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.
Pre-steady state, stopped flow analysis of Escherichia coli D-3-phosphoglycerate dehydrogenase was performed by following the fluorescence of protein tryptophan and the fluorescence resonance energy transfer from protein tryptophan to bound NADH. The results indicate that binding of substrates is ordered, with coenzyme, NADH, binding first. Furthermore, the analysis indicated that there are two sets of sites on the tetrameric enzyme that can be differentiated by their kinetic behavior. NADH binding was consistent with an initial binding event followed by a slow conformational change for each site. The slow conformational change is responsible for the apparent tight binding of NADH to the apoenzyme but is too slow to participate in the catalytic cycle when the enzyme is rapidly turning over. Subsequent binding of the substrate, alpha-ketoglutarate, was characterized by a rapid equilibrium binding event followed by a conformational change for each site. Catalysis in the direction of NAD(+) reduction showed a distinct burst of activity followed by a slow rate of turnover, indicating that the rate-limiting step is after hydride transfer. Catalysis in the direction of NADH oxidation did not display burst kinetics, indicating that the rate-limiting step is at or before the hydride transfer step. The burst data indicated that the rate of NAD(+) reduction (3.8 s(-1)) is similar to the k(cat) of the enzyme (2-3 s(-1)) in that direction. However, analysis of the reaction with deuterated NADH failed to show an effect on the velocity of the reaction with a V(H)/V(D)=1.07+/-0.06. None of the other rates determined by stopped flow analysis could account for the k(cat) of the enzyme in either direction (forward k(cat)=0.01 s(-1), reverse k(cat)=2-3 s(-1)), suggesting that the rate-limiting step in both directions is a conformational change in the enzyme that is not detected optically.  相似文献   

2.
The tetrameric molecule of pig skeletal muscle lactate dehydrogenase binds a cationic fluorescent probe, auramine O, at four equal non-interacting sites with a dissociation constant of (1.25 +/- 0.2) X 10(-4) M. Fluorescence of the dye/enzyme mixture is strongly pH-dependent, with a maximum at pH 6.3-6.8. Auramine O-binding sites are located outside the active center of the enzyme. The microenvironment of the bound dye changes upon interaction of lactate dehydrogenase with NAD+, NADH, ADP and pyruvate. The binding of specific ligands induces an increase in fluorescence of auramine O-enzyme complex. This effect was used to determine the dissociation constants of the complexes of lactate dehydrogenase with specific ligands. Pyruvate was demonstrated to bind to the apoenzyme-auramine O complex with a dissociation constant of 5.2 X 10(-4) M. With the use of auramine O, it became possible to reveal subunit interactions within the tetrameric molecule of lactate dehydrogenase. They are manifested in the changes of the microenvironment of a dye-binding site located on one of the subunits induced by the binding of ligands in the active center of a neighboring subunit.  相似文献   

3.
The tetrameric glyceraldehyde-3-phosphate dehydrogenase from rabbit muscle binds NAD+ and some of its analogues in a negatively cooperative manner, whereas other NAD+ analogues bind non-cooperatively to this enzyme. Subsequent to alkylation of a fraction of the active sites of the enzyme with the fluorescent SH reagent N-iodoacetyl-N'-(5-sulfo-1-naphthyl)-ethylenediamine, it was found that the alkylated sites bind NAD+ and NAD+ analogues with a markedly reduced affinity as compared with non-alkylated sites. It was therefore feasible to measure the fluorescence and the circular polarization of the luminescence of the enzyme-bound alkyl groups as a function of binding of NAD+ and of NAD+ analogues to the non-alkylated sites. The changes observed indicate that ligand binding to the non-alkylated sites induces changes in the fluorescence properties of the alkyl groups bound to neighbouring subunits, most likely through the protein moiety. The nature of these changes appears to depend on the structure of the coenzyme analogue. The binding of the non-cooperative binders acetyl-pyridine--adenine dinucleotide, ATP and ADP-ribose induce different conformational changes in the neighbouring vacant subunit, as monitored by the spectroscopic properties of the bound alkyl group. These results in conjunction with other data support the view that the negative cooperativity in NAD+ binding to glyceraldehyde-3-phosphate dehydrogenase results from ligand-induced conformational changes. Furthermore, these results further support the view that subtle structural changes in the coenzyme molecule determine the nature of the conformational changes induced within the enzyme tetramer.  相似文献   

4.
The V of horse liver aldehyde dehydrogenase is enhanced twofold in the presence of 0.5 mm Mg2+ ions when assayed in the dehydrogenase reaction. The mechanism of this activation appears to be related to the fact the enzyme changes from functioning with half-of-the-sites reactivity to functioning with all-of-the-sites reactivity. That is, the presteady-state burst magnitude increases from 2 mol NADH formed per mole of tetrameric enzyme to 4 mol formed per mole (K. Takahashi and H. Weiner, J. Biol. Chem., 1980, 255, 8206–8209). Whether this twofold enhancement correlates, in fact, to a change from half-of-the-sites to all-of-the-sites reactivity of the enzyme by Mg2+ ions was investigated by determining the Stoichiometry of coenzyme binding by fluorescence quenching and enhancement methods in the absence and presence of the metal ions. The biphasic Scatchard plots for NAD binding to the enzyme were similar in the absence and presence of Mg2+ ions, while that of NADH binding was monophasic (-Mg2+) and biphasic (+Mg2+). In the presence of p-methoxyacetophenone, a competitive inhibitor for substrate, the stoichiometric titration of coenzyme binding to the ternary complexes (enzyme-NAD(H)-inhibitor) revealed that only 2 mol of NAD or NADH bind in the absence of Mg2+ ions but 4 bind per mole of tetrameric enzyme in the presence of added metal. The fluorescence intensity of NAD's fluorescent derivative, 1,N6-ethenoadenine dinucleotide, bound to the enzyme was also doubled by the addition of Mg2+ ions.The combined binding data show that the stoichiometry of coenzyme binding to aldehyde dehydrogenase in the ternary complex increases from 2 to 4 mol binding per mole of tetrameric enzyme with the addition of Mg2+ ions. This increase in stoichiometry corresponds to the observed changes of burst magnitude obtained from the presteady-state and V in the steady-state kinetics assays. From both results of the kinetics and stoichiometry, we show that horse liver aldehyde dehydrogenase exhibits half-of-the-sites reactivity when in the tetrameric state in the absence of Mg2+ ions, and all-of-the-sites reactivity in the dimeric state in the presence of the metal.  相似文献   

5.
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   

6.
The unfolding and dissociation of the tetrameric enzyme fructose-1,6-bisphosphatase from pig kidney by guanidine hydrochloride have been investigated at equilibrium by monitoring enzyme activity, ANS binding, intrinsic (tyrosine) protein fluorescence, exposure of thiol groups, fluorescence of extrinsic probes (AEDANS, MIANS), and size-exclusion chromatography. The unfolding is a multistate process involving as the first intermediate a catalytically inactive tetramer. The evidence that indicates the existence of this intermediate is as follows: (1) the loss of enzymatic activity and the concomitant increase of ANS binding, at low concentrations of Gdn.HCl (midpoint at 0.75 M), are both protein concentration independent, and (2) the enzyme remains in a tetrameric state at 0.9 M Gdn.HCl as shown by size-exclusion chromatography. At slightly higher Gdn.HCl concentrations the inactive tetramer dissociates to a compact dimer which is prone to aggregate. Further evidence for dissociation of tetramers to dimers and of dimers to monomers comes from the concentration dependence of AEDANS-labeled enzyme anisotropy data. Above 2.3 M Gdn.HCl the change of AEDANS anisotropy is concentration independent, indicative of monomer unfolding, which also is detected by a red shift of MIANS-labeled enzyme emission. At Gdn.HCl concentrations higher than 3.0 M, the protein elutes from the size-exclusion column as a single peak, with a retention volume smaller than that of the native protein, corresponding to the completely unfolded monomer. In the presence of its cofactor Mg(2+), the denaturated enzyme could be successfully reconstituted into the active enzyme with a yield of approximately 70-90%. Refolding kinetic data indicate that rapid refolding and reassociation of the monomers into a nativelike tetramer and reactivation of the tetramer are sequential events, the latter involving slow and small conformational rearrangements in the refolded enzyme.  相似文献   

7.
Cytoplasmic sn-glycerol-3-phosphate dehydrogenase, labelled covalently with fluorescein isothiocyanate, shows an enzyme-concentration-dependent fluorescence anisotropy. The anisotropy versus enzyme concentration curve is shifted towards higher concentrations when substrates are present. The comparison of the dissociation constants estimated from anisotropy measurements and derived from kinetic experiments suggests that the substrate-induced dissociation of the dimeric dehydrogenase is slow with respect to the enzymatic reaction catalyzed by either its monomeric or dimeric form. The fluorescence anisotropy of the fluorescent dye-labelled dehydrogenase increase with time upon addition of unlabelled fructose-1,6-bisphosphate aldolase approaching a limiting value. This fact indicates the binding of fructose-1,6-bisphosphate aldolase aldose aldolase to glycerolphosphate dehydrogenase. A model is proposed assuming simultaneous binding of tetrameric fructose-1,6-bisphosphate aldolase to monomeric and dimeric glycerolphosphate dehydrogenase with 1:1 stoichiometry. The dissociation constants, as parameters fitted to the experimental curves, were estimated as 0.2 microM and 1 microM for aldolase-dimeric-glycerolphosphate-dehydrogenase and aldolase-monomeric-glycerolphosphate-dehydrogenase complexes respectively.  相似文献   

8.
Experimental conditions favouring the dissociation of tetrameric rabbit muscle D-glyceraldehyde-3-phosphate dehydrogenase into active monomers were elaborated. The urea-induced dissociation of the tetramer was shown to be a stepwise process (in 2 M urea only dimers are formed; an increase in urea concentration up to 3 M causes the splitting of the dimers into monomers). The specific activity of immobilized monomers in the glyceraldehyde-3-phosphate oxidation reaction does not differ from that of the parent immobilized tetrameric form. The tetrameric enzyme molecule binds the coenzyme with a negative cooperativity (the first two NAD+ molecules bind with KD below 0.1 microM; for the third and fourth molecules the dissociation constant was determined to be equal to 5.5 +/- 1.5 microM (50 mM medinal buffer, 10 mM sodium phosphate, pH 8.2). The cooperativity of NAD+ binding is preserved in the immobilized preparation of tetrameric dehydrogenase. The immobilized monomers bind NAD+ with KD of 1.6 +/- 1.0 microM. The experimental results are consistent with the hypothesis according to which the association of catalytically active subunits into a tetramer changes their coenzyme-binding properties in such a way that the first two NAD+ molecules bind more firmly to a tetramer than to a monomer, whereas the third and the fourth NAD+ molecules bind less firmly.  相似文献   

9.
The binding of NAD and NADH to electrophoretically pure 3(17)beta-hydroxysteroid dehydrogenase of Pseudomonas testosteroni was determined by Fluorescence spectroscopy and gel filtration. Four moles of cofactor are bound/mol of tetrameric enzyme; the binding sites are equivalent and independent. The dissociation constants for NAD and NADH are 16 and 0.25 micronM, respectively. As measured by gel filtration in the absence of cofactor, 0.4 mol of estradiol-17 beta is bound/mol of tetrameric enzyme. Data obtained from isotope exchange at equilibrium indicate that the binding of the cofactor to the enzyme is favored over the binding of steroid, although each may bind in the absence of the other. The rates of cofactor dissociation from the ternary complexes are slower than the rates of steroid dissociation; cofactor dissociation is probably the rate-limiting step. Cofactor analogs modified in the pyridine moiety are cosubstrates, whereas modified adenine derivatives are not. The enzyme also utilized as substrate a number of potential steroid affinity labels; no enzyme inactivation by these compounds was observed.  相似文献   

10.
The NAD-glycohydrolase/ADP-ribosyl cyclase CD38 catalyzes the metabolism of nicotinamide adenine dinucleotide (NAD) to the Ca2+ mobilizing second messengers ADP-ribose (ADPR), 2′-deoxy-ADPR, and cyclic ADP-ribose (cADPR). In the present study, we investigated binding and metabolism of NAD by a soluble fragment of human CD38, sCD38, and its catalytically inactive mutant by monitoring changes in endogenous tryptophan (Trp) fluorescence. Addition of NAD resulted in a concentration-dependent decrease in sCD38 fluorescence that is mainly caused by the Trp residue W189. Amplitude of the fluorescence decrease was fitted as one-site binding curve revealing a dissociation constant for NAD of 29 μM. A comparable dissociation constant was found with the catalytically inactive sCD38 mutant (KD 37 μM NAD) indicating that binding of NAD is not significantly affected by the mutation. The NAD-induced decrease in Trp fluorescence completely recovered in case of sCD38. Kinetics of recovery was slowed down with decreasing temperature and sCD38 concentration and increasing NAD concentration demonstrating that recovery in fluorescence is proportional to the enzymatic activity of sCD38. Accordingly, recovery in fluorescence was not observed with the catalytically inactive mutant.This article is part of a Special Issue entitled: ECS Meeting edited by Claus Heizmann, Joachim Krebs and Jacques Haiech.  相似文献   

11.
Cibacron Blue F3GA dye has been used to probe subtle conformational changes in protein structure associated with the conversion of Escherichia coli glutamine synthetase (GS) between relaxed, taut, oxidized, and dissociated forms. Binding of the dye to each form of the enzyme elicits a different spectral perturbation of the dye which can be detected by difference spectroscopy. By following time-dependent changes in the difference spectrum associated with the binding of dye to the enzyme, it was demonstrated that dissociation of subunits provoked either by urea or by relaxation of the enzyme at pH 8.5 is a multiphasic process. In the presence of 3-4 M urea, dissociation of taut GS is associated with an almost instantaneous, transient increase in absorbancy of the difference spectrum at 638 nm and, after a lag, by a progressive decrease in absorbancy at 585 nm and an increase at 700 nm. The kinetics of these changes vary as a function of temperature, pH, and the concentrations of KCl, MnCl2, and urea, probably reflecting differences in the rates of GS relaxation and in the formation of aggregates of intermediate sizes. Results of direct binding measurements show that the taut and relaxed forms of GS can bind only 1-1.3 equiv of dye per subunit, whereas dissociated subunits bind up to 3.0 equiv per subunit. The Kd of the dye-taut GS complex as calculated from binding data was 0.55 microM. The binding of dye to taut GS was inhibited by its substrate, ADP, and by the allosteric effectors AMP and tryptophan. On the basis of the abilities of ADP, AMP, and tryptophan to inhibit the binding of dye to GS, dissociation constants of the respective GS-ligand complexes were 2.4, 121, and 1170 microM, respectively, in good agreement with previously determined values. From the difference spectra obtained between a given concentration of dye in a 5.0-cm cell and 10 times that concentration in a 0.5-cm cell, it was established that at concentrations greater than 5 microM a significant fraction of the dye is present as stacked aggregates. Because only the dye monomer binds to GS, the difference spectrum between dye and dye bound to GS is due in part to GS-promoted shifts in the equilibrium between stacked and unstacked dye molecules. Consequently, with increasing dye concentrations, the amplitude of the dye vs. dye + GS difference spectrum can continue to increase, even after the GS becomes saturated with dye.(ABSTRACT TRUNCATED AT 400 WORDS)  相似文献   

12.
We have studied the conformational stability of the two homologous membrane skeletal proteins, the erythroid and non-erythroid spectrins, in their dimeric and tetrameric forms respectively during unfolding in the presence of urea and guanidine hydrochloride (GuHCl). Fluorescence and circular dichroism (CD) spectroscopy have been used to study the changes of intrinsic tryptophan fluorescence, anisotropy, far UV-CD and extrinsic fluorescence of bound 1-anilinonapthalene-8-sulfonic acid (ANS). Chemical unfolding of both proteins were reversible and could be described as a two state transition. The folded erythroid spectrin and non-erythroid spectrin were directly converted to unfolded monomer without formation of any intermediate. Fluorescence quenching, anisotropy, ANS binding and dynamic light scattering data suggest that in presence of low concentrations of the denaturants (up-to 1M) hydrogen bonding network and van der Waals interaction play a role inducing changes in quaternary as well as tertiary structures without complete dissociation of the subunits. This is the first report of two large worm like, multi-domain proteins obeying twofold rule which is commonly found in small globular proteins. The free energy of stabilization (ΔGu H 2 0) for the dimeric spectrin has been 20 kcal/mol lesser than the tetrameric from.  相似文献   

13.
The spontaneous inactivation of yeast glyceraldehyde-3-phosphate dehydrogenase was found to fit a simple two-state model at pH 8.5 and 25 degrees. The first step is a relatively rapid dissociation of the tetramer to dimers with the equilibrium largely in favor of the tetramer. In the absence of NAD+ the dimer inactivates irreversibly. The apoenzyme is quite stable with a half-life for complete activity loss proportional to the square root of the enzyme concentration. Perturbances of the protein structure (by pH, ionic strength, and specific salts), which have no effect on the tetrameric state of the molecule, result in an alteration of the cooperativity of NAD+ binding, the reactivity of the active-site sulfhydryl group, and the catalytic activity of the enzyme. Covalent modification of two of the four active-site sulfhydryl groups has profound effects on the enzymic activity which are mediated by changes in the subunit interactions. Sedimentation analysis and hybridization studies indicate that the interaction between subunits remains strong after covalent modification. Under normal physiological and equilibrium dialysis conditions the protein is a tetramer. Equilibrium dialysis studies of NAD+ binding to the enzyme at pH 8.5 and 25 degrees reveal a mixed cooperativity pattern. A model consistent with these observations and the observed half-of-the-sites reactivity is that of ligand induced sequential conformational changes which are transferred across strongly interacting subunit domains. Methods for distinguishing negatively cooperative binding patterns from mixtures of denatured enzyme and multiple species are discussed.  相似文献   

14.
Shaffer J  Sun G  Adams JA 《Biochemistry》2001,40(37):11149-11155
The COOH-terminal Src kinase (Csk) regulates a broad array of cellular processes via the specific phosphorylation and downregulation of Src family protein kinases. While Csk has been a topic for steady-state kinetic studies, the individual steps associated with substrate phosphorylation have not been investigated. To understand active-site phenomena, pre-steady-state and transient-state kinetic methods were applied to develop a catalytic pathway for substrate processing. Rapid quench flow techniques show that the phosphorylation of a substrate peptide, generated from a random library, occurs in two kinetic phases: a rapid, exponential "burst" phase followed by a slow, linear phase. The amplitude of the burst phase increases as a function of enzyme concentration, indicating that the biphasic kinetics are not the result of product inhibition. Analysis of the burst rate as a function of substrate concentration indicates that the phosphoryl transfer step is fast (k3 > or = 140 s(-1) and highly favorable (k3/k-3 > or = 6). The apparent dissociation rate constant for ADP (0.6 s(-1), measured using stopped-flow kinetic methods and a fluorescent trapping agent, mant-ATP, is close to kcat. Since the substrate dissociation constant is high, the release of phosphopeptide is not likely to limit turnover. These findings indicate that Csk rapidly delivers the gamma-phosphate of ATP to the substrate and rapidly releases the phosphoproduct. Overall rate limitation in the steady state is then attributed to the slow, net dissociation of ADP. Viscosometric studies suggest that this final event in the catalytic cycle is coupled with slow conformational changes.  相似文献   

15.
The methionyl-tRNA synthetase from Bacillus stearothermophilus is shown to be a dimer of 2 x 82,000 with identical subunits. It exhibits negative cooperativity in substrate binding and "virtual" halt-of-the-sites reactivity. The enzyme binds only 1 mol of methionine in the absence of other ligands, but several methods show that 2 mol of methionyl adenylate are bound per enzyme dimer. However, one of these adenylates is formed 480 times faster than the other (k1 = 29 sec-1 and k2 = 0.06 sec-1). While the rapid phase of the reaction follows normal saturation kinetics with respect to substrate concentration, the rate of the slow phase is independent of substrate concentrations down to 1 muM. It is suggested that the very slow rate of formation of the second adenylate reflects a rate limiting conformational change which precedes a more rapid chemical step on the second subunit.  相似文献   

16.
1. The binding parameters for NADH and NAD+ to rabbit-muscle glyceraldehyde-phosphate dehydrogenase (D-glyceraldehyde-3-phosphate:NAD+ oxidoreductase (phosphorylating), EC 1.2.1.12) have been measured by quenching of the flourescence of the protein and the NADH. 2. The fact that the degree of protein fluorescence quenching by bound NAD+ or NADH, excited at 285 nm and measured at 340 nm ('blue' tryptophans), is not linearly related to the saturation functions of these nucleotides, leads to a slight overestimation of the interaction energy and an underestimation of the concentration of sites, if linearity is assumed. 3. This is also the case for NADH, but not for NAD+, when the protein fluorescence is excited at 305 nm and measured at 390 nm ('red' tryptophans). 4. The binding of NAD+ can be described by a model in which the binding of NAD+, via negative interactions within the dimer, induces weaker binding sites, with the result that the microscopic dissociation constant is 0.08 microM at low saturation and 0.18 microM for the holoenzyme. 5. The binding of NADH can be described on the basis of the same model, the dissociation constant at low saturation being 0.5 microM and of the holoenzyme 1.0 microM. 6. The fluorescence of bound NADH is not sensitive to the conformational changes that cause the decrease in affinity of bound NAD+ or NADH. 7. The binding of NAD+ to the 3-phosphoglyceroyl enzyme can be described by a dissociation constant that is at least two orders of magnitude greater than the dissociation constants of the unacylated enzyme. The affinity of NAD+ to this form of the enzyme is in agreement with the Ki calculated from product inhibition by NAD+ of the reductive dephosphorylation of 1,3-diphosphoglycerate.  相似文献   

17.
Structural changes in the purified (Na+ + K+)-ATPase accompanying detergent inactivation were investigated by monitoring changes in light scattering, intrinsic protein fluorescence, and tryptophan to beta-parinaric acid fluorescence resonance energy transfer. Two phases of inactivation were observed using the non-ionic detergents, digitonin, Lubrol WX and Triton X-100. The rapid phase involves detergent monomer insertion but little change in protein structure or little displacement of closely associated lipids as judged by intrinsic protein fluorescence and fluorescence resonance energy transfer. Lubrol WX and Triton X-100 also caused membrane fragmentation during the rapid phase. The slower phase of inactivation results in a completely inactive enzyme in a particle of 400 000 daltons with 20 mol/mol of associated phospholipid. Fluorescence changes during the course of the slow phase indicate some dissociation of protein-associated lipids and an accompanying protein conformational change. It is concluded that non-parallel inhibition of (Na+ + K+)-ATPase and p-nitrophenylphosphate activity by digitonin (which occurs during the rapid phase of inactivation) is unlikey to require a change in the oligomeric state of the enzyme. It is also concluded that at least 20 mol/mol of tightly associated lipid are necessary for either (Na+ + K+)-ATPase or p-nitrophenylphosphatase activity and that the rate-limiting step in the slow inactivation phase involves dissociation of an essential lipid.  相似文献   

18.
The time course of the overall reaction catalyzed by the pyruvate dehydrogenase multienzyme complex produces an unexpectedly high lag (tau = 8 S) even in the presence of saturating concentrations of its substrates. The preincubation of the pyruvate dehydrogenase complex with one of the substrates alone decreases the duration of this lag, and all the substrates of the pyruvate dehydrogenase component (E1) and dihydrolipoyl transacetylase component (E2) together (pyruvate, thiamine pyrophosphate, and CoA) result in the complete disappearance of the lag. The reduction of the dihydrolipoyl dehydrogenase component (E3) of the pyruvate dehydrogenase complex with the substrates of the complex in the absence of NAD+ produces significantly different quenching in the FAD fluorescence, and then the reduction with the substrates of E3 as dihydrolipoic acid and dithioerythritol. (The formation of FADH2 was not observed in the system.) The higher fluorescence quenching in the presence of substrates of pyruvate dehydrogenase complex compared to the effect caused by the substrates of the E3 component (dihydrolipoic acid and DTE) indicates conformational changes additionally manifested in the fluorescence properties of the enzyme complex. The substrate-induced quenching of the enzyme-bound FAD fluorescence shows biphasic kinetics. The rate constant of the slow phase is comparable with the rate constant calculated from the time duration of the lag phase observed in the overall reaction. The kinetic analysis of both intensity and anisotropy decrease of the FAD fluorescence suggests a consecutive transmittance of an all substrate-coordinated, induced conformational changes directed from the pyruvate dehydrogenase-via the lipoyl transacetylase--to the lipoyl dehydrogenase. Two simultaneous conformational effects caused by binding of the substrates can be distinguished; one of them results the fluorescence of the bound FAD to be more quenched, while the other makes the FAD more mobile. The first-order rate constants of both these conformational changes were determined. The present observations suggest that the pyruvate dehydrogenase complex exists in a partially inactive state in the absence of its substrates, and it becomes active due to conformational changes caused by the binding of its substrates.  相似文献   

19.
The fluorescence of the natural coenzyme, NADH, is used to monitor the environment of the nicotinamide moiety at the active centre of rabbit muscle glyceraldehyde-3-phosphate dehydrogenase (EC 1.2.1.12). Changes of the fluorescence quantum yield and polarization of a small amount of NADH, totally bound by an excess of enzyme, show that at half-saturation of the oligomer with NAD a conformational change is induced which affects the active centre regions of the remaining subunits. This conformational transition is not effected by adenosine diphosphoribose, suggesting that the binding of the nicotinamide moiety of NAD to two subunits is essential for the change of tertiary structure of the remaining subunits that causes the observed changes of the fluorescence properties of the ADH "tracer probe". It is suggested that this conformational transition of the oligomer is responsible for the major decrease of affinity for NAD which occurs at half-saturation, and possibly for the activation by NAD+ of the reductive dephosphorylation reaction catalysed by the enzyme. It is also suggested, by analogy with haemoglobin, that the molecular basis of the negative cooperativity may be the creation of additional intersubunit bonds during the binding of the first two NAD molecules to the tetramer, and a change from a "relaxed" quaternary structure to a "tense" structure at half-saturation.  相似文献   

20.
The binding of NAD+, NADH and adenosine diphosphoribose (Ado-PP-Rib) to a stable, highly active and nucleotide-free preparation of rabbit muscle glyceraldehyde-3-phosphate dehydrogenase (D-glyceraldehyde-3-phosphate: NAD+ oxidoreductase (phosphorylating), EC 1.2.1.12) has been studied. All three nucleotides quench the protein fluorescence to the same extent when they bind to the enzyme, and this property has been used to measure the dissociation constants for the two high-affinity binding sites for the nucleotides. The results indicate negative interactions between, or non-identify of, these two binding sites, to which NAD+ and NADH bind with similar affinity. The binding of NAD+ to the enzyme has been studied by spectrophotometric titrations at 360 nm. It appears that the binding of NAD+ to each of the four subunits of the enzyme contributes equally to the intensity of this 'Racker' band. The dissociation constants associated with the binding of the third and fourth molecules of NAD+ estimated from such titrations confirm some previous estimates. The binding of NADH to the enzyme causes a decrease of intensity of the absorbance of the coenzyme at 340 nm, and the dissociation constants for binding of the third and fourth molecules of NADH have been estimated from spectrophotometric titrations. They are the same as those for NAD+. Judging by the apparent dissociation constants, negative interactions on binding the third molecule of NAD+ or NADH are more marked than those associated with the binding of the second and fourth molecules, suggesting that a major conformational change occurs at half-saturation of the tetramer with coenzyme.  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