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.
A hydrophobic amine, SCH 28080, 2-methyl-8-(phenylmethoxy)imidazo(1,2a)pyridine-3-acetonitrile, previously shown to inhibit gastric acid secretion in vivo and in vitro, was also shown to inhibit basal and stimulated aminopyrine accumulation in isolated gastric glands when histamine, high K+ concentrations, or dibutyryl cAMP were used as secretagogues. Stimulated, but not basal, oxygen consumption was also inhibited. Neutralization of the acid space of the parietal cell by high concentrations of the weak base, imidazole, reduced the potency of the drug, suggesting that SCH 28080 was active when protonated. Studies on the isolated H+,K+-ATPase showed that the compound inhibited the enzyme competitively with K+, whether ATP or p-nitrophenyl phosphate were used as substrates. In contrast, the inhibition was mixed with respect to p-nitrophenyl phosphate and uncompetitive with respect to ATP. The drug reduced the steady state level of the phosphoenzyme but not the observed rate constant for phosphoenzyme formation in the absence of K+ nor the quantity of phosphoenzyme reacting with K+. The drug quenched the fluorescence of fluorescein isothiocyanate-modified enzyme and also inhibited the ATP-independent K+ exchange reaction of the H+,K+-ATPase. Its action on gastric acid secretion can be explained by inhibition of the H+,K+-ATPase by reversible complexation of the enzyme. This class of compound, therefore, acts as a reversible inhibitor of gastric acid secretion.  相似文献   

2.
Acylphosphatase, one of the smallest enzymes, is expressed in all organisms. It displays hydrolytic activity on acyl phosphates, nucleoside di- and triphosphates, aryl phosphate monoesters, and polynucleotides, with acyl phosphates being the most specific substrates in vitro. The mechanism of catalysis for human acylphosphatase (the organ-common type isoenzyme) was investigated using both aryl phosphate monoesters and acyl phosphates as substrates. The enzyme is able to catalyze phosphotransfer from p-nitrophenyl phosphate to glycerol (but not from benzoyl phosphate to glycerol), as well as the inorganic phosphate-H(2)18O oxygen exchange reaction in the absence of carboxylic acids or phenols. In short, our findings point to two different catalytic pathways for aryl phosphate monoesters and acyl phosphates. In particular, in the aryl phosphate monoester hydrolysis pathway, an enzyme-phosphate covalent intermediate is formed, whereas the hydrolysis of acyl phosphates seems a more simple process in which the Michaelis complex is attacked directly by a water molecule generating the reaction products. The formation of an enzyme-phosphate covalent complex is consistent with the experiments of isotope exchange and transphosphorylation from substrates to glycerol, as well as with the measurements of the Br?nsted free energy relationships using a panel of aryl phosphates with different structures. His-25 involvement in the formation of the enzyme-phosphate covalent complex during the hydrolysis of aryl phosphate monoesters finds significant confirmation in experiments performed with the H25Q mutated enzyme.  相似文献   

3.
The complete time course of the hydrolysis of p-nitrophenyl phosphate catalyzed by the low molecular weight (acid) phosphotyrosyl protein phosphatase from bovine heart was elucidated and analyzed in detail. Burst titration kinetics were demonstrated for the first time with this class of enzyme. At pH 7.0, 4.5 degrees C, a transient pre-steady-state "burst" of p-nitrophenol was formed with a rate constant of 48 s-1. The burst was effectively stoichiometric and corresponded to a single enzyme active site/molecule. The burst was followed by a slow steady-state turnover of the phosphoenzyme intermediate with a rate constant of 1.2 s-1. Product inhibition studies indicated an ordered uni-bi kinetic scheme for the hydrolysis. Partition experiments conducted for several substrates revealed a constant product ratio. Vmax was constant for these substrates, and the overall rate of hydrolysis was increased greatly in the presence of alcohol acceptors. An enzyme-catalyzed 18O exchange between inorganic phosphate and water was detected and occurred with kcat = 4.47 x 10(-3) s-1 at pH 5.0, 37 degrees C. These results were all consistent with the existence of a phosphoenzyme intermediate in the catalytic pathway and with the breakdown of the intermediate being the rate-limiting step. The true Michaelis binding constant Ks = 6.0 mM, the apparent Km = 0.38 mM, and the rate constants for phosphorylation (k2 = 540 s-1) and dephosphorylation (k3 = 36.5 s-1) were determined under steady-state conditions with p-nitrophenyl phosphate at pH 5.0 and 37 degrees C in the presence of phosphate acceptors. The energies of activation for the enzyme-catalyzed hydrolysis at pH 5.0 and 7.0 were 13.6 and 14.1 kcal/mol, respectively. The activation energy for the enzyme-catalyzed medium 18O exchange between phosphate and water was 20.2 kcal/mol. Using the available equilibrium and rate constants, an energetic diagram was constructed for the enzyme-catalyzed reaction.  相似文献   

4.
N10-Formyltetrahydrofolate synthetase from bacteria and yeast catalyzes a slow formate-dependent ADP formation in the absence of H4folate. The synthesis of formyl phosphate by the enzyme was detected by trapping the intermediate as formyl hydroxamate. That the "formate kinase" activity was part of the catalytic center of N10-formyltetrahydrofolate synthetase was shown by demonstrating coordinate inactivation of the "kinase" and synthetase activities by heat and a sulfhydryl reagent, similar effects of monovalent cations, similar Km values for substrates, and similar Ki values for the inhibitor phosphonoacetaldehyde for both activities. The relative rates of the kinase activities for the bacterial and yeast enzymes are about 10(-4) and 4 x 10(-6) of their respective synthetase activities. These slow rates for the kinase reaction can be explained by the slow dissociation of ADP and formyl phosphate from the enzyme. This conclusion is supported by rapid-quench studies where a "burst" of ADP formation (6.4 s-1) was observed that is considerably faster than the steady-state rate (0.024 s-1). The demonstration of enzyme-bound products by a micropartition assay and the lack of a significant formate-stimulated exchange between ADP and ATP provide further evidence for the slow release of the products from the enzyme. The synthesis of N10-CHO-H4folate when H4folate was added to the E-formyl phosphate-ADP complex is also characterized by a "burst" of product formation. The rate of this burst phase at 5 degrees C occurs with a rate constant of 18 s-1 compared to 14 s-1 for the overall reaction at the same temperature. These results provide further evidence for formyl phosphate as an intermediate in the reaction and are consistent with the sequential mechanism of the normal catalytic pathway. Positional isotope exchange experiments using [beta,gamma-18O]ATP showed no evidence for exchange during turnover experiments in the presence of either H4folate or the competitive inhibitor pteroyltriglutamate. The absence of scrambling of the 18O label as observed by 31P NMR suggests that the central complex may impose restraints to limit free rotation of the P beta oxygens of the product ADP.  相似文献   

5.
1. The properties and kinetics of the 2,3-diphosphoglycerate-dependent phosphoglycerate mutases are discussed. There are at least three possible mechanisms for the reaction: (i) a phosphoenzyme (Ping Pong) mechanism; (ii) an intermolecular transfer of phosphate from 2,3-diphosphoglycerate to the substrates (sequential mechanism); (iii) an intramolecular transfer of phosphate. It is concluded that these mechanisms cannot be distinguished by conventional kinetic measurements. 2. The fluxes for the different mechanisms are calculated and it is shown that it should be possible to distinguish between the mechanisms by appropriate induced-transport tests and by comparing the fluxes of (32)P- and (14)C-labelled substrates at chemical equilibrium. 3. With (14)C-labelled substrates no induced transport was found over a wide concentration range, and with (32)P-labelled substrates co-transport occurred that was independent of concentration over a twofold range. (14)C-labelled substrates exchange at twice the rate of (32)P-labelled substrates at chemical equilibrium. The results were completely in accord with a phosphoenzyme mechanism and indicated a rate constant for the isomerization of the phosphoenzyme of not less than 4x10(6)s(-1). The intramolecular transfer of phosphate (and intermolecular transfer between two or more molecules of substrate) were completely excluded. The intermolecular transfer of phosphate from 2,3-diphosphoglycerate would have been compatible with the results only if the K(m) for 2-phosphoglycerate had been over 7.5-fold smaller than the observed value and if an isomerization of the enzyme-2,3-diphosphoglycerate complex had been the major rate-limiting step in the reaction. 4. The very rapid isomerization of the phosphoenzyme that the experiments demonstrate suggests a mechanism that does not involve a formal isomerization. According to this new scheme the enzyme is closely related mechanistically and perhaps evolutionarily to a 2,3-diphosphoglycerate diphosphatase.  相似文献   

6.
The intrinsic fluorescence of sarcoplasmic reticulum vesicles was measured under conditions allowing ATPase phosphorylation from inorganic phosphate. Significant fluorescence enhancement of up to 4% resulted from gradient-independent enzyme phosphorylation at pH 6, in the absence of KCl. The equilibrium fluorescence data obtained at various magnesium and phosphate concentrations agree with a reaction scheme in which Mg2+, as direct activator, and free phosphate, as the true substrate, bind to the enzyme in random order to give a noncovalent ternary complex (Mg.*E.Pi), in equilibrium with the covalent phosphoenzyme (Mg.*E-P). The transient kinetics of the fluorescence rise was also studied, and the resulting data were generally consistent with the above scheme, assuming that binding reactions are fast compared to covalent phosphoenzyme formation. This, however, might be valid only as a first approximation. At 20 degrees C and pH 6, the phosphate concentration for half-maximum phosphorylation rate constant, at 20 mM magnesium, was higher than 20 mM. Similarly, the magnesium concentration for half-maximum phosphorylation rate constant, at 20 mM phosphate, was also higher than 20 mM. The maximum phosphorylation rate was faster than 25 s-1, and the phosphoenzyme hydrolysis rate constant was 1.5-2 s-1 under these conditions, so that the equilibrium constant between Mg.*E.Pi and Mg.*E-P largely favors the phosphoenzyme.  相似文献   

7.
We have developed two methods for quantitatively measuring inorganic pyrophosphate (PPi) in the presence of 10(3)--10(4) molar excesses of inorganic phosphate (Pi) and used them to measure the extent of enzyme-bound pyrophosphate (EPPi) formation in solutions of yeast inorganic pyrophosphatase and Pi. We have also measured the rate of enzyme-catalyzed H2O--phosphate oxygen exchange. We find both processes to have essentially identical dependence on Mg2+ and Pi concentrations, thus providing important confirmation for the recent proposal by Janson et al. (1979) that oxygen exchange proceeds via EPPi formation. Our results are consistent with a model in which three Mg2+ per active site are required for EPPi formation but inconsistent with a model requiring only two Mg2+ per active site and permit the formulation of an overall scheme for inorganic pyrophosphatase catalysis of PPi--Pi equilibration as well as the evaluation of equilibrium and rate constants in this scheme. The major results and conclusions of our work are the following: (a) the equilibrium constant for PPi (enzyme-bound) in equilibrium with 2Pi (enzyme-bound) is 4.8; (b) following PPi hydrolysis, the first released Pi contains an oxygen from solvent water; (c) the steps for PPi hydrolysis on the enzyme and for release of both product Pi's are all partially rate determining in overall enzyme-catalyzed PPi hydrolysis; (d) PPi formation on the enzyme is rate determining for H2O--Pi oxygen exchange; (e) PPi dissociation from the enzyme is very slow and is the rate-determining step in Pi--PPi exchange (Cohn, 1958; Janson et al., 1979). This also accounts for the observation that the calculated dissociation constant for MgPPi complex binding to enzyme is considerably lower than the measured Km for enzyme-catalyzed MgPPi hydrolysis.  相似文献   

8.
The stereochemical course of the phospho group transfer catalyzed by mutant (S102C) alkaline phosphatase from Escherichia coli was investigated by using 31P nuclear magnetic resonance spectroscopy. Transphosphorylation from 4-nitrophenyl (Rp)-[16O, 17O, 18O]phosphate to (S)-propane-1,2-diol occurs with overall retention of configuration at phosphorus. This result is consistent with the view that the hydrolysis of substrates by this mutant enzyme proceeds by way of a covalent phosphoenzyme intermediate in the same manner as the wild-type alkaline phosphatase.  相似文献   

9.
Thiamin phosphate synthase catalyzes the formation of thiamin phosphate from 4-amino-5-(hydroxymethyl)-2-methylpyrimidine pyrophosphate and 5-(hydroxyethyl)-4-methylthiazole phosphate. Several lines of evidence suggest that the reaction proceeds via a dissociative mechanism. The previously determined crystal structure of thiamin phosphate synthase in complex with the reaction products, thiamin phosphate and magnesium pyrophosphate, provided a view of the active site and suggested a number of additional experiments. We report here seven new crystal structures primarily involving crystals of S130A thiamin phosphate synthase soaked in solutions containing substrates or products. We prepared S130A thiamin phosphate synthase with the intent of characterizing the enzyme-substrate complex. Surprisingly, in three thiamin phosphate synthase structures, the active site density cannot be modeled as either substrates or products. For these structures, the best fit to the electron density is provided by a model that consists of independent pyrimidine, pyrophosphate, and thiazole phosphate fragments, consistent with a carbenium ion intermediate. The resulting carbenium ion is likely to be further stabilized by proton transfer from the pyrimidine amino group to the pyrophosphate to give the pyrimidine iminemethide, which we believe is the species that is observed in the crystal structures.  相似文献   

10.
A general method to study the phosphate group of phosphoenzymes with infrared difference spectroscopy by helper enzyme-induced isotope exchange was developed. This allows the selective monitoring of the phosphate P-O vibrations in large proteins, which provides detailed information on several band parameters. Here, isotopic exchange was achieved at the oxygen atoms of the catalytically important phosphate group that transiently binds to the sarcoplasmic reticulum Ca(2+)-ATPase (SERCA1a). [gamma-(18)O(3)]ATP phosphorylated the ATPase, which produced phosphoenzyme that was initially isotopically labeled. The helper enzyme adenylate kinase regenerated the substrate ATP from ADP (added or generated upon ATP hydrolysis) with different isotopic composition than used initially. With time this produced the unlabeled phosphoenzyme. The method was tested on the ADP-insensitive phosphoenzyme state of the Ca(2+)-ATPase for which the vibrational frequencies of the phosphate group are known, and it was established that the helper enzyme is effective in mediating the isotope exchange process.  相似文献   

11.
The fatty acid of acyl dihydroxyacetone phosphate can be exchanged enzymatically for another fatty acid. It has been shown that this reaction proceeds by cleavage of the oxygen bound to C-1 of the dihydroxyacetone phosphate (DHAP) moiety rather than by the more common cleavage at the acyl to oxygen bond. In the present study, the stereochemistry of this reaction was defined further; using deuterated substrates and fast atom bombardment-mass spectrometry, it was shown that the fatty acid exchange involves the stereospecific labilization of the pro-R hydrogen at C-1 of the DHAP moiety of acyl DHAP. The mechanism of ether bond formation, in which acyl DHAP is converted to O-alkyl DHAP, also proceeds via labilization of the pro-R hydrogen and cleavage of the fatty acid at the C-1 to oxygen bond. In addition, other workers have provided evidence that the enzyme responsible for the exchange reaction is O-alkyl DHAP synthetase. Therefore, the present results support the hypothesis that the acyl exchange is the reverse reaction of the first step in O-alkyl DHAP synthesis; in both of these reactions the pro-R hydrogen of C-1 of the DHAP moiety of acyl DHAP and the fatty acid moiety are labilized with cleavage of the fatty acid at the DHAP C-1 to oxygen bond.  相似文献   

12.
The steady state rate of ATP hydrolysis (v) by the gastric H,K-ATPase and the steady state level of phosphoenzyme (E-P) have been measured at 0 and 10 mM KCl; both v and E-P have a nonhyperbolic dependence on the ATP concentration that is consistent with negative cooperativity. The ratio of the rate of hydrolysis to phosphoenzyme (v/[E-P]) was found to vary with the concentration of ATP. Thus, for the rate law v = [E-P].k, k must be a function of the ATP concentration. This requires that ATP be able to bind to E-P or to an enzyme form that occurs after E-P but prior to an irreversible step, such as the loss of inorganic phosphate (Pi). At low ATP concentrations, product inhibition by Pi gives concave downward plots of 1/v against Pi concentration. Pi increases the apparent Km and decreases the apparent Vm. At saturating ATP concentrations, Pi is a noncompetitive inhibitor. These data show that ATP and Pi can bind to the H,K-ATPase simultaneously. They are inconsistent with mechanisms where the binding of ATP and Pi is mutually exclusive.  相似文献   

13.
Native solium and potassium adenosine triphosphatase from guinea pig kidney accepted a phosphate group from radioactive inorganic phosphate to form an acyl phosphate bond at the active site in the presence or absence of sodium ion. Magnesium ion was always required. In the presence of sodium ion and absence of adenosine triphosphate, there was no phosphorylation by inorganic phosphate. Addition of unlabeled adenosine triphosphate produced a potassium-sensitive phosphoenzyme which exchanged its phosphate-group with radioactive inorganic phosphate. The dephosphoenzyme was an intermediate in this exchange. The rate constant for dephosphorylation was about 0.05 per second. Addition of rubidium ion, a congener of potassium ion, to the potassium-sensitive phosphoenzyme produced a phosphoenzyme labeled from inorganic phosphate with a corresponding rate constant of 0.26 per s. This was a rubidium-complexed phosphoenzyme. Addition of magnesium ion to potassium-sensitive phosphoenzyme converted it into insensitive phosphoenzyme, the splitting of which was not accelerated by potassium ion or by adenosine diphosphate. Its rate constant was 0.07 per s. In the absence of sodium ion and adenosine triphosphate, inorganic phosphate was incorporated directly into a similar insensitive phosphoenzyme. In the presence of potassium ion or rubidium ion, inorganic phosphate was incorporated into a potassium-complexed or rubidium-complexed phosphoenzyme which exchanged 32-P with inorganic phosphate completely in less than 3 s. Incorporation of inorganic phosphate into a complex of the enzyme with the inhibitor, ouabain, is already described in the literature. Its rate constant was about 0.02 per s. Thus there appear to be at least four reactive states of the phosphoenzyme which equilibrate measurably with inorganic phosphate, namely, potassium-sensitive phosphoenzyme, potassium-complexed phosphoenzyme, insensitive phosphoenzyme, and ouabain phosphoenzyme. Two of these reactive states are functional intermediates in native sodium and potassium ion transport adenosine triphosphatase. The results are compatible with control of the reactivity of the active site by conformational changes in the surrounding active center and with regulation of the energy level of the phosphate group according to the kind of monovalent cation bound to the enzyme.  相似文献   

14.
Initial velocity and isotope exchange studies confirmed that the over-all reaction, like that catalyzed by pyruvate carboxylase purified from rat liver and chicken liver, was a nonclassical Ping Pong Bi Bi Uni Uni sequence with ATP and HCO3-binding randomly in the Bi Bi partial reaction. Three possible mechanisms for the coupling of ATP hydrolysis and CO2 fixation are considered: (i) Mechanism i, a concerted mechanism without the formation of a kinetically significant or detectable intermediate; (ii) Mechanism ii, activation of the enzyme by ATP to form an activated phosphoenzyme complex which can react with HCO3- by formation of a phosphorylated intermediate. On the basis of other evidence, an activated intermediate containing the ADP moiety was considered improbable. Evidence is presented which indicates that an isotopic exchange between ATP and ADP in the absence of added orthophosphate is not a property of the sheep kidney enzyme. This observation removed the need to postulate either that this exchange is an abortive reaction, or that there is a compulsory formation of a phosphoenzyme intermediate. Two analogues of ADP, alpha,beta-methylene adenosine diphosphate, and adenosine 5'-phosphosulfate, have been used to provide further evidence against Mechanism ii. Both compounds were competitive inhibitors with respect to MgATP2- (Ki values respectively, 0.58 mM and 3.0 mM, compared with 0.17 mM for ADP), but neither could be phosphorylated by the enzyme. Neither analogue could replace ADP in the HCO3-: oxalacetate isotopic exchange reaction, indicating that phosphorylation of ADP is necessary for this exchange to occur, and that Mechanism ii is not applicable. Since Mechanism iii involves formation of a carbonly phosphate intermediate, analogues of this compound, namely, carbamyl phosphate and phosphonacetic acid were used to examine this pathway. The fact that the enzyme catalyzed the synthesis of ATP from ADP and carbamyl phosphate, and that phosphonacetic acid was a noncompetitive inhibitor with respect to MgATP2- (Ki = 0.5 mM) provides strong evidence that a carbonyl phosphate derivative is involved in the reaction mechanism. However, we have not found from initial velocity studies evidence for the formation of this intermediate, and it may therefore have only a transient existence in an essentially concerted reaction.  相似文献   

15.
1. The kinetics of phosphoglucomutases from different sources are discussed and it is concluded that on the available evidence there are in all cases three possible mechanisms for the reaction. These are an indirect transfer of phosphate involving the phosphoenzyme (mechanism 1), a direct transfer of phosphate (mechanism 2), and an intermolecular transfer of phosphate from glucose 1,6-diphosphate to the substrate (mechanism 3). Conventional net flux measurements are shown not to differentiate between these mechanisms. 2. Flux equations are developed and it is shown that there are three flux ratios that characterize and distinguish between the mechanisms. 3. To examine these flux ratios induced-transport tests are described with 14C- and 32P-labelled substrates. The fluxes determined with 14C- and 32P-labelled substrates are also compared at chemical equilibrium. 4. With rabbit muscle phosphoglucomutase the results of these tests were completely consistent with mechanism 1 and unequivocally excluded any substantial part of the reaction proceeding by mechanism 2 or mechanism 3. Evidence was also obtained for an isomerization of the phosphoenzyme with an apparent rate constant about 4·5×107sec.−1. Taking into account the activity coefficients of the substrates the true rate constant appears to be about one-sixth of this value. 5. Isotope effects and non-ideal behaviour of the solutions are discussed and the activity coefficients of the substrates are shown to be equal by measurement of the depression of freezing point. It is concluded that these factors do not influence the tests significantly. 6. Alternative mechanisms are considered and it is concluded that the tests show that the glucose residue is transferred directly, that the phosphate is transferred indirectly with one intermediate phosphate, and that there is an isomerization of the free phosphoenzyme without reference to any other details of the reaction. Further, no assumptions are required about the constancy of rate constants. 7. The relative merits of induced transport and product inhibition for detecting isomerization of the enzyme are discussed. It is concluded that the induced-transport test is more sensitive and that its interpretation is less equivocal. 8. The application of the tests to other enzyme systems is briefly considered.  相似文献   

16.
Z Y Zhang  R L Van Etten 《Biochemistry》1991,30(37):8954-8959
The kcat and Km values for the bovine heart low molecular weight phosphotyrosyl protein phosphatase catalyzed hydrolysis of 16 aryl phosphate monoesters and of five alkyl phosphate monoesters having the structure Ar(CH2)nOPO3H2 (n = 1-5) were measured at pH 5.0 and 37 degrees C. With the exception of alpha-naphthyl phosphate and 2-chlorophenyl phosphate, which are subject to steric effects, the values of kcat are effectively constant for the aryl phosphate monoesters. This is consistent with the catalysis being nucleophilic in nature, with the existence of a common covalent phosphoenzyme intermediate, and with the breakdown of this intermediate being rate-limiting. In contrast, kcat for the alkyl phosphate monoesters is much smaller and the rate-limiting step for these substrates is interpreted to be the phosphorylation of the enzyme. A single linear correlation is observed for a plot of log (kcat/Km) vs leaving group pKa for both classes of substrates at pH 5.0: log (kcat/Km) = -0.28pKa + 6.88 (n = 19, r = 0.89), indicating a uniform catalytic mechanism for the phosphorylation event. The small change in effective charge (-0.28) on the departing oxygen of the substrate is similar to that observed in the specific acid catalyzed hydrolysis of monophosphate monoanions (-0.27) and is consistent with a strong electrophilic interaction of the enzyme with this oxygen atom in the transition state. The D2O solvent isotope effect and proton inventory experiments indicate that only one proton is "in flight" in the transition state of the phosphorylation process and that this proton transfer is responsible for the reduction of effective charge on the leaving oxygen.  相似文献   

17.
Evidence is presented based on requirements for exchange in the partial reactions, initial velocity and exchange kinetics and product inhibition, that the pyruvate, phosphate dikinase reaction of propionibacteria occurs by a nonclassical Tri Uni Uni Ping Pong mechanism. The mechanism involves a pyrophosphoryl enzyme, a phosphoryl enzyme, and the free enzyme, and three functionally distinct and independent substrate sites. On the first site, there is pyrophosphorylation of the enzyme by ATP with subsequent release of AMP. The pyrophosphoryl moiety then reacts at the second site with Pi yielding the product PPi and the phosphoryl from of the enzyme. At the third site pyruvate is phosphorylated yielding P-enolpyruvate and the free enzyme. The three catalytic sites are proposed to be linked by a histidyl residue which functions as a pyrophosphoyrl- and phosphoryl-carrier between the three sites. This proposal is based on the following observations. (A) The patterns of the double reciprocal plots of the initial velocities were all parallel; (b) product inhibition between each pair of substrates and products of the three partial reactions were competitive, i.e. ATP against AMP, Pi against PPi, and pyruvate against P-enolpyruvate; (c) the other product inhibitions, with one exception, were noncompetitive as required by the nonclassical ping-pong mechanism; (d) ATP or P-enolpyruvate was required for the Pi in equilibrium PPi exchange reaction which is in accord with the participation of a pyrosphosphoryl or phosphoryl form of the enzyme in this exchange; (e) the ATP in equilibrium AMP exchange and pyruvate in equilibrium P-enolpyruvate exchange did not require additional substrates. In addition, the inhibition and participation in the exchange reactions of the alpha,beta and beta,gamma-methylene analogues of ATP and of the methylene analogue of inorganic pyrophosphate were investigated and the results were in accord with the proposed mechanism. The combined evidence provides a well documented example of a three site nonclassical Tri Uni Uni Ping Pong mechanism.  相似文献   

18.
Steady state initial velocity studies were carried out to determine the kinetic mechanism of human liver aldehyde dehydrogenase. Intersecting double reciprocal plots obtained in the absence of inhibitors demonstrated that the dehydrogenase reaction proceeded by sequential addition of both substrates prior to release of products. Dead end inhibition patterns obtained with coenzyme and substrate analogues (e.g. thionicotinamide-AD+ and chloral hydrate) indicated that NAD+ and aldehyde can bind in random fashion. The patterns of inhibition by the product NADH and of substrate inhibition by glyceraldehyde were also consistent with this mechanism. However, comparisons between kinetic constants associated with the dehydrogenase and esterase activities of this enzyme suggested that most of the dehydrogenase reaction flux proceeds via formation of an initial binary NAD+-enzyme complex over a wide range of substrate and coenzyme concentrations.  相似文献   

19.
p-Hydroxybenzoate hydroxylase (EC 1.14.13.2) from Pseudomonas fluorescens is a NADPH-dependent, FAD-containing monooxygenase catalyzing the hydroxylation of p-hydroxybenzoate to form 3,4-dihydroxybenzoate in the presence of NADPH and molecular oxygen. The mechanism of this three-substrate reaction was investigated in detail at pH 6.6, 4 degrees C, by steady state kinetics, stopped flow spectrophotometry, and equilibrium binding experiments. The initial velocity patterns are consistent with a ping-pong type mechanism which involves two ternary complexes between the enzyme and substrates. The first ternary complex is formed by random addition of p-hydroxybenzoate and NADPH to the enzyme, followed by the release of the first product (NADP+). The reduced enzyme . p-hydroxybenzoate complex now reacts with oxygen, the third substrate, to form the second ternary complex. The enzyme-bound p-hydroxybenzoate then reacts with the activated oxygen to give 3,4-dihydroxybenzoate which is released regenerating the oxidized enzyme for the next cycle. The binding of p-hydroxybenzoate to the oxidized enzyme to form a 1:1 complex causes large, characteristic spectral perturbations and fluorescence quenching. The dissociation constant for the enzyme . substrate complex was obtained by titrations in which absorbance and/or fluorescence quenching was measured. The binding constants of NADPH to the enzyme with and without p-hydroxybenzoate were determined kinetically by measuring the rate of reduction of the enzyme at different concentrations of NADPH. The reduction of the enzyme proceeds extremely slowly in the absence of p-hydroxybenzoate. The presence of the substrate causes a dramatic stimulation (140,000-fold) in the rate of enzyme reduction. The anaerobic reduction of the enzyme by NADPH in the presence of p-hydroxybenzoate produces a transient charge-transfer intermediate. On the basis of the proposed mechanism, the dissociation constants for p-hydroxybenzoate and NADPH as well as the Michaelis constants for all the three substrates were calculated from the initial velocity data. The agreement obtained between various kinetic parameters from the initial rate measurements and those calculated from the individual rate constants determined in rapid reactions, strongly supports the proposed mechanism for the p-hydroxybenzoate hydroxylase reaction.  相似文献   

20.
We re-examined the kinetics of the bisphosphatase reaction of rat hepatic 6-phosphofructo-2-kinase/fructose-2,6-bisphosphatase after depleting the enzyme of bound fructose 6-phosphate and found a hyperbolic dependence on fructose 2,6-bisphosphate at concentrations below 100 nM. The Michaelis constant was 4 nM, the Vmax was about 12 nmol X mg-1 X min-1 at 22 degrees C but the substrate inhibited at concentrations above 100 nM. Both phosphate and alpha-glycerol phosphate strongly inhibited phosphoenzyme formation and hydrolytic rate below 100 nM, but relieved the inhibition by substrate at higher concentrations probably by antagonizing substrate binding. A number of observations support the proposition that the phosphoenzyme is a necessary participant in catalysis. 1) The amount of phosphoenzyme measured during steady-state hydrolysis as a function of substrate concentration correlated with the velocity profile. 2) Rapid mixing experiments demonstrated that over a broad range of substrate concentrations phosphoenzyme formation was faster than the net rate of hydrolysis. 3) Both phosphate and alpha-glycerol phosphate inhibited the rate of phosphoenzyme formation and, at low substrate concentrations, reduced the steady-state phosphoenzyme levels. The latter correlated with inhibition of substrate hydrolysis. 4) Both phosphate and alpha-glycerol phosphate stimulate the rate of phosphoenzyme breakdown, consistent with their stimulation of substrate hydrolysis at high substrate concentrations. 5) The fractional rate of phosphoenzyme breakdown, which was pH and substrate dependent, multiplied by the amount of phosphoenzyme obtained in the steady state at that pH and substrate concentration approximated the observed rate of hydrolysis. We conclude that the phosphoenzyme is a reaction intermediate in the hepatic fructose-2,6-bisphosphatase reaction.  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