0.4 and 0.62-0.73 mM, respectively, at pH 7. KCl does not increase the affinity of vanadate to the inhibitory sites as was found with (Na+ + K+)-ATPase. On the other hand, the effect of Mg2+ upon vanadate binding is similar to that upon (Na+ + K+)-ATPase, and the corresponding affinity constants of Mg2+ and vanadate for the two enzymes are of the same order of magnitude.  相似文献   

5.
S Liu  M J Gresser  A S Tracey 《Biochemistry》1992,31(10):2677-2685
The formation of complexes of vanadate with 2-phosphoglycerate and 3-phosphoglycerate have been studied using 51V nuclear magnetic resonance spectroscopy. Signals attributed to two 2,3-diphosphoglycerate analogues, 2-vanadio-3-phosphoglycerate and 2-phospho-3-vanadioglycerate, were detected but were not fully resolved from signals of inorganic vanadate and the anhydride formed between vanadate and the phosphate ester moieties of the individual phosphoglycerates. Equilibrium constants for formation of the two 2,3-bisphosphate analogues were estimated as 2.5 M-1 for 2-vanadio-3-phosphoglycerate and 0.2 M-1 for 2-phospho-3-vanadioglycerate. The results of the binding study are fully consistent with non-cooperativity in the binding of vanadiophosphoglycerate to the two active sites of phosphoglycerate mutase (PGM). 2-Vanadio-3-phosphoglycerate was found to bind to the dephospho form of phosphoglycerate mutase with a dissociation constant of about 1 x 10(-11) M at pH 7 and 7 x 10(-11) M at pH 8. Three signals attributed to histidine residues were observed in the 1H NMR spectrum of phosphoglycerate mutase. Two of these signals and also an additional signal, tentatively attributed to a tryptophan, underwent a chemical shift change when the vanadiophosphoglycerate complex was bound to the enzyme. The results obtained here are in accord with these vanadate-phosphoglycerate complexes being much more potent inhibitors of phosphoglycerate mutase than either monomeric or dimeric vanadate. The dissociation constant of 10(-11) M for 2-vanadio-3-phosphoglycerate is about 4 orders of magnitude smaller than the Km for PGM, a result in accordance with the vanadiophosphoglycerates being transition state analogues for the phosphorylation of PGM by 2,3-diphosphoglycerate.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

6.
Time-dependent 31P saturation-transfer studies were conducted with the Cd2+-activated form of muscle phosphoglucomutase to probe the origin of the 100-fold difference between its catalytic efficiency (in terms of kcat) and that of the more efficient Mg2+-activated enzyme. The present paper describes the equilibrium mixture of phosphoglucomutase and its substrate/product pair when the concentration of the Cd2+ enzyme approaches that of the substrate and how the nine-spin 31P NMR system provided by this mixture was treated. It shows that the presence of abortive complexes is not a significant factor in the reduced activity of the Cd2+ enzyme since the complex of the dephosphoenzyme and glucose 1,6-bisphosphate, which accounts for a large majority of the enzyme present at equilibrium, is catalytically competent. It also shows that rate constants for saturation transfer obtained at three different ratios of enzyme to free substrate are mutually compatible. These constants, which were measured at chemical equilibrium, can be used to provide a quantitative kinetic rationale for the reduced steady-state activity elicited by Cd2+ relative to Mg2+ [cf. Ray, W.J., Post, C.B., & Puvathingal, J.M. (1989) Biochemistry (following paper in this issue)]. They also provide minimal estimates of 350 and 150 s-1 for the rate constants describing (PO3-) transfer from the Cd2+ phosphoenzyme to the 6-position of bound glucose 1-phosphate and to the 1-position of bound glucose 6-phosphate, respectively. These minimal estimates are compared with analogous estimates for the Mg2+ and Li+ forms of the enzyme in the accompanying paper.  相似文献   

7.
Inorganic vanadate (Vi) activates catalysis by glucose-6-phosphate dehydrogenase of the oxidation of glucose by NADP+. As the concentration of Glu-6-P dehydrogenase is increased, the rate of the vanadate-activated glucose oxidation becomes less sensitive to increases in enzyme concentration. The rate of glucose oxidation in the absence of Vi increases linearly with Glu-6-P dehydrogenase concentration. These results are interpreted in terms of nonenzymic formation of glucose 6-vanadate. At high enzyme concentration, vanadate ester formation becomes partially rate-limiting, and extrapolation to infinite Glu-6-P dehydrogenase concentration allows determination of the second order rate constant for formation of the ester. In separate experiments designed to test the proposed mechanism, it was found that Vi, at concentrations at which it strongly activates catalysis by Glu-6-P dehydrogenase of glucose oxidation, has no effect on the rates of oxidation of glucose 6-phosphate or 6-deoxyglucose catalyzed by Glu-6-P dehydrogenase. Sulfate, which is known to activate glucose oxidation and to inhibit glucose 6-phosphate oxidation, strongly activates 6-deoxyglucose oxidation. These experiments show that the 6-hydroxyl group of glucose is essential for the observed activation by Vi and are also consistent with the formation of glucose 6-vanadate. Also, the rate of the sulfate-activated glucose oxidation increases linearly with Glu-6-P dehydrogenase concentration. These results are consistent with the proposed mechanism for sulfate activation which involves sulfate binding to the enzyme (Anderson, W. B., Horne, R. N., and Nordlie, R. C. (1968) Biochemistry 7, 3997-4004). The second order rate constant calculated for formation of glucose 6-vanadate at pH 7.0 is 2.4 M-1 s-1. The corresponding values for glucose 6-phosphate and glucose 6-arsenate formation are approximately 9 X 10(-11) M-1 s-1 and 6.3 X 10(-6) M-1 s-1 (Lagunas, R. (1980) Arch. Biochem. Biophys. 205, 67-75).  相似文献   

8.
The phosphorylation of sarcoplasmic reticulum ATPase with Pi in the absence of Ca2+ was studied by equilibrium and kinetic experimentation. The combination of these measurements was then subjected to analysis without assumptions on the stoichiometry of the reactive sites. The analysis indicates that the species undergoing covalent interaction is the tertiary complex E X Pi X Mg formed by independent interaction of the two ligands with the enzyme. The binding constant of Pi or Mg2+ to either free or partially associated enzyme is approximately equal to 10(2) M-1, and no significant synergistic effect is produced by one ligand on the binding of the other; the equilibrium constant (Keq) for the covalent reaction E X Pi X Mg E-P X Mg is approximately equal to 16, with kphosph = 53 s-1, and khyd = 3-4 s-1 (25 degrees C, pH 6.0, no K+). The phosphorylation reaction of sarcoplasmic reticulum ATPase with Pi is highly H+ dependent. Such a pH dependence involves the affinity of enzyme for different ionization states of Pi, as well as protonation of two protein residues per enzyme unit in order to obtain optimal phosphorylation. The experimental data can then be fitted satisfactorily assuming pK values of 5.7 and 8.5 for the two residues in the nonphosphorylated enzyme (changing to 7.7 for one of the two residues, following phosphorylation) and values of 50.0 and 0.58 for the equilibrium constants of the H2(E X HPO4) in equilibrium with H(E-PO3) + H2O and H(E X HPO4) in equilibrium with E-PO3 + H2O reactions, respectively. In addition to the interdependence of H+ and phosphorylation sites, an interdependence of Ca2+ and phosphorylation sites is revealed by total inhibition of the Pi reaction when two high affinity calcium sites per enzyme unit are occupied by calcium. Conversely, occupancy of the phosphate site by vanadate (a stable transition state analogue of phosphate) inhibits high affinity calcium binding. The known binding competition between the two cations and their opposite effects on the phosphorylation reaction suggest that interdependence of phosphorylation site, H+ sites, and Ca2+ sites is a basic mechanistic feature of enzyme catalysis and cation transport.  相似文献   

9.
The interaction of vanadate ions with the Ca-ATPase from sarcoplasmic reticulum vesicles was studied in a native and a fluorescein-labeled ATPase preparation (Pick, U., and Karlish, S. J. D. (1980) Biochim. Biophys. Acta 626, 255-261). Vanadate induced a fluorescence enhancement in a fluorescein-labeled enzyme, indicating that it shifts the equilibrium between the two conformational states of the enzyme by forming a stable E2-Mg-vanadate complex (E2 is the low affinity Ca2+ binding conformational state of the sarcoplasmic reticulum Ca-ATPase). Indications for tight binding of vanadate to the enzyme (K1/2 = 10 microM) in the absence of Ca2+ and for a slow dissociation of vanadate from the enzyme in the presence of Ca2+ are presented. The enzyme-vanadate complex was identified by the appearance of a time lag in the onset of Ca2+ uptake and by a slowing of the fluorescence quenching response to Ca2+. Ca2+ prevented the binding of vanadate to the enzyme. Pyrophosphate (Kd = 2 mM) and ATP (Kd = 25 microM) competitively inhibited the binding of vanadate, indicating that vanadate binds to the low affinity ATP binding site. Binding of vanadate inhibited the high affinity Ca2+ binding to the enzyme at 4 degrees C. Vanadate also inhibited the phosphorylation reaction by inorganic phosphate (Ki = 10 microM) but had no effect on the phosphorylation by ATP. It is suggested that vanadate binds to a special region in the low affinity ATP binding site which is exposed only in the E2 conformation of the enzyme in the absence of Ca2+ and which controls the rate of the conformation transition in the dephosphorylated enzyme. The implications of these results to the role of the low affinity ATP binding sites are discussed.  相似文献   

10.
Vanadate is able to promote the binding of ouabain to (Na+ +K+)-ATPase and it is shown that vanadate is trapped in the enzyme-ouabain complex. Also ouabain-bound enzyme, the formation of which was facilitated by (Mg2+ +Na+ +ATP) or (Mg2+ +Pi), is accessible to vanadate when washed free of competing ligands used for the promotion of ouabain binding. For vanadate binding to (Na+ +K+)-ATPase and to enzyme-ouabain complexes a divalent cation (Mg2+ or Mn2+) is indispensable, indicating that the cation does not remain attached to the ouabain-bound enzyme. K+ further increases vanadate binding in the absence of ouabain, but seems to have no additional role in case of vanadate binding to enzyme-ouabain complexes. Mn2+ is more efficient than Mg2+ in promoting binding of vanadate and ouabain to (Na+ +K+)-ATPase. That K+ in combination with Mn2+, in analogy with the effect in combination with Mg2+, increases the equilibrium binding level of vanadate and decreases that of ouabain does not seem to favour the hypothesis of selection of a special E2-subconformation by Mn2+. The vanadate-trapped enzyme-ouabain complex was examined for simultaneous nucleotide binding which could demonstrate a two-substrate mechanism per functional unit of the enzyme. The acceleration by (Na+ +ATP) of ouabain release from the (Mg2+ +Pi)-facilitated enzyme-ouabain complex does not, as anticipated, support such a mechanism. On the other hand, the deceleration of vanadate release as well as of ouabain release from a (Mg2+ +vanadate)-promoted complex could be consistent with a two-substrate mechanism working out-of-phase.  相似文献   

11.
Kidney alkaline phosphatase is an enzyme which requires two types of metals for maximal activity: zinc, which is essential, and magnesium, which is stimulatory. The main features of the Mg2+ stimulation have been analyzed. The stimulation is pH-dependent and is observed mainly between pH 7.5 and 10.5. Mg2+ binding to native alkaline phosphatase is characterized by a dissociation constant of 50 muM at pH 8.5,25 degrees. Binding of Zn2+ is an athermic process. Both the rate constants of association, ka, and of dissociation, kd, have low values. Typical values are 7 M(-1) at pH 8.0, 25 degrees, for ka and 4.10(-4) S(-1) at pH 8.0, 25 degrees, for kd. The on and off processes have high activation energies of 29 kcal mol (-1). Mg2+ can be replaced at its specific site by Mn2+, Co2+, Ni2+, and Zn2+. Zinc binding to the Mg2+ site inhibits the native alkaline phosphatase. Mn2+, Co2+, and Ni2+ also bind to the Mg2+ site with a stimulatory effect which is nearly identic-al with that of Mg2+, Mn2+ is the stimulatory cation which binds most tightly to the Mg2+ site; the dissociation constant of the Mn2+ kidney phosphatase complex is 2 muM at pH 8.5. The stoichiometry of Mn2+ binding has been found to be 1 eq of Mn2+ per mol of dimeric kidney phosphatase. The native enzyme displays absolute half-site reactivity for Mn2+ binding. Mg2+ binding site and the substrate binding sites are distinct sites. The Mg2+ stimulation corresponds to an allosteric effect. Mg2+ binding to its specific sites does not affect substrate recognition, it selectively affects Vmax values. Quenching of the phosphoenzyme formed under steady state conditions with [32P]AMP as a substrate as well as stopped flow analysis of the catalyzed hydrolysis of 2,4-dinitrophenyl phosphate or p-nitrophenyl phosphate have shown that the two active sites of the native and of the Mg2+-stimulated enzyme are not equivalent. Stopped flow analysis indicated that one of the two active sites was phosphorylated very rapidly whereas the other one was phosphorylated much more slowly at pH 4.2. Half of the sites were shown to be reactive at pH 8.0. Quenching experiments have shown that only one of the two sites is phosphorylated at any instant; this result was confirmed by the stopped flow observation of a burst of only 1 mol of nitrophenol per mol of dimeric phosphatase in the pre-steady state hydrolysis of p-nitrophenyl phosphate. The half-of-the-sites reactivity observed for the native and for the Mg2+-stimulated enzyme indicates that the same type of complex, the monophosphorylated complex, accumulates under steady state conditions with both types of enzymes. Mg2+ binding to the native enzyme at pH 8.0 increases considerably the dephosphorylation rate of this monophosphorylated intermediate. A possible mechanism of Mg2+ stimulation is discussed.  相似文献   

12.
The interaction of Ca2+ and vanadate with fluorescein isothiocyanate (FITC) labeled sarcoplasmic reticulum (SR) Ca2+-ATPase has been studied by following the kinetics of changes in the reporter group fluorescence and equilibrium fluorescence levels. The vanadate species bound to the enzyme is clearly monomeric orthovanadate, probably H2VO4-. Vanadate binding is noncooperative, suggesting an absence of interactions between the Ca2+-ATPase subunits. The fluorescence experiments confirm the existence of a calcium-enzyme-vanadate complex (in the presence of magnesium). On the basis of the fluorescence properties of this complex, it is similar in its conformation to the calcium-enzyme complex, i.e., "E1-like" rather than "E2-like". However, Ca2+ binds to the enzyme-vanadate complex via sites that are only accessible from the interior of the SR vesicles. The complex Ca2E*Van, which is rapidly formed, isomerizes very slowly (t1/2 approximately 1 min) to the stable ternary complex. The mutual destabilization between bound vanadate and two bound Ca2+ ions is only 1.6 kcal/mol, much smaller than that produced by the interaction of calcium and phosphate.  相似文献   

13.
The effects of vanadate (Na3VO4) on pancreatic B-cell function were studied in normal mouse islets. Vanadate did not affect basal insulin release but potentiated the effect of 7-30 mM glucose at concentrations of 0.1-1 mM. This effect was progressive and slowly reversible. It was abolished by omission of extracellular Ca2+ but unaffected by blockers of adrenergic or muscarinic receptors. Comparison of the changes in membrane potential, 86Rb efflux and 45Ca efflux that vanadate and ouabain produced in B-cells made it possible to exclude the hypothesis that vanadate increases insulin release by blocking the sodium pump. Vanadate was also without effect on cAMP levels. On the other hand, it markedly changed the characteristics of the Ca(2+)-dependent electrical activity and of the oscillations of cytoplasmic Ca2+ recorded in B-cells stimulated by 15 mM glucose. In the steady state, Ca2+ influx was increased by vanadate, and this resulted in a rise in cytoplasmic Ca2+. The exact mechanisms underlying these changes could not be established but a blockade of K channels was excluded. In the presence of LiCl, vanadate markedly increased inositol phosphate levels in islet cells. This effect was attenuated but not suppressed by omission of Ca2+. A small increase in inositol bisphosphate was still produced by vanadate in the absence of LiCl. These results suggest that vanadate both stimulates phosphoinositide breakdown and inhibits inositol phosphate degradation. In conclusion, vanadate does not induce insulin release, but markedly potentiates the stimulation by glucose. This property is not due to an inhibition of the sodium pump or to a rise in cAMP concentration. It results from a complex interplay between changes in B-cell membrane potential, phosphoinositide metabolism and Ca2+ handling.  相似文献   

14.
Deng H  Callender R  Huang Z  Zhang ZY 《Biochemistry》2002,41(18):5865-5872
Vanadate can often bind to phosphoryl transfer enzymes to form a trigonal-bipyramidal structure at the active site. The enzyme-vanadate dissociation constants in these enzymes are much lower than those for phosphate. Therefore, enzyme-bound vanadate moieties are often considered as transition state analogues. To test whether the enzyme-vanadate complex is a true transition state analogue beyond the simple geometry and binding affinity arguments and whether the bond orders of the VO bonds in the complex approach those of the PO bonds in the transition state, the binding properties of vanadate in the Yersinia protein-tyrosine phosphatase (PTPase) and its T410A, D356N, W354A, R409K, and D356A mutants have been studied by steady-state kinetic measurements and by difference Raman measurements. The results of the kinetic measurements show no correlation between K(I) and kcat or kcat/K(m) in these mutants. In addition, our analysis of the Raman data shows that the bond order change of the nonbridging V--O bonds in the vanadate complexes does not correlate with the kinetic parameters in a number of PTPase variants as predicted by the transition state binding paradigm. Furthermore, the ionization state of the bound vanadate moiety is not invariant across the PTPase variants studied, and the average bond order of the nonbridging V--O bonds decreased by 0.06-0.07 valence unit in the wild type and all of the mutant PTPases, either in dianionic or in monoanionic form. Thus the complex would resemble an associative transition state, contrary to the previously determined dissociative structure of the transition state. Therefore, it is concluded that vanadate is not a true transition state analogue for the PTPase reactions.  相似文献   

15.
The structure of vanadate, a phosphate analogue which was suggested to function in the presence of tightly bound ADP and divalent cations as a transition state inhibitor of CF1-ATPase, was investigated by X-ray absorption spectroscopy. Analysis of the vanadium K-edge was used for determination of the structure of vanadate bound to a single site in CF1-ATPase containing a single tightly bound ADP. There was a decrease in the intensity of the 1s-3d pre-edge transition and a change in the shape of two other shoulders at the edge region upon binding of vanadate to CF1 in the presence of Mg2+ ions. The changes are due to alteration in the structure of vanadium from tetrahedral to a five-coordinated trigonal bipyramidal geometry. Comparison of the pre-edge peak intensity of ADP-vanadate complex, and model compound resolved by crystallography support the proposed structure of CF1-bound vanadate. 51V NMR measurements were used to verify the pentacoordinated structure of ADP-vanadate complex used as a model in the X-ray absorption studies. The inhibition of a single and multiple site activity by vanadate and by MgADP was measured. Vanadate inhibition of CF1-ATPase activity decreased more than 90 fold in the presence of MgADP. A differential specificity of the inhibition in single and multiple mode of activity was observed. It is suggested that ADP-vanadate binds to the active sites of the enzyme as a pentacoordinated vanadium having approximate trigonal bipyramidal geometry. This structure is analogous to the proposed transition state of the phosphate during the synthesis and the hydrolysis of ATP by CF1.  相似文献   

16.
The binding and conformational properties of the divalent cation site required for H+,K(+)-ATPase catalysis have been explored by using Ca2+ as a substitute for Mg2+. 45Ca2+ binding was measured with either a filtration assay or by passage over Dowex cation exchange columns on ice. In the absence of ATP, Ca2+ was bound in a saturating fashion with a stoichiometry of 0.9 mol of Ca2+ per active site and an apparent Kd for free Ca2+ of 332 +/- 39 microM. At ATP concentrations sufficient for maximal phosphorylation (10 microM), 1.2 mol of Ca2+ was bound per active site with an apparent Kd for free Ca2+ of 110 +/- 22 microM. At ATP concentrations greater than or equal to 100 microM, 2.2 mol of Ca2+ were bound per active site, suggesting that an additional mole of Ca2+ bound in association with low affinity nucleotide binding. At concentrations sufficient for maximal phosphorylation by ATP (less than or equal to 10 microM), APD, ADP + Pi, beta,gamma-methylene-ATP, CTP, and GTP were unable to substitute for ATP. Active site ligands such as acetyl phosphate, phosphate, and p-nitrophenyl phosphate were also ineffective at increasing the Ca2+ affinity. However, vanadate, a transition state analog of the phosphoenzyme, gave a binding capacity of 1.0 mol/active site and the apparent Kd for free Ca2+ was less than or equal to 18 microM. Mg2+ displaced bound Ca2+ in the absence and presence of ATP but Ca2+ was bound about 10-20 times more tightly than Mg2+. The free Mg2+ affinity, like Ca2+, increased in the presence of ATP. Monovalent cations had no effect on Ca2+ binding in the absence of ATP but dit reduce Ca2+ binding in the presence of ATP (K+ = Rb+ = NH4 + greater than Na+ greater than Li+ greater than Cs+ greater than TMA+, where TMA is tetramethylammonium chloride) by reducing phosphorylation. These results indicate that the Ca2+ and Mg2+ bound more tightly to the phosphoenzyme conformation. Eosin fluorescence changes showed that both Ca2+ and Mg2+ stabilized E1 conformations (i.e. cytosolic conformations of the monovalent cation site(s)) (Ca.E1 and Mg.E1). Addition of the substrate acetyl phosphate to either Ca.E1 or Mg.E1 produced identical eosin fluorescence showing that Ca2+ and Mg2+ gave similar E2 (extracytosolic) conformations at the eosin (nucleotide) site. In the presence of acetyl phosphate and K+, the conformations with Ca2+ or Mg2+ were also similar. Comparison of the kinetics of the phosphoenzyme and Ca2+ binding showed that Ca2+ bound prior to phosphorylation and dissociated after dephosphorylation.(ABSTRACT TRUNCATED AT 400 WORDS)  相似文献   

17.
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   

18.
In the presence of Mg2+ vanadate was shown to facilitate ouabain binding to (Na+ + K+)-ATPase in much the same way as Pi does. Thus the hypothesis that vanadate interacts with the phosphate site of the enzyme seems to be supported by ouabain binding experiments. At given ouabain concentrations maximum binding is achieved at microM concentrations of vanadate whereas mM concentrations of Pi are needed. Na+ as well as K+ counteract ouabain binding but some cardiac glycoside binding is still possible at in vivo concentrations of these cations. A minor contamination of the enzyme preparations with vanadate could explain the in vitro binding of ouabain that can be obtained with Mg2+ and in the absence of Pi.  相似文献   

19.
Ribonuclease T1 (RNase-T1) from Aspergillus Oryzae cleaves ribonucleic acid specifically at guanosine to yield oligonucleotides with terminal guanosine-3'-phosphate. It forms a complex with vanadate (association constant K approximately 145 +/- 30 M-1; delta (51V) = -514 ppm) with spectral features similar to the less stable complexes obtained with di- and tripeptides (Gly-His, Pros-His-Ala, Gly-His-Lys, Val-Glu) containing amino acids that are constituents at the active site of the enzyme. Guanosine also forms a (sparingly soluble) complex with vanadate. Its role is mimicked by inosine, which yields two soluble complexes with vanadate, characterized by delta values of -511 (K = 94 M-1) and -523 ppm (K = 305 M-1 in TRIS buffer and 685 m-1 in buffer-free solution). Comparison with literature values leads to an assignment of the delta = -523 signal to a complex where monovanadate, possibly in a trigonal bipyramidal geometry suggested for the transition state of the phosphate analogue, is coordinated to the 2'- and 3'-oxygens of the ribose ring. A drastic increase of complex stability is observed in the ternary vanadate (12-16 mM)/inosine(10.5 mM)/RNase-T1(5.4 mM) system. An approximate lower limit for the association constant is 1.5.10(5) M-2. The spectral characteristics of the main component of the binary vanadate/inosine complex are essentially maintained (delta = -525 ppm, half-width = 960 Hz), suggesting vanadate binding to the enzyme through hydrogen bonds.  相似文献   

20.
The activation kinetics of purified Rhodospirillum rubrum ribulose bisphosphate carboxylase were analysed. The equilibrium constant for activation by CO(2) was 600 micron and that for activation by Mg2+ was 90 micron, and the second-order activation constant for the reaction of CO(2) with inactive enzyme (k+1) was 0.25 X 10(-3)min-1 . micron-1. The latter value was considerably lower than the k+1 for higher-plant enzyme (7 X 10(-3)-10 X 10(-3)min-1 . micron-1). 6-Phosphogluconate had little effect on the active enzyme, and increased the extent of activation of inactive enzyme. Ribulose bisphosphate also increased the extent of activation and did not inhibit the rate of activation. This effect might have been mediated through a reaction product, 2-phosphoglycolic acid, which also stimulated the extent of activation of the enzyme. The active enzyme had a Km (CO2) of 300 micron-CO2, a Km (ribulose bisphosphate) of 11--18 micron-ribulose bisphosphate and a Vmax. of up to 3 mumol/min per mg of protein. These data are discussed in relation to the proposed model for activation and catalysis of ribulose bisphosphate carboxylase.  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