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.
Reactive intermediate enzyme complexes are difficult to study directly and the use of physical methods requiring observation periods of more than a second has not been possible heretofore. Here we introduce a simple approach, the "Le Chatelier forcing method" which does for the first time produce significant concentrations of such kinetically competent central intermediates observable for extended periods of time. The method involves only the forcing of the accumulation of intermediate complexes at thermodynamic equilibrium by the use of high reactant concentrations working against a high concentration of a product, combined with a valid and applicable method of analysis. We demonstrate this approach using the glutamate dehydrogenase catalyzed reaction with the reaction product ammonia as a "dam" to oppose the forward driving force of NADP and l-glutamate. We demonstrate the accumulation of substantial amounts measurable amounts of stable enzyme-NADPH-alpha-carbinolamine and alpha-iminoglutarate complexes in three different alpha-amino acid dehydrogenases. We describe the manipulation of such Le Chatelier forced equilibria to increase the prominence of particular species and discuss the implications of these findings for previously unattainable experimental approaches.  相似文献   

2.
The binding reactions of two heterocyclic analogs of protocatechuate (PCA), 2-hydroxyisonicotinic acid N-oxide and 6-hydroxynicotinic acid N-oxide, to Brevibacterium fuscum protocatechuate 3,4-dioxygenase have been characterized. These analogs were synthesized as models for the ketonized tautomer of PCA which we have previously proposed as the form which reacts with O2 in the enzyme complex (Que, L., Jr., Lipscomb, J.D., Munck, E., and Wood, J.M. (1977) Biochim. Biophys. Acta 485, 60-74). Both analogs have much higher affinity for the enzyme than PCA. Repetitive scan optical spectra of each binding reaction show that at least one intermediate is formed. The spectra of the intermediates are red-shifted (lambda max = 500 nm) relative to that of native enzyme (lambda max = 435 nm) but are similar to that of the anaerobic enzyme-PCA complex. In contrast, the spectrum of the final, deadend complex formed by each analog is significantly blue-shifted (lambda max less than 340 nm) resulting in an apparent bleaching of the chromophore of the enzyme. A transient intermediate exhibiting a similar bleached spectrum has been detected in the enzyme reaction cycle immediately after O2 is added to the enzyme-PCA complex (Bull C., Ballou D.P., and Otsuka, S. (1981) J. Biol. Chem. 256, 12681-12686). Stopped flow measurements of the analog binding reactions show that a relatively weak enzyme complex is initially formed followed by at least two isomerizations leading to the bleached, high affinity complexes. EPR spectra of both the early and final complexes reveal only high spin Fe3+ with negative zero field splitting, showing that the optical bleaching is not due to Fe reduction. The studies show that the ketonized analogs are poor models for the enzyme-substrate complex but do successfully mimic many features of the first oxy complex of the reaction cycle. We propose that substrate ketonization occurs coincident with or after O2 binding and may be involved directly in the O2 insertion reaction.  相似文献   

3.
Substrates containing electron-withdrawing groups were reacted with protocatechuate 3,4-dioxygenase and oxygen. Haloprotocatechuates (5-fluoro-, 5-chloro-, 5-bromo-, 2-chloro-, and 6-chloroprotocatechuates) are oxygenated by the enzyme at rates 28- to 3000-fold lower than that with the native substrate. These lower rates are due to both deactivation of substrate to O2 attack, and to the formation of abortive enzyme-substrate (ES) complexes. Such ES complexes with haloprotocatechuates are spectrally distinct from the normal ES complex. 6-Chloroprotocatechuate produces changes more like those due to protocatechuate. The abortive ES complexes, when rapidly mixed with oxygen, decay to free enzyme and product monophasically, and the dependence of the rates on O2 concentration shows that a rate-limiting step precedes reaction with O2. Thus these complexes are rather unreactive toward O2, and the rate-limiting step in oxygenation is their conversion to active complexes. In contrast, the reaction of O2 with the enzyme and 6-chloroprotocatechuate is biphasic, the first phase being dependent on O2 concentration (2 X 10(4) M-1 S-1) and the second not (7 S-1). The intermediate formed after the first phase strongly resembles the second intermediate seen in the reaction of enzyme with protocatechuate and O2 (Bull, C., Ballou, D. P., and Otsuka, S., (1981) J. Biol. Chem. 256, 12681-12686), implying that the electron-withdrawing effect of the chlorine slows the O2 addition step considerably while the conversion to the second intermediate is hardly affected. When the enzyme cycles through several turnovers with 6-chloroprotocatechuate, an enzyme species is formed that resembles the unreactive ES complexes seen with the other haloprotocatechuates, indicating that a small amount of the unreactive complex is in equilibrium with the reactive complex and that during successive turnovers the enzyme is slowly converted into the unreactive form. The formation of this form correlates with the observation that in assays the rate of product formation gradually decreases with time.  相似文献   

4.
The rate of inactivation of succinyl-CoA:3-ketoacid coenzyme A transferase by thiol reagents is increased 3 to 100 times by very low concentrations of acyl-CoA substrates. The same maximum inactivation rate is found with acetoacetyl-CoA and succinyl-CoA. The enhanced rate of inactivation is caused by the stoichiometric formation of the enzyme-CoA intermediate and an accompanying conformation change of the enzyme. The inactivation rate provides a simple assay for the amount of enzyme present as the enzyme-CoA intermediate, using only catalytic concentrations of enzyme. This technique has been utilized to measure (a) a rate constant for hydrolysis of the enzyme-CoA intermediate of 0.10 min-1 at pH 8.1; (b) a stoichiometry of two active sites per enzyme molecule; and (c) the equilibrium constants for formation of the enzyme-CoA intermediate from dilute solutions of substrates (and hence for the overall reaction) by determining the ratio of [enzyme-CoA]/[enzyme] in the presence of a series of substrate "buffers" at different ratios of [RCOO-]/[RCOSCoA]. As the total concentration of acyl-CoA and carbosylate substrates is increased, the inactivation rate is decreased. This indicates that the Michaelis complexes are protected against inactivation.  相似文献   

5.
Direct evidence for an enzyme-bound intermediate in the EPSP synthase reaction pathway has been obtained by rapid chemical quench-flow studies. The transient-state kinetic analysis has led to the following complete scheme: (formula; see text) Values for all 12 rate constants were obtained. Substrate trapping experiments in the forward and reverse reactions established the kinetically preferred order of binding and release of substrates and products and showed that shikimate 3-phosphate (S3P) and 5-enolpyruvoylshikimate 3-phosphate (EPSP) dissociate at rates greater than turnover in each direction. Pre-steady-state bursts of product formation were observed in the reaction in each direction indicating a rate-limiting step following catalysis. Single turnover experiments with enzyme in excess over substrate demonstrated the formation of a transient intermediate in both the forward and reverse reactions. In these experiments, the enzymatic reaction was observed by employing a radiolabel in the enol moiety of either phosphoenol pyruvate (PEP) or EPSP. The separation and quantitation of reaction products were accomplished by HPLC monitoring radioactivity. The intermediate was observed as the transient production of radiolabeled pyruvate, formed due to the breakdown of the intermediate in the acid quench used to stop the reaction. The intermediate was observed within 5-10 ms after the substrates were mixed with enzyme and decayed in a reaction paralleling the formation of product in each direction. Thus, the kinetics demonstrate directly the kinetic competence of the presumed intermediate. No pyruvate was formed, on a time scale which is relevant to catalysis, after incubation of the enzyme with dideoxy-S3P and PEP or with EPSP in the absence of phosphate; and so, the intermediate does not accumulate under these conditions. The intermediate broke down to form PEP and EPSP in addition to pyruvate when the reaction was quenched with base rather than acid; therefore, the intermediate must contain the elements of each product. Other experiments were designed to measure directly the phosphate binding rate and further constrain the PEP binding rate. The overall solution equilibrium constant in the forward direction was determined to be 180 by quantitation of radiolabeled reactants and products in equilibrium after incubation with a low enzyme concentration. The internal, active site equilibrium constant was obtained by incubation of radiolabeled S3P with excess enzyme and high concentrations of phosphate and PEP to provide the ratio of [EPSP]/[S3P] = 2.3, which is largely a measure of K4.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

6.
The denaturant-induced equilibrium unfolding transition of equine beta-lactoglobulin was investigated by ultraviolet absorption, fluorescence, and circular dichroism (CD) spectra. An equilibrium intermediate populates at moderate denaturant concentrations, and its CD spectrum is similar to that of the molten globule state previously observed for this protein at acid pH [Ikeguchi, M., Kato, S., Shimizu, A., and Sugai, S. (1997) Proteins: Struct., Funct., Genet. 27, 567-575]. The unfolding and refolding kinetics were also investigated by the stopped-flow CD and fluorescence. A significant change in the CD intensity was observed within the dead time of measurements (25 ms) when the refolding reaction was initiated by diluting the urea-unfolded protein solution, indicating the transient accumulation of the folding intermediate. The CD spectrum of this burst-phase intermediate agrees well with that of the molten globule state at acid pH. The stability of the burst-phase intermediate was also estimated from the urea-concentration dependence of the burst-phase amplitude, and it shows a fair agreement with that of the equilibrium intermediate. These results indicate that the molten globule state of equine beta-lactoglobulin populates at moderate urea concentration as well as at acid pH and it is equivalent with the kinetic folding intermediate.  相似文献   

7.
1,4,5,6-Tetrahydronicotinamide adenine dinucleotide (H2NADH) has been investigated as a reduced coenzyme analog in the reaction between trans-4-N,N-dimethylaminocinnamaldehyde (I) (lambdamax 398 nm, epsilonmax 3.15 X 10-4 M-minus 1 cm-minus 1) and the horse liver alcohol dehydrogenase-NADH complex. These equilibrium binding and temperature-jump kinetic studies establish the following. (i) Substitution of H2NADH for NADH limits reaction to the reversible formation of a new chromophoric species, lambdamax 468 nm, epsilonmax 5.8 x 10-4 M-minus 1 cm-minus 1. This chromophore is demonstrated to be structurally analogous to the transient intermediate formed during the reaction of I with the enzyme-NADH complex [Dunn, M. F., and Hutchison, J. S. (1973), Biochemistry 12, 4882]. (ii) The process of intermediate formation with the enzyme-NADH complex is independent of pH over the range 6.13-10.54. Although studies were limited to the pH range 5.98-8.72, a similar pH independence appears to hold for the H2NADH system. (iii) Within the ternary complex, I is bound within van der Waal's contact distance of the coenzyme nicotinamide ring. (iv) Formation of the transient intermediate does not involve covalent modification of coenzyme. Based on these findings, we conclude that zinc ion has a Lewis acid function in facilitating the chemical activation of the aldehyde carbonyl for reduction, and that reduced coenzyme plays a noncovalent effector role in this substrate activating step.  相似文献   

8.
The mechanism of stimulation of platelets by thrombin and other proteases was studied by following kinetics of secretion of Ca2+ or ATP. The progress-time curves of secretion were analyzed for rate and total amount released. The reaction of thrombin was perturbed by addition of hydroxylamine or a competitive inhibitor and by variation of pH and it was compared with the reactions of other proteases. Trypsin and papain, with specificities for arginyl residues, induced secretion with a time course that was nearly identical with that induced by thrombin when saturating levels of enzyme were used. At low levels of enzyme, trypsin and papain gave extended lags in the progress-time curves. Higher concentrations of trypsin and papain were required for saturation of the measured parameters. Human plasmin (lysly specificity) and bovine chymotrypsin (aromatic amino acid specificity) failed to induce platelet secretion. Active site inhibited thrombin was also ineffective. Both yield and kinetics depended on pH, with the pH profile for each enzyme similar to its profile for hydrolysis of synthetic substrates. Studies at low pH also showed that the early part of the reaction undergoes a change in rate-determining step from enzyme dependent at low enzyme to enzyme indepdenent at high enzyme. Hydroxylamine, a nucleophile that would be expected to accelerate hydrolytic reactions, actually decreased both the rate of initial reactions and yield. A competitive inhibitor of thrombin also decreased both rate and yield; a calculated inhibition constant was in agreement with the value for a synthetic substrate, suggesting that the interaction of thrombin with platelets is analogous to reaction with substrates. A modification of our previous model is proposed in order to accommodate the results described here and to reaoncile the apparent contradictions that enzyme was found not to turn over in the reaction (Detwiler, T. C., and Feinman, R. D. (1973), Biochemistry 12, 282), that catalytic activity is required (Davey, M. G., and Luscher, E. F. (1967), Nature (London) 216, 875; this paper), and that the reaction is characterized by an apparent equilibrium binding (Tollefsen, D. M., Feagler J. R., and Majerus, P. W. (1974), J. Biol. Chem. 249, 2646). The essential feature is a reversible catalytic step with no dissociation of enzyme from product. This is followed by irreversible, thrombin-independent platelet processes leading to secretion, with yield dependent on the equilibrium concentration of the thrombin product. The model thus has aspects of catalysis, stoichiometry, and an agonist-receptor equilibrium.  相似文献   

9.
The dissociations of porcine heart mitochondrial, bovine heart mitochondrial, and porcine heart cytoplasmic malate dehydrogenase dimers (L-malate: NAD+oxidoreductase, EC 1.1.1.37) have been examined by Sephadex G-100 gel filtration chromatography and sedimentation velocity ultracentrifugation. The porcine mitochondrial enzyme was found to chromatograph as subunits when applied to a gel filtration column at a concentration of .02 muM or less at pH 7.0. The presence of coenzymes shifted the dissociation equilibrium at low enzyme concentrations in favor of dimer formation. Monomer formation was also favored when procine mitochondrial enzyme was incubated at pH 5.0 even at concentrations as high as 120 muM. This shift in equilibrium has been correlated with the increased rate and specificity of sulfhydryl residue modification with N-ethylmaleimide at pH 5.0 (Gregory, E.M., Yost, F.J.,Jr., Rohrbach, M.S., and Harrison, J.H. (1971)J. Biol. Chem. 246, 5491-5497). Bovine mitochondrial enzyme did not exhibit a concentration-dependent disociation under the conditions examined. However, at pH5.0 monomer formation was favored, and correlations could again be drawn with sulfhydryl residue modification (Gregory, E.M. (1975)J.Biol. Chem. 250, 5470-5474). In both mitochondrial enzymes, coenzyme binding was found capable of overcoming the effects of pH on the dissociation equilibrium, and dimer formation was favored. Unlike either of the above mentioned enzymes, porcine cytoplasmic malate dehydrogenase did not dissociate into its monomeric form under any conditions investigated.  相似文献   

10.
M Merle  P V Graves  B Labouesse 《Biochemistry》1984,23(8):1716-1723
The formation of tryptophanyl adenylate catalyzed by tryptophanyl-tRNA synthetase from beef pancreas has been studied by stopped-flow analysis under conditions where the concentration of one of the substrates was largely decreasing during the time course of the reaction. Under such conditions a nonlinear regression analysis of the formation of the adenylate (adenylate vs. time curve) at several initial tryptophan and enzyme concentrations gave an accurate determination of both binding constants of this substrate. The use of the jackknife procedure according to Cornish - Bowden & Wong [ Cornish - Bowden , A., & Wong , J.J. (1978) Biochem. J. 175, 969-976] gave the limit of confidence of these constants. This approach confirmed that tryptophanyl-tRNA synthetase presents a kinetic anticooperativity toward tryptophan in the activation reaction that closely parallels the anticooperativity found for tryptophan binding at equilibrium. Both sites are simultaneously forming the adenylate. The dissociation constants obtained under the present pre-steady-state conditions for tryptophan are KT1 = 1.6 +/- 0.5 microM and KT2 = 18.5 +/- 3.0 microM at pH 8.0, 25 degrees C. The rate constant kf of adenylate formation is identical for both active sites (kf = 42 +/- 5 s-1). The substrate depletion method presently used, linked to the jackknife procedure, proves to be particularly suitable for the determination of the kinetic constants and for the discrimination between different possible kinetic models of dimeric enzyme with high substrate affinity. In such a case this method is more reliable than the conventional method using substrate concentrations in high excess over that of the enzyme.  相似文献   

11.
W J Caspary  D A Lanzo  C Niziak 《Biochemistry》1981,20(13):3868-3875
We have previously shown that the bleomycin-induced autooxidation of ferrous iron follows Michaelis--Menten kinetics which are characteristic of enzymatic reactions [Caspary, W. J., Lanzo, D. A., Niziak, C., Friedman, R., & Bachur, N. R. (1979) Mol. Pharmacol. 16, 256]. In this paper, we identify the iron complexes formed during this reaction. The first is a ferrous iron--bleomycin complex which can be considered the catalyst substrate complex. The product of this reaction is a ferric iron--bleomycin complex which is found in a low-spin and a high-spin form. The relative concentrations of these two forms are a function of pH. Glutathione, a biologically relevant reducing agent, binds to the ferric iron--bleomycin complex, reduces it, and may serve as a model for the reduction of the ferric iron--bleomycin complex to the ferrous state during the catalytic cycle. Oxygen uptake induced by bleomycin and ferrous iron is not inhibited by superoxide dismutase (SOD) or catalase. In the absence of bleomycin, catalase strongly inhibits oxygen uptake. This suggests the presence of a relatively stable intermediate in which the superoxide radical is not readily accessible to superoxide dismutase. At pH 9.3, we are able to observe a transient species by electron spin resonance (ESR). When potassium superoxide is added to the ferric iron--bleomycin complex, the same ESR spectrum is produced. We suggest that a transient species composed of a ferric iron, the superoxide ion, and bleomycin is formed. The precise nature of the binding cannot be determined from the data presented.  相似文献   

12.
Evolutionary optimization of the catalytic effectiveness of an enzyme   总被引:5,自引:0,他引:5  
The kinetic and thermodynamic features of reactions catalyzed by present-day enzymes appear to be the consequence of the evolution of these proteins toward maximal catalytic effectiveness. These features are identified and analyzed (in detail for one substrate-one product enzymes) by using ideas that link the energetics of the reaction catalyzed by an enzyme to the maximization of its catalytic efficiency. A catalytically optimized enzyme will have a value for the "internal" equilibrium constant (Kint, the equilibrium constant between the substrates and the products of the enzyme when all are bound productively) that depends on how close to equilibrium the enzyme maintains its reaction in vivo. Two classes are apparent. For an enzyme that operates near equilibrium, the catalytic efficiency is sensitive to the value of Kint, and the optimum value of Kint is near unity. For an enzyme that operates far from equilibrium, the catalytic efficiency is less sensitive to the value of Kint, and Kint assumes a value that ensures that the rate of the chemical transformation is equal to the rate of product release. In each of these cases, the internal thermodynamics is "dynamically matched", where the concentrations of substrate- and product-containing complexes are equal at the steady state in vivo.  相似文献   

13.
Two species of adenylate kinase isozymes (ATP:AMP phosphotransferase, EC 2.7.4.3) from human Duchenne dystrophic serum were separated by Blue Sepharose CL-6B affinity column chromatography. One of these species was the "aberrant" adenylate kinase isozyme, found specifically in the Duchenne type of this disease (Hamada, M., Okuda, H., Oka, K., Watanabe, T., Ueda, K., Nojima, M., Kuby, S.A., Manship, M., Tyler, F., and Ziter, F. (1981) Biochim. Biophys. Acta 660, 227-237). The separated aberrant form possessed a molecular size of 98,000 (+/- 1,500), whereas the normal serum species of the enzyme was 87,000 (+/- 1,600) by sodium dodecyl sulfate-polyacrylamide gel electrophoresis, by gel filtration, and by sedimentation equilibrium. The sedimentation coefficient of each species was found to be 5.8 S for the aberrant form and 5.6 S for the normal form, respectively. The subunit size (Mr = 24,700) of the aberrant enzyme in 8 M urea proved to be very similar to that of the normal human liver enzyme (Hamada, M., Sumida, M., Okuda, H., Watanabe, T., Nojima, M., and Kuby, S.A. (1982) J. Biol. Chem. 257, 13120-13128), and the normal species subunit (Mr = 21,700) was found to be very similar to that of the normal human muscle enzyme (Kuby, S.A., Fleming, G., Frischat, A., Cress, M.C., and Hamada, M. (1983) J. Biol. Chem. 258, 1901-1907). Both species were tetrameric enzymes in the serum. The amino acid composition for the normal species was similar to that for the muscle-type enzyme, and that for the aberrant species was similar to the liver enzyme, but with some notable exceptions in both cases. Thus, the normal species had no tryptophan and two half-cystine residues/subunit; whereas, there was 1 tryptophan and 4 half-cystine residues/subunit of the aberrant molecule. The amino acid composition of both serum isozymes when compared to their respective muscle or liver-type enzyme differed mainly in the content of Glu, Asp, His, Leu, Ile, Gly. Kinetic properties of the two forms of human serum adenylate kinase were studied at limiting concentrations of both ADP3- and MgADP- in the reverse reaction and of AMP2- and MgATP2- in the forward reaction. The type of reaction mechanism compatible with the data was a two-substrate random quasiequilibrium type of mechanism without independent binding of the substrates and with a rate-limiting step largely at the interconversion of the ternary complexes.  相似文献   

14.
Z X Wang  B Preiss  C L Tsou 《Biochemistry》1988,27(14):5095-5100
Kinetics of inactivation and modification of the reactive thiol groups of creatine kinase by 5,5'-dithiobis(2-nitrobenzoic acid) or iodoacetamide have been compared, the former by following the substrate reaction in presence of the inactivator [Wang, Z.-X., & Tsou, C.-L. (1987) J. Theor. Biol. 127, 253]. The microscopic constants for the reaction of the inactivators with the free enzyme and with the enzyme-substrate complexes were determined. From the results obtained it appears that with respect to ATP both inactivators are noncompetitive whereas for creatine iodoacetamide is competitive but DTNB is not. The formation of the ternary complex protects against the inactivation by both DTNB and iodoacetamide. The inactivation kinetics is monophasic with both inactivators, but under similar conditions, the modification reactions in the presence of the transition-state analogue of creatine-ADP-Mg2+-nitrate show biphasic kinetics as also reported by Price and Hunter [Price, N.C., & Hunter, M.G. (1976) Biochim. Biophys. Acta 445, 364]. If the reactive ternary complex and the enzyme complexed with the transition-state analogue react in the same way with these reagents, the modification of one fast-reacting thiol group for each enzyme molecule leads to complete inactivation, indicating that the enzyme has to be in the dimeric state to be active.  相似文献   

15.
The soluble cytochrome o from Vitreoscilla contains two identical subunits and two hemes. The reduced form binds 2 mol of CO in a cooperative manner with a Hill coefficient near 2 (Tyree, B., and Webster, D. A. (1978) J. Biol. Chem. 253, 6988-6991). This carbonyl compound was photolysed with a dye laser and recombination followed at 437 or 420 nm where maximal absorbance changes were registered. Recombination kinetics were biphasic, and the fast phase was approximately 10 times the rate of the slow phase. Apparent rate constants of both phases showed a nonlinear dependence on CO concentration, respectively, in conformity with a reaction scheme which assumes the transient formation of an intermediate species in both slow and fast reactions. A study of temperature dependence of the reactions gave EA = 2.7 kcal/mol for the slow reaction and EA = 3.2 kcal/mol for the fast reaction below 23 degrees C; above this temperature the slope of the Arrhenius plot for the fast reaction became positive. Maximal rates for both phases were around pH 6.5 and fell to approximately 40% of maximal at pH 12. The binding reaction was affected by even a low concentration of sodium dodecyl sulfate (0.0025%), which changed both the kinetic constant of each phase and the relative contribution of each phase to the reaction. A model which assumes the existence of fast and slow reaction conformers in equilibrium is proposed.  相似文献   

16.
The tetrahedral intermediate formed at the active site of 5-enolpyruvoylshikimate-3-phosphate synthase by reaction of shikimate 3-phosphate with phosphoenolpyruvate was isolated, and its properties in solution and in reaction with enzyme were examined. The intermediate was moderately stable at pH 7.0, with a half-life of 45 min, and showed increasing lifetimes with increasing pH (t1/2 greater than 48 h at pH greater than or equal to 12). The intermediate bound to the enzyme rapidly, with a second order rate constant of 5 x 10(7) M-1 s-1. Upon binding to the enzyme, it reacted to form both products (5-enolpyruvoylshikimate 3-phosphate, Pi) and substrates (shikimate 3-phosphate, phosphoenolpyruvate) in proportions predicted by the rate constants defined previously for reactions occurring at the active enzyme site (Anderson, K.S. Sikorski, J.A., and Johnson, K. A. (1988b) Biochemistry 27, 7395-7406). The kinetics of binding and dissociation of stable phosphonate analogs of the tetrahedral intermediate (Alberg, D., and Bartlett, P.A. (1989) J. Am. Chem. Soc. 111, 2337) were also examined. In comparison to the intermediate, the analogs bound to the enzyme 300-10,000 fold more slowly and at least 300-20,000 times mroe weakly. These results clarify the definitions for kinetic competence of enzyme intermediates and call into question the significance of the slow binding of analogs of transition states or enzyme intermediates.  相似文献   

17.
Net rate constants that define the steady-state rate through a sequence of steps and the corresponding effective energy barriers for two (PO3-)-transfer steps in the phosphoglucomutase reaction were compared as a function of metal ion, M, where M = Mg2+ and Cd2+. These steps involve the reaction of either the 1-phosphate or the 6-phosphate of glucose 1,6-bisphosphate (Glc-P2) bound to the dephosphoenzyme (ED) to produce the phosphoenzyme (EP) and the free monophosphates, glucose 1-phosphate (Glc-1-P) or glucose 6-phosphate (Glc-6-P): EP.M + Glc-1-P----ED.M.Glc-P2----EP.M.Glc-6-P6. Before this comparison was made, net rate constants for the Cd2+ enzyme, obtained at high enzyme concentration via 31P NMR saturation-transfer studies [Post, C. B., Ray, W. J., Jr., & Gorenstein, D. G. (1989) Biochemistry (preceding paper in this issue)], were appropriately scaled by using the observed constants to calculate both the expected isotope-transfer rate at equilibrium and the steady-state rate under initial velocity conditions and comparing the calculated values with those measured in dilute solution. For the Mg2+ enzyme, narrow limits on possible values of the corresponding net rate constants were imposed on the basis of initial velocity rate constants for the forward and reverse directions plus values for the equilibrium distribution of central complexes, since direct measurement is not feasible. The effective energy barriers for both the Mg2+ and Cd2+ enzymes, calculated from the respective net rate constants, together with previously values for the equilibrium distribution of complexes in both enzymic systems [Ray, W. J., Jr., & Long, J. W. (1976) Biochemistry 15, 4018-4025], show that the 100-fold decrease in the kappa cat for the Cd2+ relative to the Mg2+ enzyme is caused by two factors: the increased stability of the intermediate bisphosphate complex and the decreased ability to cope with the phosphate ester involving the 1-hydroxyl group of the glucose ring. In fact, it is unlikely that the efficiency of (PO3-) transfer to the 6-hydroxyl group of bound Glc-1-P (thermodynamically favorable direction) is reduced by more than an order of magnitude in the Cd2+ enzyme. By contrast, the efficiency of the Li+ enzyme in the same (PO3-)-transfer step is less than 4 x 10(-8) that of the Mg2+ enzyme.(ABSTRACT TRUNCATED AT 400 WORDS)  相似文献   

18.
Nature of the fast and slow refolding reactions of iron(III) cytochrome c   总被引:2,自引:0,他引:2  
The fast and slow refolding reactions of iron(III) cytochrome c (Fe(III) cyt c), previously studied by Ikai et al. (Ikai, A., Fish, W. W., & Tanford, C. (1973) J. Mol. Biol. 73, 165--184), have been reinvestigated. The fast reaction has the major amplitude (78%) and is 100-fold faster than the slow reaction in these conditions (pH 7.2, 25 degrees C, 1.75 M guanidine hydrochloride). We show here that native cyt c is the product formed in the fast reaction as well as in the slow reaction. Two probes have been used to test for formation of native cyt c. absorbance in the 695-nm band and rate of reduction of by L-ascorbate. Different unfolded species (UF, US) give rise to the fast and slow refolding reactions, as shown both by refolding assays at different times after unfolding ("double-jump" experiments) and by the formation of native cyt c in each of the fast and slow refolding reactions. Thus the fast refolding reaction is UF leads to N and the slow refolding reaction is Us leads to N, where N is native cyt c, and there is a US in equilibrium UF equilibrium in unfolded cyt c. The results are consistent with the UF in equilibrium US reaction being proline isomerization, but this has not yet been tested in detail. Folding intermediates have been detected in both reactions. In the UF leads to N reaction, the Soret absorbance change precedes the recovery of the native 695-nm band spectrum, showing that Soret absorbance monitors the formation of a folding intermediate. In the US leads to N reaction an ascorbate-reducible intermediate has been found at an early stage in folding and the Soret absorbance change occurs together with the change at 695 nm as N is formed in the final stage of folding.  相似文献   

19.
Silinski P  Fitzgerald MC 《Biochemistry》2002,41(13):4480-4491
4-Oxalocrotonate tautomerase (4-OT) is a multimeric, bacterial enzyme comprised of 6 identical 62-amino acid subunits, which associate under native conditions to form a homo-hexameric structure stabilized entirely by noncovalent interactions. We have previously shown that the GuHCl-induced equilibrium unfolding of 4-OT at pH 8.5 is well modeled as a two-state process involving only hexamer and unfolded monomer; and we have obtained spectroscopic evidence that intermediate state(s) is (are) populated in the equilibrium unfolding reaction at pHs 6.0 and 7.4 [Silinski, P., Allingham, M. J., and Fitzgerald, M. C. (2001) Biochemistry 40, 4493-4502]. Here, we report on the pH-induced equilibrium unfolding of 4-OT using size-exclusion chromatography (SEC), far-UV-circular dichroism (CD) spectroscopy, and catalytic activity measurements over the pH range from 1.5 to 10.1. Our results indicate that the native hexamer of 4-OT is the predominant species in solution at pHs > or =6.2, that a partially folded dimeric state of 4-OT is stabilized in solution at pH 4.8, and that the enzyme is largely denatured in strongly acidic solutions (pH < or =3.1). GuHCl-induced equilibrium unfolding studies on 4-OT at pH 4.8 indicate that the folded 4-OT dimer populated at this pH is stabilized by 11.7 kcal.mol(-1). The results of biophysical studies on a fluorescent analogue of the enzyme, 4-OT(F50Y), and the results of UV photo-cross-linking studies on a synthetically derived 4-OT analogue, 4-OT(P1Bpa), suggest the polypeptide chains in the 4-OT dimer are nativelike in structure with the exception of their C-termini.  相似文献   

20.
The inactivation of glycosidases by 2-deoxy-2-fluoroglycosides has been shown previously to occur via the accumulation of a covalent 2-deoxy-2-fluoro-alpha-D-glucopyranosyl enzyme intermediate [Withers, S. G., & Street, I. P. (1988) J. Am. Chem. Soc. 110, 8551]. Further characterization of this process with Agrobacterium beta-glucosidase is described, and the range of glycosides engaging in this behavior is examined. Inactivation is shown to be accompanied by the release of a stoichiometric "burst" of aglycon, thereby providing a new class of active site titration agents for glycosidases. The rate of inactivation is shown to be very strongly dependent on the leaving group ability of the aglycon, the slowest inactivator studied (p-nitrophenyl2-deoxy-2-fluoro-beta-D-glucopyranoside) yielding only partial inactivation due to turnover of the intermediate becoming competitive with its formation. Such turnover of the intermediate is shown to be greatly accelerated by transglycosylation to a suitable glycoside bound in the aglycon site, resulting in the release of a disaccharide product which was isolated and characterized. The pH dependences of both the formation and the hydrolysis of the 2-deoxy-2-fluoroglycosyl-enzyme closely resemble those of each step for normal catalysis, indicating that the same catalytic groups are involved in both processes. A model system for the partial "steady-state" inactivation observed previously [Withers, S. G., Rupitz, K., & Street, I. P. (1988) J. Biol. Chem. 263, 7929] with certain other glycosidases was established by incubating the enzyme with an inactivator known to undergo relatively rapid transglycosylation in the presence of various concentrations of a suitable reactivator.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

设为首页 | 免责声明 | 关于勤云 | 加入收藏

Copyright©北京勤云科技发展有限公司  京ICP备09084417号