首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 15 毫秒
1.
Stopped-flow fluorescence spectroscopy has been used to study the reaction of human alpha-thrombin with recombinant hirudin variant 1 (rhir) at 37 degrees C and an ionic strength of 0.125 M. A 35% enhancement in intrinsic fluorescence accompanied formation of the thrombin-rhir complex. Over one third of this enhancement corresponded to a structural change that could be induced by binding of either the NH2-terminal fragment (residues 1-51) or the COOH-terminal fragment (residues 52-65) of rhir. Three kinetic steps were detected for reaction of thrombin with rhir. At high rhir concentrations (greater than or equal to 3 microM), two intramolecular steps with observed rate constants of 296 +/- 5 s-1 and 50 +/- 1 s-1 were observed. By using the COOH-terminal fragment of rhir as a competitive inhibitor, it was possible to obtain an estimate of 2.9 x 10(8) M-1 s-1 for the effective association rate constant at low rhir concentrations. At higher ionic strengths, this rate constant was lower, which is consistent with the formation of the initial complex involving an ionic interaction. The mechanism for the reaction of both the COOH- and NH2-terminal fragments of rhir appeared to involve two steps. When thrombin was reacted with the COOH-terminal fragment at high concentrations (greater than or equal to 6 microM), the bimolecular step occurred within the dead time of the spectrometer and only one intramolecular step, with a rate constant of 308 +/- 5 s-1 was observed. At concentrations of NH2-terminal fragment below 50 microM, its binding to thrombin appeared to be a bimolecular reaction with an association rate constant of 8.3 x 10(5) M-1 s-1. In the presence of saturating concentrations of the COOH-terminal fragment, a 1.7-fold increase in this rate constant was observed. At concentrations of NH2-terminal fragment greater than 50 microM, biphasic reaction traces were observed which suggests a two-step mechanism. By comparing the reaction amplitudes and dissociation constants observed with rhir and its COOH-terminal fragment, it was possible to obtain approximate estimates for the values of the rate constants of different steps in the formation of the rhir-thrombin complex.  相似文献   

2.
An analysis of steady-state kinetics of purified rat liver transketolase shows that the reaction proceeds according to a two-stroke substitution ("ping-pong") mechanism. Based on the kinetic data, a competitive relationship was shown to exist between xylulose-5-phosphate and ribose-5-phosphate for the sites of substrate binding by the substituted form of the enzyme with the formation of a non-productive abortive complex (kd = 125 microM). The values of constants of two monomolecular steps of the reaction (k2 = 42 s-1; k4 = 9.4 s-1) were determined. It was assumed that the maximum rate-limiting step of the transketolase reaction is the degradation of the substituted form of transketolase--ribose-5-phosphate complex having a rate constant of k4.  相似文献   

3.
The rate-limiting step of ethanol oxidation by alcohol dehydrogenase (E) at substrate inhibitory conditions (greater than 500 mM ethanol) is shown to be the dissociation rate of NADH from the abortive E-ethanol-NADH complex. The dissociation rate constant of NADH decreased hyperbolically from 5.2 to 1.4 s-1 in the presence of ethanol causing a decrease in the Kd of NADH binding from 0.3 microM for the binary complex to 0.1 microM for the abortive complex. Correspondingly, ethanol binding to E-NADH (Kd = 37 mM) was tighter than to enzyme (Kd = 109 mM). The binding rate of NAD+ (7 X 10(5) M-1s-1) to enzyme was not affected by the presence of ethanol, further substantiating that substrate inhibition is totally due to a decrease in the dissociation rate constant of NADH from the abortive complex. Substrate inhibition was also observed with the coenzyme analog, APAD+, but a single transient was not found to be rate limiting. Nevertheless, the presence of substrate inhibition with APAD+ is ascribed to a decrease in the dissociation rate of APADH from 120 to 22 s-1 for the abortive complex. Studies to discern the additional limiting transient(s) in turnover with APAD+ and NAD+ were unsuccessful but showed that any isomerization of the enzyme-reduced coenzyme-aldehyde complex is not rate limiting. Chloride increases the rate of ethanol oxidation by hyperbolically increasing the dissociation rate constant of NADH from enzyme and the abortive complex to 12 and 2.8 s-1, respectively. The chloride effect is attributed to the binding of chloride to these complexes, destabilizing the binding of NADH while not affecting the binding of ethanol.  相似文献   

4.
K L Kim  D S Kang  L B Vitello  J E Erman 《Biochemistry》1990,29(39):9150-9159
The steady-state kinetics of the cytochrome c peroxidase catalyzed oxidation of horse heart ferrocytochrome c by hydrogen peroxide have been studied at both pH 7.0 and pH 7.5 as a function of ionic strength. Plots of the initial velocity versus hydrogen peroxide concentration at fixed cytochrome c are hyperbolic. The limiting slope at low hydrogen peroxide give apparent bimolecular rate constants for the cytochrome c peroxidase-hydrogen peroxide reaction identical with those determined directly by stopped-flow techniques. Plots of the initial velocity versus cytochrome c concentration at saturating hydrogen peroxide (200 microM) are nonhyperbolic. The rate expression requires squared terms in cytochrome c concentration. The maximum turnover rate of the enzyme is independent of ionic strength, with values of 470 +/- 50 s-1 and 290 +/- 30 s-1 at pH 7.0 and 7.5, respectively. The limiting slope of velocity versus cytochrome c concentration plots provides a lower limit for the association rate constant between cytochrome c and the oxidized intermediates of cytochrome c peroxidase. The limiting slope varies from 10(6) M-1 s-1 at 300 mM ionic strength to 10(8) M-1 s-1 at 20 mM ionic strength and extrapolates to 5 x 10(8) M-1 s-1 at zero ionic strength. The data are discussed in terms of both a two-binding-site mechanism and a single-binding-site, multiple-pathway mechanism.  相似文献   

5.
Bromopyruvate behaves as an active-site-directed inhibitor of the pyruvate decarboxylase (E1) component of the pyruvate dehydrogenase complex of Escherichia coli. It requires the cofactor thiamin pyrophosphate (TPP) and acts initially as an inhibitor competitive with pyruvate (Ki ca. 90 microM) but then proceeds to react irreversibly with the enzyme, probably with the thiol group of a cysteine residue. E1 catalyzes the decomposition of bromopyruvate, the enzyme becoming inactivated once every 40-60 turnovers. Bromopyruvate also inactivates the intact pyruvate dehydrogenase complex in a TPP-dependent process, but the inhibition is more rapid and is mechanistically different. Under these conditions, bromopyruvate is decarboxylated, and the lipoic acid residues in the lipoate acetyltransferase (E2) component become reductively bromoacetylated. Further bromopyruvate then reacts with the new thiol groups thus generated in the lipoic acid residues, inactivating the complex. If reaction with the lipoic acid residues is prevented by prior treatment of the complex with N-ethylmaleimide in the presence of pyruvate, the mode of inhibition reverts to irreversible reaction with the E1 component. In both types of inhibition of E1, reaction of 1 mol of bromopyruvate/mol of E1 chain is required for complete inactivation, and all the evidence is consistent with reaction taking place at or near the pyruvate binding site.  相似文献   

6.
3-Benzyl-6-chloromethyl-3,4-dihydrocoumarin inhibits human leucocyte elastase (HLE) and porcine pancreatic elastase (PPE) through a mechanism-based process characterized by the following apparent enzyme-inhibitor dissociation constants, Ki, and limiting inactivation rate constants k2: 200 microM (HLE), 69 microM (PPE) and 5.10(-2) s-1 (HLE), 17.7.10(-2) s-1 (PPE) at pH 8.0, 37 degrees C. Bis(4-acyloxyphenyl)methane derivatives with a benzylic halogen as potential leaving group have also been synthesized and studied. They transiently inactivate PPE and HLE through the formation of an acyl-enzyme.  相似文献   

7.
The mechanism of acyl enzyme formation from acyl-CoA derivatives was studied for chicken liver fatty acid synthase in 0.1 M potassium phosphate (pH 7.0) and 1 mM EDTA at 23 degrees C. Three mechanistically important acyl-binding sites exist: a cysteine, 4'-phosphopantetheine, and a hydroxyl (serine). The cysteine was specifically labeled with iodoacetamide, and chemical modification of this labeled enzyme with chloroacetyl-CoA resulted in additional covalent labeling of 4'-phosphopantetheine. Reaction of the enzyme with acetyl-CoA results in 47% oxyester formation, whereas with malonyl-CoA and butyryl-CoA, 57 and 80% are oxyesters, respectively, as judged by treatment of the denatured enzyme with hydroxylamine. Limited proteolysis with trypsin followed by sodium dodecyl sulfate-polyacrylamide gel electrophoresis indicates that the reactive hydroxyl and cysteine are on the same peptide. Butyryl-CoA is a relatively poor primer for steady state fatty acid synthesis, probably because transfer from the hydroxyl-binding site to 4'-phosphopantetheine is inefficient. Quenched flow studies indicate that the rate constants for transfer of acetyl from enzyme-bound acetyl-CoA to native, iodoacetamide-labeled, and iodoacetamide-chloroacetyl-labeled enzyme are 43, 110, and 150 s-1. These results can be interpreted in terms of a random acylation of the hydroxyl, 4'-phosphopantetheine, and cysteine by enzyme-bound acetyl-CoA with rate constants of 150 s-1, less than 110 s-1, and less than 43 s-1, respectively. Alternatively the latter two rate constants could be characteristic of intramolecular transfer between enzyme acylation sites. Structural constraints apparently prevent all three acylation sites from being occupied simultaneously. The rate of deacetylation of the acetylated enzyme by enzyme-bound CoA also is most rapid for the iodoacetamide-chloroacetyl-labeled enzyme.  相似文献   

8.
The mechanism of inhibition of ornithine transcarbamoylase by the bacterial toxin phaseolotoxin [N-delta-(phosphosulphamyl)ornithylalanylhomoarginine] was investigated. Ornithine transcarbamoylase was purified by affinity chromatography from Escherichia coli W argR- by using N-delta-(phosphonoacetyl)ornithine as the ligand. Under steady-state conditions phaseolotoxin inhibition was reversible and exhibited mixed kinetics with respect to carbamoyl phosphate. The apparent Ki and apparent K'i were 0.2 microM and 10 microM respectively. Inhibition with respect to ornithine was noncompetitive, with an apparent Ki of 0.9 microM. These data are consistent with competitive binding of phaseolotoxin to the carbamoyl phosphate-binding site of the enzyme. The toxin also appears to be able to bind to the enzyme-carbamoyl phosphate complex, although, since K'i is 50 times greater than Ki, this event is kinetically much less significant. In the presence of phaseolotoxin ornithine transcarbamoylase exhibited a transient phase of activity before a steady state. This is consistent with low rates of association and dissociation for the toxin with enzyme and the enzyme-toxin complex. Rate constants of 2.5 X 10(4)M-1 X s-1 and 5 X 10(-3)s-1 were estimated for the association and dissociation constants respectively.  相似文献   

9.
Interaction of plasma gelsolin with ADP-actin   总被引:3,自引:0,他引:3  
In the presence of Ca2+, gelsolin forms a very tight, stoichiometric complex with 2 molecules of ADP-G-actin. Removal of free Ca2+ causes the 1:2 complex to dissociate to a 1:1 complex. Gelsolin accelerates the very slow polymerization of ADP-actin, apparently by accelerating the rate of nucleation, but the number concentration of filaments formed is probably less than the gelsolin concentration, indicating that the GA2 complex is not a true nucleus. These results are similar to those obtained for the interaction of gelsolin with ATP-G-actin. Both kinetic and equilibrium measurements demonstrate that the critical concentration of gelsolin-capped ADP-actin filaments (8 microM in 1 mM MgCl2 and 0.2 mM ADP) is the same as for the uncapped filaments, proving that the critical concentration is the same at both ends of the equilibrium polymer in ADP as predicted by theory. The association and dissociation rate constants for the addition of ADP-G-actin at the pointed end of an ADP-F-actin filament are estimated to be 4.6 X 10(4) M-1 s-1 and 0.4 s-1, respectively, about 15-fold lower than the rate constants at the barbed end.  相似文献   

10.
Temperature jump relaxation kinetics of the P-450cam spin equilibrium   总被引:1,自引:0,他引:1  
M T Fisher  S G Sligar 《Biochemistry》1987,26(15):4797-4803
The ferric spin-state equilibrium and relaxation rate of cytochrome P-450 has been examined with temperature jump spectroscopy using a number of camphor analogues known to induce different mixed spin states in the substrate-bound complexes [Gould, P., Gelb, M., & Sligar, S. G. (1981) J. Biol. Chem. 256, 6686]. All temperature-induced spectral changes were monophasic, and the spin-state relaxation rate reached a limiting value at high substrate concentrations. The ferric spin equilibrium constant, Kspin, is defined in terms of the rate constants k1 and k-1 via Kspin = k1/k-1 = [P-450(HS)]/[P-450(LS)] where HS and LS represent high-spin (S = 5/2) and low-spin (S = 1/2) ferric iron, respectively, and the spectrally observed spin-state relaxation rate by kobsd = k1 + k-1. A strong correlation between the fraction of high-spin species and the rate constant, k-1, is observed. For a 3 degrees C temperature jump (from 10 to 13 degrees C), the 23% high-spin tetramethylcyclohexanone complex (Kd = 45 +/- 20 microM) is characterized by a ferric spin relaxation rate of kobsd = 1990 s-1, while the rates for the d-fenchone (41% high spin, Kd = 42 +/- 10 microM) and kobsd = 1990 s-1, while the rates for the d-fenchone (41% high spin, Kd = 42 +/- 10 microM) and camphoroquinone (75% high spin, Kd = 15 +/- 5 microM) complexes are 1430 and 346 s-1, respectively.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

11.
N Stahl  W P Jencks 《Biochemistry》1987,26(24):7654-7667
Phosphorylation of the sarcoplasmic reticulum calcium ATPase, E, is first order with kb = 70 +/- 7 s-1 after free enzyme was mixed with saturating ATP and 50 microM Ca2+; this is one-third the rate constant of 220 s-1 for phosphorylation of enzyme preincubated with calcium, cE.Ca2, after being mixed with ATP under the same conditions (pH 7.0, Ca2+-loaded vesicles, 100 mM KCl, 5 mM Mg2+, 25 degrees C). Phosphorylation of E with ATP and Ca2+ in the presence of 0.25 mM ADP gives approximately 50% E approximately P.Ca2 with kobsd = 77 s-1, not the sum of the forward and reverse rate constants, kobsd = kf + kr = 140 s-1, that is expected for approach to equilibrium if phosphorylation were rate limiting. These results show that (1) kb represents a slow conformational change, rather than phosphoryl transfer, and (2) different pathways are followed for the phosphorylation of E and of cE.Ca2. The absence of a lag for phosphorylation of E with saturating ATP and Ca2+ indicates that all other steps, including the binding of Ca2+ ions and phosphoryl transfer, have rate constants of greater than 500 s-1. Chase experiments with unlabeled ATP or with ethylene glycol bis(beta-aminoethyl ether)-N,N,N',N'-tetraacetic acid (EGTA) show that the rate constants for dissociation of [gamma-32P]ATP and Ca2+ are comparable to kb. Dissociation of ATP occurs at 47 s-1 from E.ATP.Ca2+ and at 24 s-1 from E.ATP. Approximately 20% phosphorylation occurs following an EGTA chase 4.5 ms after the addition of 300 microM ATP and 50 microM Ca2+ to enzyme. This shows that Ca2+ binds rapidly to the free enzyme, from outside the vesicle, before the conformational change (kb). The fraction of Ca2+-free E.[gamma-32P]ATP that is trapped to give labeled phosphoenzyme after the addition of Ca2+ and a chase of unlabeled ATP is half-maximal at 6.8 microM Ca2+, with a Hill slope of n = 1.8. The calculated dissociation constant for Ca2+ from E.ATP.Ca2 is approximately 2.2 X 10(-10) M2 (K0.5 = 15 microM). The rate constant for the slow phase of the biphasic reaction of E approximately P.Ca2 with 1.1 mM ADP increases 2.5-fold when [Ca2+] is decreased from 50 microM to 10 nM, with half-maximal increase at 1.7 microM Ca2+. This shows that Ca2+ is dissociating from a different species, aE.ATP.Ca2, that is active for catalysis of phosphoryl transfer, has a high affinity for Ca2+, and dissociates Ca2+ with k less than or equal to 45 s-1.(ABSTRACT TRUNCATED AT 400 WORDS)  相似文献   

12.
Voltage-jump and light-flash experiments have been performed on isolated Electrophorus electroplaques exposed simultaneously to nicotinic agonists and to the photoisomerizable compound 2,2'-bis-[alpha-(trimethylammonium)methyl]-azobenzene (2BQ). Dose-response curves are shifted to the right in a nearly parallel fashion by 2BQ, which suggests competitive antagonism; dose-ratio analyses show apparent dissociation constants of 0.3 and 1 microM for the cis and trans isomers, respectively. Flash-induced trans----cis concentration jumps produce the expected decrease in agonist-induced conductance; the time constant is several tens of milliseconds. From the concentration dependence of these rates, we conclude that the association and dissociation rate constants for the cis-2BQ-receptor binding are approximately 10(8) M-1 s-1 and 60 s-1 at 20 degrees C; the Q10 is 3. Flash-induced cis----trans photoisomerizations produce molecular rearrangements of the ligand-receptor complex, but the resulting relaxations probably reflect the kinetics of buffered diffusion rather than of the interaction between trans-2BQ and the receptor. Antagonists seem to bind about an order of magnitude more slowly than agonists at nicotinic receptors.  相似文献   

13.
Allosamidin, a glycoside antibiotic, is shown to be a strong, competitive inhibitor of semi-purified chitinase from yeast cells of Candida albicans. The inhibitory potency of allosamidin was pH-dependent, with IC50 values of 280 nM at pH 5.0 and 21 nM at pH 7.5. At higher, micromolar, concentrations, allosamidin inactivated this chitinase in a time- and concentration-dependent manner. Kinetic studies of this inactivation provided evidence for the formation of a reversible complex between allosamidin and chitinase, characterized by Kinact = 5 microM, followed by irreversible modification of the enzyme with velocity constant k2 = 4.6 x 10(-3) s-1. Chemical modification studies with the use of group-specific reagents suggested the presence of Glu/Asp carboxyl group(s) at or near the active site, that were important for enzyme activity. The carboxyl-specific reagent, 1-ethyl-3(3-dimethylaminopropyl)-carbodiimide, inactivated the chitinase in a single step process, with apparent second-order rate constant of 0.014 M-1 s-1.  相似文献   

14.
"Suicide" inactivation occurs during catalysis by thromboxane synthase. Loss of enzymatic activity, accompanying thromboxane B2 formation, was proportional to the substrate concentration. Inactivation was directly related to product formation: for several different experimental protocols 50% loss of thromboxane synthase activity corresponded with formation of 454 +/- 79 ng of thromboxane B2/mg protein. The time course of inactivation was pseudo-first-order and obeyed saturation kinetics. Inactivation (KI) and first-order rate constants (ki) were 18 microM and 0.18 s-1 for prostaglandin H2. Prostaglandin H1, a poor substrate for turnover, was also a site-directed inactivator with KI = 28 microM and ki = 0.09 s-1. Competitive inhibitors, typified by U63557a and U46619, preserved the enzyme activity by slowing the rate of inactivation from 0.18 to 0.05 s-1. Loss of the hemoprotein Soret absorbance did not correlate quantitatively or temporally with the loss of thromboxane synthase activity. A similar, irreversible inactivation accompanied thromboxane formation by intact platelets. Loss of activity was proportional to substrate concentration and catalytic activity. For a pool of 25 separate donors, thromboxane synthase activity declined exponentially as a function of thromboxane B2 formation: 50% loss of activity corresponded to 23 ng of thromboxane B2/10(7) platelets. The data conform to criteria for a specific, mechanism-based process in which thromboxane synthase participates in two parallel reactions, one leading to thromboxane formation and the other to suicide inactivation. The specific, rather than indiscriminate, nature of the process, and its occurrence in intact platelets may have implications for the cell biology of thrombosis. Depletion of thromboxane synthase activity may be a factor in the choice and effectiveness of antithrombotic agents.  相似文献   

15.
By gel filtration and titration on DEAE-cellulose filters we show that Escherichia coli tryptophanyl-tRNA synthetase forms tryptophanyl adenylate as an initial reaction product when the enzyme is mixed with ATP-Mg and tryptophan. This reaction precedes the synthesis of the tryptophanyl-ATP ester known to be formed by this enzyme. The stoichiometry of tryptophanyl adenylate synthesis is 2 mol per mole of dimeric enzyme. When this reaction is studied either by the stopped-flow method, by the fluorescence changes of the enzyme, or by radioactive ATP depletion, three successive chemical processes are identified. The first two processes correspond to the synthesis of the two adenylates, at very different rates. The rate constants of tryptophanyl adenylate synthesis are respectively 146 +/- 17 s-1 and 3.3 +/- 0.9 s-1. The third process is the synthesis of tryptophanyl-ATP, the rate constant of which is 0.025 s-1. The Michaelis constants for ATP and for tryptophan in the activation reaction are respectively 179 +/- 35 microM and 23.9 +/- 7.9 microM, for the fast site, and 116 +/- 45 microM and 3.7 +/- 2.2 microM, for the slow site. No synergy between ATP and tryptophan can be evidenced. The data are interpreted as showing positive cooperativity between the subunits associated with conformational changes evidenced by fluorometric methods. The pyrophosphorolysis of tryptophanyl adenylate presents a Michaelian behavior for both sites, and the rate constant of the reverse reaction is 360 +/- 10 s-1 with a binding constant of 196 +/- 12 microM for inorganic pyrophosphate (PPi).(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

16.
1. The disulfide of thioinosine triphosphate, (SnoPPP)2, is a substrate of the Ca2+-pump and the Ca2+-ATPase of sarcoplasmic reticulum (Km = 400 microM). 2. Inactivation of Ca2+-ATPase by the beta,gamma-methylene diphosphonate analogue of the disulfide of thioinosine triphosphate, (SnoPP[CH2]P)2, in the presence of (Ca2+ + Mg2+ + K+) is preceeded by a dissociable enzyme inhibitor complex with a dissociation constant of 130 microM for a low-affinity binding site. ATP protected Ca2+-ATPase against the inactivation under these conditions with a dissociation constant of 140 microM. 3. Kinetic analysis of the inactivations of Ca2+-ATPase by (SnoPP[CH2]P)2 in the absence of Ca2+ and Mg2+ but the presence of K+ and EGTA led to the appearance of two nucleotide binding sites with two different inactivation velocities. Inactivation rate constants k2 were found for the rapid inactivating part (k2' = 1.44 X 10(-2) s-1) and the slow inactivating part (k2" = 1.15 X 10(-3) s-1). From the protective effect of ATP under these conditions a high-affinity (Kd = 48.78 microM) and a low-affinity ATP binding site (Kd = 114 microM) were apparent. 4. The affinity of the analogues to the enzyme is decreased in the sequence: (SnoPPP)2 > (SnoPP[NH]P)2 > (SnoPP[CH2]P)2 > (SnoP)2. 5. (SnoPPP)2-inactivated Ca2+-ATPase was reactivated by incubation with dithiothreitol. 6. Inactivation of Ca2+-ATPase by [gamma-32P](SnoPPP)2 in the presence of (Mg2+ + K+ + Ca2+) or (EGTA + K+) was accompanied by the incorporation of hydroxylamine-insensitive radioactivity into the acid-precipitable protein. The enzyme-bound [gamma-32P]SnoPPP was cleaved by dithiothreitol. 7. It is concluded that (SnoPPP)2 and its non-hydrolyzable analogues (SnoPP[NH]P)2 and (SnoPP[CH2]P)2 act as ATP affinity labels and form mixed disulfides with a sulfhydryl group within the active site.  相似文献   

17.
Using a modified procedure for measuring the time of fibrin clot lysis, the kinetics of Glu- and Lys-plasminogen activation by the tissue activator was studied. Within the plasminogen concentration range of 0.4-100 nM the rate of activation of both protein forms obeys the Michaelis-Menten kinetics. At Lys-plasminogen concentration equimolar to that of fibrin, the rate of activation of the former decreases down to that of Glu-plasminogen activation. The kinetic constants for Glu- and Lys-plasminogen activation (Km) are equal to 0.055 and 0.013 microM; k = 0.19 and 0.21 s-1, respectively. The Km values for fibrin-bound Glu- and Lys-plasminogen are equal to 0.25 nM and 8 nM, respectively (k = 0.08 and 0.26 s-1, respectively). It is assumed that the tissue activator exhibits a higher affinity for the Glu-plasminogen--fibrin complex than for the Lys-plasminogen-fibrin complex.  相似文献   

18.
The kinetics of glucose-fructose oxidoreductase from Zymomonas mobilis   总被引:5,自引:0,他引:5  
Glucose-fructose oxidoreductase operates by a classic ping-pong mechanism with a single site for all substrates: glucose, fructose, gluconolactone and sorbitol. The Km values for these substrates were determined. The values of kcat are 200 s-1 and 0.8 s-1 for the forward and reverse directions respectively. The overall catalytic process consists of two half-reactions with alternate reduction of NADP+ and oxidation of NADPH tightly bound to the enzyme. Reduction of enzyme-NADP+ by glucose and oxidation of enzyme-NADPH by gluconolactone involve single first-order processes. The values of the rate constants at saturating substrate are 2100 s-1 and 8 s-1 respectively; deuterium isotope effects indicate that these are for the hydrogen transfer step. Oxidation of enzyme-NADPH by fructose is first order with a limiting rate constant of at least 430 s-1. The reaction of enzyme-NADP+ with sorbitol is biphasic, with rate constants for both phases less than 1 s-1. This behaviour is explained by a mechanism in which the slow cyclisation of the acyclic form of fructose follows its dissociation from the enzyme. The rate-determining steps for the overall reaction are probably dissociation of gluconolactone in the forward direction and hydrogen transfer from sorbitol to enzyme-bound NADP+ in the reverse direction.  相似文献   

19.
The reduction kinetics of both the resting and redox-cycled forms of the nitrite reductase from the anaerobic rumen bacterium Wolinella succinogenes were studied by stopped-flow reaction techniques. Single-turnover reduction of the enzyme by dithionite occurs in two kinetic phases for both forms of the enzyme. When the resting form of the enzyme is subjected to a single-turnover reduction by dithionite, the slower of the two kinetic phases exhibits a hyperbolic dependence of the rate constant on the square root of the reductant concentration, the limiting value of which (approximately 4 s-1) is assigned to a slow internal electron-transfer process. In contrast, when the redox-cycled form of the enzyme is reduced by dithionite in a single-turnover experiment, both kinetic phases exhibit linear dependences of the rate on the square root of dithionite concentration, with associated rate constants of 150 M-1/2.s-1 and 6 M-1/2.s-1. Computer simulations of both the reduction processes shows that no unique set of rate constants can account for the kinetics of both forms, although the kinetics of the redox-cycled species is consistent with a much enhanced rate of internal electron transfer. Under turnover conditions the time course for reduction of the enzyme, in the presence of millimolar levels of nitrite and 100 mM-dithionite, is extremely complex. A working model for the mechanism of the turnover activity of the enzyme is proposed which very closely describes the reaction kinetics over a wide range of substrate concentrations, as shown by computer simulation. The similarity in the action of the nitrite reductase enzyme and mammalian cytochrome c oxidase is commented upon.  相似文献   

20.
Phosphoenolpyruvate carboxylase [EC 4.1.1.31] from Escherichia coli W was alkylated by incubation with bromopyruvate, substrate analog, leading to irreversible inactivation. The reaction followed pseudo-first-order kinetics. Mg2+, an essential cofactor for catalysis, enhanced the inactivation, and the enhancing effect increased as the pH increased. The inactivation rate showed a tendency to saturate with increasing concentrations of bromopyruvate, indicating that an enzyme-bromopyruvate complex was formed prior to the alkylation. DL-Phospholactate, a potent competitive inhibitor with respect to phosphoenolpyruvate, protected the enzyme from inactivation in a competitive manner. Examination of the acid hydrolysate of the enzyme modified with [14C]bromopyruvate by paper chromatography showed that radioactivity was solely incorporated into carboxyhydroxyethyl cysteine. In addition, determination of sulfhydryl groups of the native and modified enzymes with 5,5'-dithiobis(2-nitrobenzoate) showed that inactivation occurred concomitant with the modification of one cysteinyl residue per subunit. The results indicate that bromopyruvate reacted with the enzyme as an active-site-directed reagent.  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