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.
Stopped flow spectrophotometry was used to investigate the kinetics of the transition of the phosphoglycerate dehydrogenase (3-phosphoglycerate: NAD oxidoreductase, EC 1.1.1.95) reaction from the active to the inhibited rate upon the addition of the physiological inhibitor serine. The transition was characterized by a single first order rate constant (kobs,i) which was independent of enzyme concentration. At pH 8.5, kobs,i increased in a hyperbolic manner with serine concentration from 2 to 8 s-1. The increase in kobs,i occurred at serine concentrations where the steady state inhibition was virtually complete. These results indicate that serine inhibition is an allosteric process involving a conformational change in the enzyme. A model is presented in which serine at low concentrations binds exclusively to the inhibited state of the enzyme and shifts the equilibrium toward that state; at high serine concentrations, serine binds to the active state, facilitating its conversion to the inhibited state. An alternative model, which we favor, proposes two classes of inhibitor binding sites. The kinetics of the fluorescence quenching of enzyme-bound NADH by serine (Sugimoto, E., and Pizer, L.I. (1968) J. Biol. Chem. 243, 2090-2098), measured by stopped flow fluorimetry, was also characterized by a single first order rate constant (kobs,f.q.) which was independent of enzyme concentration. At pH 8.5, kobs,f.q. ranged from 0.4 s-1 at low serine concentrations to 1.1 s-1 at high serine concentrations. These results indicate that the fluorescence quenching induced by serine is a manifestation of a structural change in the enzyme. Enzyme and excess NADH were mixed with substrate and serine in the stopped flow instrument, and enzyme-bound NADH fluorescence was monitored by exciting through the protein at 285 nm. A rapid fluorescence quenching process, which occurred within the mixing time, was followed by a slower fluorescence enhancement process which terminated in a steady state level corresponding to the quenched fluorescence of the enzyme NADH serine complex. The rapid quenching was the result of substrate binding (Dubrow, R., and Pizer, L.I. (1977) J. Biol. Chem. 252, 1539-1551). The fluorescence enhancement was characterized by a single first order rate constant whose value for a given serine concentration corresponded with Kobs,j. This data shows that the quenched state of the enzyme-NADH-complex is the state which is directly responsible for the inhibition of enzyme activity. During catalysis the quenched state is achieved from a different initial conformation, and consequently at a different rate, than in the absence of substrate. kobs,j and kobs,f.q. were also measured using glycine, another inhibitor. The ultraviolet difference spectrum between enzyme and enzyme plus serine was determined and proposed to be the result of the same structural change which is responsible for the fluorescence quenching by serine.  相似文献   

2.
The catalytic mechanism of the phosphoglycerate dehydrogenase reaction in both directions was investigated by studying: (a) pre-steady state transients in reduced coenzyme appearance or disappearance or disappearance and in protein fluorescence; (b) deuterium isotope effects on the transients and on the steady state reactions; and (c) the partial reaction between the enzyme-NADH complex and hydroxypyruvate-P. These studies led to the scheme below for the ternary complex interconversion. E1-NADH-hydroxypyruvate-P(1)equilibriumE2-NADH-hydroxypyruvate-P(2)equilibriumE3-NADH-hydroxypyruvate-P + H+(3)equilibriumE3-NAD+-3-phosphoglycerate(4)equilibriumE4-NAD+-3-phosphoglycerate Steps 1,2, and 4 are ternary complex isomerizations. Step 3 is the hydride transfer. Under steady state conditions isomerization 2 is the rate-determining step in the direction of hydroxypyruvate-P reduction at higher pH values. At lower pH values, the hydride transfer step is also partially rate-determining. The rate-determining step in the direction of 3-phosphoglycerate oxidation occurs subsequent to the hydride transfer step at higher pH values. At lower pH values the rate is determined by both isomerization 4 and the hydride transfer step. Isomerizations 1, 2, and 4 were inhibited by serine, an allosteric inhibitor, indicating that the inactive conformation of the enzyme is incapable of performing any of the steps of the ternary complex interconversion. Phosphoglycerate dehydrogenase corresponds to a V-type allosteric enzyme. When the enzyme-NADH complex was mixed with hydroxypyruvate-P at pH 8.5, a rapid quenching of enzymebound NADH fluorescence occurred. This process was studied under pseudo-first order conditions and shown to be the result of hydroxypyruvate-P binding.  相似文献   

3.
The time-resolved and steady state fluorescence properties were measured for pig heart cytoplasmic malate dehydrogenase at pH 6.0 and 8.0. The fluorescence decay can be described by two rate processes, according to the functions: I(t) = 0.7e(-t/1.0) + 0.3e(-t/4.4) for the free enzyme and I(t) = 0.7e(-t/0.8) + 0.3e(-t/2.0 for the enzyme . NADH complex. Quenching by NADH of the tryptophan fluorescence is linear. The only effect of pH is to change the association constant for NADH binding; the fluorescence of the free enzyme and the fluorescence quenching by NADH, I-, and acrylamide are unaffected by pH. Thus there are no changes in conformation of the free enzyme or of the NADH complex over the range of pH 6 to 8.  相似文献   

4.
1. The NADP-dependent glutamate dehydrogenase of Neurospora crassa undergoes slow reversible structural transitions, with half-times in the order of a few minutes, between active and inactive states. The inactive state of the enzyme, which predominates at pH values below 7.0, has an intrinsic tryptophan fluorescence 25% lower than that of the active state, which predominates at pH values above 7.6. The inactive state can be activated either by an increase in pH or by addition of activators such as succinate. 2. The kinetics of the slow transitions that follow activating and inactivating rapid changes in conditions have been monitored by measurements of protein fluorescence. The results show that the slow reversible conformational change detected by the change in fluorescence is the rate-limiting process for enzyme activation and inactivation. 3. In both directions this conformational change follows apparent first-order kinetics and the rate constant is independent of protein concentration. These kinetics and published measurements of molecular weight are indicative of an isomerization process. 4. In both directions the changes show a large energy of activation and a large positive entropy of activation, consistent with a considerable disturbance of conformation in the transition state. 5. Comparisons of the fluorescence emission spectra of the active and inactive states indicate that the difference in fluorescence is produced by quenching, possibly intramolecular, in the inactive conformation. Iodide ions cause similar quenching. 6. In some mutationally altered forms of the enzyme comparable but modified conformational changes can be followed by protein fluorescence.  相似文献   

5.
Citrate synthase from Escherichia coli enhances the fluorescence of its allosteric inhibitor, NADH, and shifts the peak of emission of the coenzyme from 457 to 428 nm. These effects have been used to measure the binding of NADH to this enzyme under various conditions. The dissociation constant for the NADH-citrate synthase complex is about 0.28 muM at pH 6.2, but increases toward alkaline pH as if binding depends on protonation of a group with a pKa of about 7.05. Over the pH range 6.2-8.7, the number of binding sites decreases from about 0.65 to about 0.25 per citrate synthase subunit. The midpoint of this transition is at about pH 7.7, and it may be one reflection of the partial depolymerization of the enzyme which is known to occur in this pH range. A gel filtration method has been used to verify that the fluorescence enhancement technique accurately reveals all of the NADH molecules bound to the enzyme in the concentration range of interest. NAD+ and NADP+ were weak competitive inhibitors of NADH binding at pH 7.8 (Ki values greater than 1 mM), but stronger inhibition was shown by 5'-AMP and 3'-AMP, with Ki values of 83 +/- 5 and 65 +/- 4 muM, respectively. Acetyl-CoA, one of the substrates, and KCl, an activator, also inhibit the binding in a weakly cooperative manner. All of these effects are consistent with kinetic observations on this system. We interpret our results in terms of two types of binding site for nucleotides on citrate synthase: an active site which binds acetyl-CoA, the substrate, or its analogue 3'-AMP; and an allosteric site which binds NADH or its analogue 5'-AMP and has a lesser affinity for other nicotinamide adenine dinucloetides. When the active site is occupied, we propose that NADH cannot bind to the allosteric site, but 5'-AMP can; conversely, when NADH is the in the allosteric site, the active site cannot be occupied. In addition to these two classes of sites, there must be points for interaction with KCl and other salts. Oxaloacetate, the second substrate, and alpha-ketoglutarate, an inhibitor whose mode of action is believed to be allosteric, have no effect on NADH binding to citrate synthase at pH 7.8. When NADH is bound to citrate synthase, it quenches the intrinsic tryptophan fluorescence of the enzyme. The amount of quenching is proportional to the amount of NADH bound, at least up to a binding ratio of 0.50 NADH per enzyme subunit. This amount of binding leads to the quenching of 53 +/- 5% of the enzyme fluorescence, which means that one NADH molecule can quench all the intrinsic fluorescence of the subunit to which it binds.  相似文献   

6.
Inorganic phosphate, a strong activator of glutamate dehydrogenase at pH 8.0–9.0, is an inhibitor at pH 6.0–7.6. The extent of inhibition increases with the decrease of pH. The same effect is shown by other electrolytes, including Tris-hydroxymethyl-aminomethane and NaCl.The combined effect of pH and ionic strength also alters the allosteric characteristics of the enzyme. Lowering the pH minimizes the activation by high concentrations of NAD; phosphate partially restores this activation. The allosteric activation by ADP disappears at pH around neutrality; in the pH range 6.0–7.0, ADP becomes a strong inhibitor, the inhibition being enhanced by the addition of ionic compounds. Similarly, the extent of allosteric inhibition by guanosine 5′-triphosphate (pyro) (GTP), which is maximal at pH 9.0, decreases at lower pH values and a slight activation is observed in the presence of electrolytes at pH 6.0.Glutamate dehydrogenase, selectively desensitized by dinitrophenylation in the presence of ADP, can be activated by ADP at pH 9.0, but is no longer inhibited by the same effector at pH 6.0, high salt concentration. The densensitized enzyme is not inhibited by GTP at pH 9.0, but is activated by this effector at pH 6.0 in the presence of ionic compounds. Conversely, GTP-protected dinitrophenylated glutamate dehydrogenase is desensitized only to the effect of the activating modifier, ADP at pH 9.0, GTP at pH 6.0, high salt concentration. These findings suggest that the conformation of each allosteric site of glutamate dehydrogenase is changed by pH and ionic strength so that it keeps its specificity for the ligand which brings about a given effect, activation or inhibition, independently from its chemical structure.  相似文献   

7.
Lipoamide dehydrogenase, a component of the bovine adrenal ketoglutarate dehydrogenase complex, catalyzes the oxidation of NADH by p-quinones and ferricyanide. The kinetics of oxidation obey the ping-pong mechanism. At pH 7.0, the constants for the active center oxidation by quinones (kox) are equal to 1.1 X 10(4)-5.3 X 10(5) M-1s-1 and increase as the acceptor potential rises. The values of kox for quinones change insignificantly within the pH range of 7.7-5.0, whereas that for ferricyanide increases 10-fold with a decrease of pH from 7.0 to 5.0. The value of the catalytic constant for the enzyme (kcat) reaches its maximum at pH 5.5. The quinones interact with the thiol groups of lipoamide dehydrogenase by inhibiting the fluorescence of FAD and diaphorease activity. The reaction is catalyzed by a basic amino acid (pK 6.7) within the composition of the enzyme.  相似文献   

8.
The binding of NADH and NAD+ to the human liver cytoplasmic, E1, and mitochondrial, E2, isozymes at pH 7.0 and 25 degrees C was studied by the NADH fluorescence enhancement technique, the sedimentation technique, and steady-state kinetics. The binding of radiolabeled [14C]NADH and [14C]NAD+ to the E1 isozyme when measured by the sedimentation technique yielded linear Scatchard plots with a dissociation constant of 17.6 microM for NADH and 21.4 microM for NAD+ and a stoichiometry of ca. two coenzyme molecules bound per enzyme tetramer. The dissociation constant, 19.2 microM, for NADH as competitive inhibitor was found from steady-state kinetics. With the mitochondrial E2 isozyme, the NADH fluorescence enhancement technique showed only one, high-affinity binding site (KD = 0.5 microM). When the sedimentation technique and radiolabeled coenzymes were used, the binding studies showed nonlinear Scatchard plots. A minimum of two binding sites with lower affinity was indicated for NADH (KD = 3-6 microM and KD = 25-30 microM) and also for NAD+ (KD = 5-7 microM and KD = 15-30 microM). A fourth binding site with the lowest affinity (KD = 184 microM for NADH and KD = 102 microM for NAD+) was observed from the steady-state kinetics. The dissociation constant for NAD+, determined by the competition with NADH via fluorescence titration, was found to be 116 microM. The number of binding sites found by the fluorescence titration (n = 1 for NADH) differs from that found by the sedimentation technique (n = 1.8-2.2 for NADH and n = 1.2-1.6 for NAD+).(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

9.
1. L-Lactate dehydrogenase from lettuce (Lactuca sativa) leaves was purified to electrophoretic homogeneity by affinity chromatography. 2. In addition to its NAD(H)-dependent activity with L-lactate and pyruvate, the enzyme also catalyses the reduction of hydroxypyruvate and glyoxylate. The latter activities are not due to a contamination of the enzyme preparations with hydroxypyruvate reductase. 3. The enzyme shows allosteric properties that are markedly by the pH. 4. ATP is a potent inhibitor of the enzyme. The kinetic data suggest that the inhibition by ATP is competitive with respect to NADH at pH 7.0 and 6.2. The existence of regulatory binding sites for ATP and NADH is discussed. 5. Bivalent metal cations and fructose 6-phosphate relieve the ATP inhibition of the enzyme. 6. A function of leaf L-lactate dehydrogenase is proposed as a component of the systems regulating the cellular pH and/or controlling the concentration of reducing equivalents in the cytoplasm of leaf cells.  相似文献   

10.
Laser flash photolysis has been used to investigate the kinetics of reduction of trimethylamine dehydrogenase by substoichiometric amounts of 5-deazariboflavin semiquinone, and the subsequent intramolecular electron transfer from the FMN cofactor to the Fe4S4 center. The initial reduction event followed second-order kinetics (k = 1.0 x 10(8) M-1 s-1 at pH 7.0 and 6.4 x 10(7) M-1 s-1 at pH 8.5) and resulted in the formation of the neutral FMN semiquinone and the reduced iron-sulfur cluster (in a ratio of approximately 1:3). Following this, a slower, protein concentration independent (and thus intramolecular) electron transfer was observed corresponding to FMN semiquinone oxidation and iron-sulfur cluster reduction (k = 62 s-1 at pH 7.0 and 30 s-1 at pH 8.5). The addition of the inhibitor tetramethylammonium chloride to the reaction mixture had no effect on these kinetic properties, suggesting that this compound exerts its effect on the reduced form of the enzyme. Treatment of the enzyme with phenylhydrazine, which introduces a phenyl group at the 4a-position of the FMN cofactor, decreased both the rate constant for reduction of the protein and the extent of FMN semiquinone production, while increasing the amount of iron-sulfur center reduction, consistent with the results obtained with the native enzyme. Experiments in which the kinetics of reduction of the enzyme were determined during various stages of partial reduction were also consistent with these results, and further indicated that the FMN semiquinone form of the enzyme is more reactive toward the deazariboflavin reductant than is the oxidized FMN.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

11.
Pig heart lactate dehydrogenase was studied in the direction of pyruvate and NADH formation by recording rapid changes in extinction, proton concentration, nucleotide fluorescence and protein fluorescence. Experiments measuring extinction changes show that there is a very rapid formation of NADH within the first millisecond and that the amplitude of this phase (phase 1) increases threefold over the pH range 6-8. A second transient rate (phase 2) can also be distinguished (whose rate is pH-dependent), followed by a steady-state rate (phase 3) of NADH production. The sum of the amplitudes of the first two phases corresponds to 1mol of NADH produced/mol of active sites of lactate dehydrogenase. Experiments that measured the liberation of protons by using Phenol Red as an indicator show that no proton release occurs during the initial very rapid formation of NADH (phase 1), but protons are released during subsequent phases of NADH production. Fluorescence experiments help to characterize these phases, and show that the very rapid phase 1 corresponds to the establishment of an equilibrium between E(NAD) (Lactate) right harpoon over left harpoon H(+)E(NADH) (Pyruvate). This equilibrium can be altered by changing lactate concentration or pH, and the H(+)E(NADH) (Pyruvate) species formed has very low nucleotide fluorescence and quenched protein fluorescence. Phase 2 corresponds to the dissociation of pyruvate and a proton from the complex with a rate constant of 1150s(-1). The observed rate constant is slower than this and is proportional to the position of the preceding equilibrium. The E(NADH) formed has high nucleotide fluorescence and quenched protein fluorescence. The reaction, which is rate-limiting during steady-state turnover, must then follow this step and be involved with dissociation of NADH from the enzyme or some conformational change immediately preceding dissociation. Several inhibitory complexes have also been studied including E(NAD+) (Oxamate) and E(NADH) (Oxamate') and the abortive ternary complex E(NADH) (Lactate). The rate of NADH dissociation from the enzyme was measured and found to be the same whether measured by ligand displacement or by relaxation experiments. These results are discussed in relation to the overall mechanism of lactate dehydrogenase turnover and the independence of the four binding sites in the active tetramer.  相似文献   

12.
Enzymatic properties, renaturation and metabolic role of mannitol-1-phosphate dehydrogenase from Escherichia coli. D-mannitol-1-phosphate dehydrogenase was purified to homogeneity from Escherichia coli, and its physicochemical and enzymatic properties were investigated. The molecular weight of the polypeptide chain is 45,000 as determined by polyacrylamide gel electrophoresis in denaturing conditions. High performance size exclusion chromatography gives an apparent molecular weight of 47,000 for the native enzyme, showing that D-mannitol-1-phosphate dehydrogenase is a monomeric NAD-dependent dehydrogenase. D-mannitol-1-phosphate dehydrogenase is rapidly denatured by 6 M guanidine hydrochloride. Non-superimposable transition curves for the loss of activity and the changes in fluorescence suggest the existence of a partially folded inactive intermediate. The protein can be fully renatured after complete unfolding, and the regain of both native fluorescence and activity occurs rapidly within a few seconds at pH 7.5 and 20 degrees C. Such a high rate of reactivation is unusual for a protein of this size. D-mannitol-1-phosphate dehydrogenase is specific for mannitol-1-phosphate (or fructose-6-phosphate) as a substrate and NAD+ (or NADH) as a cofactor. Zinc is not required for the activity. The affinity of D-mannitol-1-phosphate dehydrogenase for the reduced or oxidized form of its substrate or cofactor remains constant with pH. The affinity for NADH is 20-fold higher than for NAD+. The forward and reverse catalytic rate constants of the reaction: mannitol-1-phosphate + NAD+ in equilibrium fructose-6-phosphate + NADH have different pH dependences. The oxidation of mannitol-1-phosphate has an optimum pH of 9.5, while the reduction of fructose-6-phosphate has its maximum rate at pH 7.0. At pH values around neutrality the maximum rate of reduction of fructose-6-phosphate is much higher than that of oxidation of mannitol-1-phosphate. The enzymatic properties of isolated D-mannitol-1-phosphate dehydrogenase are discussed in relation to the role of this enzyme in the intracellular metabolism.  相似文献   

13.
D H Ozturk  I Park  R F Colman 《Biochemistry》1992,31(43):10544-10555
A new guanosine nucleotide has been synthesized and characterized: guanosine 5'-O-[S-(3-bromo-2-oxopropyl)]thiophosphate (GMPSBOP), with a reactive functional group which can be placed at a position equivalent to the pyrophosphate region of GTP. This new analog is negatively charged at neutral pH and is similar in size to GTP. GMPSBOP has been shown to react with bovine liver glutamate dehydrogenase with an incorporation of 2 mol of reagent/mol of subunit. The modification reaction desensitizes the enzyme to inhibition by GTP, activation by ADP, and inhibition by high concentrations of NADH, but does not affect the catalytic activity of the enzyme. The rate constant for reaction of GMPSBOP with the enzyme exhibits a nonlinear dependence on reagent concentration with KD = 75 microM. The addition to the reaction mixture of alpha-ketoglutarate, GTP, ADP, or NADH alone results in little decrease in the rate constant, but the combined addition of 5 mM NADH with 0.4 mM GTP or with 10 mM alpha-ketoglutarate reduces the reaction rate approximately 6-fold. GMPSBOP modifies peptides containing Met-169 and Tyr-262, of which Tyr-262 is not critical for the decreased sensitivity of the enzyme toward allosteric ligands. The presence of 0.4 mM GTP plus 5 mM NADH protects the enzyme against reaction at both Met-169 and Tyr-262, but yields enzyme with 1 mol of reagent incorporated/mol of subunit which is modified at an alternate site, Met-469. In the presence of 0.2 mM GTP + 0.1 mM NADH, protection against modification of Tyr-262, but only partial protection against labeling of Met-169, is observed. In contrast, the presence of 10 mM alpha-ketoglutarate + 5 mM NADH protect only against reaction with Met-169. The results suggest that GMPSBOP reacts at the GTP-dependent NADH regulatory site [Lark, R. H., & Colman, R. F. (1986) J. Biol. Chem. 261, 10659-10666] of bovine liver glutamate dehydrogenase, which markedly affects the sensitivity of the enzyme to GTP inhibition. The reaction of GMPSBOP with Met-169 is primarily responsible for the altered allosteric properties of the enzyme.  相似文献   

14.
1. No discontinuities were observed during the continuous titration with NADH of the lactate dehydrogenases of ox muscle, pig heart, pig muscle, rabbit muscle, dogfish muscle or lobster tail muscle. The binding was monitored by either the enhanced fluorescence of bound NADH or the quenched fluorescence of the protein. A single macroscopic dissociation constant, independent of protein concentration, could be used to describe the binding to each enzyme, and there was no need to postulate the involvement of molecular relaxation effects. 2. The affinity for NADH decreases only threefold between pH6 and 8.5. Above pH9 the affinity decreases more rapidly with increasing pH and is consistent with a group of about pK9.5 facilitating binding. Muscle enzymes bind NADH more weakly than does the pig heart enzyme. 3. Increasing temperature and increasing concentrations of ethanol both weaken NADH binding. 4. NADH binding is weakened by increasing ionic strength. NaCl is more effective than similar ionic strengths derived from sodium phosphate or sodium pyrophosphate. 5. Commercial NAD(+) quenches the protein fluorescence of the heart and muscle isoenzymes. Highly purified NAD(+) does not, and its binding was monitored by competition for the NADH-binding sites. A single macroscopic dissociation constant is sufficient to describe NAD(+) binding at the concentrations tested. The dissociation constant is about 0.3mm and is not sensitive to changed ionic strength and to changed pH in the range pH6-8.5.  相似文献   

15.
NAD-dependent 1,2-propanediol dehydrogenase (EC 1.1.1.4) activity was detected in cell-free crude extracts of various propane-grown bacteria. The enzyme activity was much lower in 1-propanol-grown cells than in propane-grown cells of Pseudomonas fluorescens NRRL B-1244, indicating that the enzyme may be inducible by metabolites of propane subterminal oxidation. 1,2-Propanediol dehydrogenase was purified from propane-grown Ps. fluorescens NRRL B-1244. The purified enzyme fraction shows a single-protein band upon acrylamide gel electrophoresis and has a molecular weight of 760,000. It consists of 10 subunits of identical molecular weight (77,600). It oxidizes diols that possess either two adjacent hydroxy groups, or a hydroxy group with an adjacent carbonyl group. Primary and secondary alcohols are not oxidized. The pH and temperature optima for 1,2-propanediol dehydrogenase are 8.5 and 20-25 degrees C, respectively. The activation energy calculated is 5.76 kcal/mol. 1,2-Propanediol dehydrogenase does not catalyze the reduction of acetol or acetoin in the presence of NADH (reverse reaction). The Km values at 25 degrees C, pH 7.0, buffer solution for 1,2-propan1,2-propanediol dehydrogenase are 8.5 and 20-25 degrees C, respectively. The activation energy calculated is 5.76 kcal/mol. 1,2-Propanediol dehydrogenase does not catalyze the reduction of acetol or acetoin in the presence of NADH (reverse reaction). The Km values at 25 degrees C, pH 7.0, buffer solution for 1,2-propan1,2-propanediol dehydrogenase are 8.5 and 20-25 degrees C, respectively. The activation energy calculated is 5.76 kcal/mol. 1,2-Propanediol dehydrogenase does not catalyze the reduction of acetol or acetoin in the presence of NADH (reverse reaction). The Km values at 25 degrees C, pH 7.0, buffer solution for 1,2-propanediol and NAD are 2 X 10(-2) and 9 X 10(-5) M, respectively. The 1,2-propanediol dehydrogenase activity was inhibited by strong thiol reagents, but not by metal-chelating agents. The amino acid composition of the purified enzyme was determined. Antisera prepared against purified 1,2-propanediol dehydrogenase from propane-grown Ps. fluorescens NRRL B-1244 formed homologous precipitin bands with isofunctional enzymes derived from propane-grown Arthrobacter sp. NRRL B-11315, Nocardia paraffinica ATCC 21198, and Mycobacterium sp. P2y, but not from propane-grown Pseudomonas multivorans ATCC 17616 and Brevibacterium sp. ATCC 14649, or 1-propanol-grown Ps. fluorescens NRRL B-1244. Isofunctional enzymes derived from methane-grown methylotrophs also showed different immunological and catalytic properties.  相似文献   

16.
Submitochondrial particles from bovine heart in which NADH dehydrogenase is reduced by either addition of NADH and rotenone or by reversed electron transfer generate 0.9 +/- 0.1 nmol of O2-/min per mg of protein at pH 7.4 and at 30 degrees C. When NADH is used as substrate, rotenone, antimycin and cyanide increase O2- production. In NADH- and antimycin-supplemented submitochondrial particles, rotenone has a biphasic effect: it increases O2- production at the NADH dehydrogenase and it inhibits O2- production at the ubiquinone-cytochrome b site. The generation of O2- by the rotenone, the uncoupler carbonyl cyanide rho-trifluoromethoxyphenylhydrazone and oligomycin at concentrations similar to those required to inhibit energy-dependent succinate-NAD reductase. Cyanide did not affect O2- generation at the NADH dehydrogenase, but inhibited O2- production at the ubiquinone-cytochrome b site. Production of O2- at the NADH dehydrogenase is about 50% of the O2- generation but the ubiquinone-cytochrome b area at pH 7.4. Additivity of the two mitochondrial sites of O2- generation was observed over the pH range from 7.0 to 8.8. AN O2- -dependent autocatalytic process that requires NADH, submitochondrial particles and adrenaline is described.  相似文献   

17.
Lactate dehydrogenase in Phycomyces blakesleeanus.   总被引:1,自引:1,他引:0       下载免费PDF全文
1. An NAD-specific L(+)-lactate dehydrogenase (EC 1.1.1.27) from the mycelium of Phycomyces blakesleeanus N.R.R.L. 1555 (-) was purified approximately 700-fold. The enzyme has a molecular weight of 135,000-140,000. The purified enzyme gave a single, catalytically active, protein band after polyacrylamide-gel electrophoresis. It shows optimum activity between pH 6.7 and 7.5. 2. The Phycomyces blakesleeanus lactate dehydrogenase exhibits homotropic interactions with its substrate, pyruvate, and its coenzyme, NADH, at pH 7.5, indicating the existence of multiple binding sites in the enzyme for these ligands. 3. At pH 6.0, the enzyme shows high substrate inhibition by pyruvate. 3-hydroxypyruvate and 2-oxovalerate exhibit an analogous effect, whereas glyoxylate does not, when tested as substrates at the same pH. 4. At pH 7.5, ATP, which inhibits the enzyme, acts competitively with NADH and pyruvate, whereas at pH 6.0 and low concentrations of ATP it behaves in a allosteric manner as inhibitor with respect to NADH, GTP, however, has no effect under the same experimental conditions. 5. Partially purified enzyme from sporangiophores behaves in entirely similar kinetic manner as the one exhibited by the enzyme from mycelium.  相似文献   

18.
Since coupling between phosphorylation and electron transport cannot be measured directly in intact chloroplasts capable of high rates of photosynthesis, attempts were made to determine ATP/2 e ratios from the quamdum requirements of glycerate and phosphoglycerate reduction and from the extent of oxidation of added NADH via the malate shuttle during reduction of phosphoglycerate in light. These different approaches gave similar results. The quantum requirement of glycerate reduction, which needs 2 molecules of ATP per molecule of NADPH oxidized was found to be pH-dependent. 9-11 quanta were required at pH 7.6, and only about 6 at pH 7.0. The quantum requirement of phosphoglycerate reduction, which consumes ATP and NADPH in a 1/1 ratio, was about 4 both at pH 7.6 ant at 7.0. ATP/2 e ratios calculated from the quantum requirements and the extent of phosphoglycerate accumulation during glycerate reduction were usually between 1.2 and 1.4, occasionally higher, but they never approached 2. Although the chloroplast envelope is impermeable to pyridine nucleotides, illuminated chlrooplasts reduced added NAD via the malate shuttle in the absence of electron acceptors and also during the reduction of glycerate or CO2. When phosphoglycerate was added as the substrate, reduction of pyridine-nucleotides was replaced by oxidation and hydrogen was shuttled into the chloroplasts to be used for phosphoglycerate reduction even under light which was rate-limiting for reduction. This indicated formation of more ATP than NADPH by the electron transport chain. From the rates of oxidation of external NADH and of phosphoglycerate reduction at very low light intensities ATP/2e ratios were calculated to be between 1.1 and 1.4. Fully coupled chloroplasts reduced oxaloacetate in the light at rates reaching 80 and in some instances 130 mumoles times mg-1 chlorophyll times h-1 even though ATP is not consumed in this reaction. The energy transfer inhibitor phlorizin did not significantly suppress this reduction at concentrations which completely inhibited photosynthesis. Uncouplers stimulated oxaloacetate reduction by factors ranging from 1.5 to more than 10. Chloroplasts showing little uncoupler-induced stimulation of oxaloacetate reduction were highly active in photoreducing CO2. Measurements of light intensity dependence of quantum requirements for oxaloacetate reduction gave no indication for the existence of uncoupled or basal electron flow in intact chloroplasts. Rather reduction is brought about by loosely coupled electron transport. It is concluded that coupling of phosphorylation to electron transport in intact chloroplasts is flexible, not tight. Calculated ATP/2e ratios were obtained under con a decreENG  相似文献   

19.
Some kinetic properties of the D(-)-lactate dehydrogenase (EC 1.1.1.28) of Escherichia coli have been investigated. There were marked differences between the kinetic properties of the enzyme studied in situ compared with the in vitro D(-)-lactate dehydrogenase. D(-)-Lactate dehydrogenase in situ showed high substrate inhibition with pyruvate over the pH range 6.0–7.0, whereas the enzyme in vitro did not. The pH optimum for pyruvate reduction by the in situ D(-)-lactate dehydrogenase ranged between pH 7.5 and 7.8, whereas the in vitro enzyme showed its pH optimum between pH 6.8 and 7.0. The pK values of the prototropic groups that controlled the enzymatic activity shift to the acidic region for the in vitro enzyme with respect to the in situ enzyme. In vitro D(-)-lactate dehydrogenase exhibits homotropic interactions with its substrate, pyruvate and its coenzyme, NADH, at pH values ranging between pH 6.0 and 8.5, but the in situ enzyme showed homotropic interactions neither with pyruvate nor with NADH at all pH values studied.  相似文献   

20.
The activity of alpha-ketoglutarate dehydrogenase complex from pigeon breast muscle is controlled by ADP and the reaction products, i. e. succinyl-CoA and NADH. ADP activates the alpha-ketoglutarate dehydrogenase component of the complex, whereas NADH inhibits alpha-ketoglutarate dehydrogenase and lipoyl dehydrogenase. In the presence of NADH the kinetic curve of the complex with respect to alpha-ketoglutarate and NAD and the dependence of upsilon versus [NAD] and upsilon versus [Lip (SH)2] in the lipoyl dehydrogenase reaction are S-shaped. In the absence of inhibitor ADP had no activating effect on lipoyl dehydrogenase; however, in the presence of NADH ADP decreases the cooperativity for NAD. The cooperative kinetics of the constituent enzymes of the complex are indicative of its allosteric properties. Isolation of the alpha-ketoglutarate dehydrogenase complex and its lipoyl dehydrogenase and alpha-ketoglutarate dehydrogenase components in a desensitized state confirms their allosteric nature. It is assumed that NADH effects of isolated alpha-ketoglutarate dehydrogenase is due to a shift in the equilibrium between different oligomeric forms of the enzyme.  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