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.
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   

2.
The displacement of NADH from cytoplasmic aldehyde dehydrogenase (EC 1.2.1.3) from sheep liver was studied by using NAD+, 1,10-phenanthroline, ADP-ribose, deamino-NAD+ and pyridine-3-aldehyde-adenine dinucleotide as displacing agents, by following the decrease in fluorescence as a function of time. The data obtained could be fitted by assuming two first-order processes were occurring, a faster process with an apparent rate constant of 0.85 +/- 0.20 s-1 and a relative amplitude of 60 +/- 10% and a slower process with an apparent rate constant of 0.20 +/- 0.05 s-1 and a relative amplitude of 40 +/- 10% (except for pyridine-3-aldehyde-adenine dinucleotide, where the apparent rate constant for the slow process was 0.05 s-1). The displacement rates did not change significantly when the pH was varied from 6.0 to 9.0. Kinetic data are also reported for the dependence of the rate of binding of NADH to the enzyme on the total concentration of NADH. Detailed arguments are presented based on the isolation and purification procedures, the equilibrium coenzyme-binding studies and the kinetic data, which lead to the following model for the release of NADH from the enzyme: (formula: see article). The parameters that best fit the data are: k + 1 = 0.2 s-1; k - 1 = 0.05 s-1; k + 2 = 0.8 s-1 and k - 2 = 5 X 10(5)litre-mol-1-s-1. The slow phase of the NADH release is similar to the steady-state turnover number for substrates such as acetaldehyde and propionaldehyde and appears to contribute significantly to the limitation of the steady-state rate.  相似文献   

3.
J W Burgner  W J Ray 《Biochemistry》1984,23(16):3620-3626
The binary complex of NAD and lactate dehydrogenase reacts reversibly with cyanide to produce a complex (E X NAD-CN) whose noncovalent interactions are similar to those in the E X NADH complex (where E is one-fourth of the tetrameric dehydrogenase). The reaction apparently is a simple bimolecular nucleophilic addition at the 4 position of the bound nicotinamide ring; viz., cyanide does not bind to the enzyme prior to reaction. The value of the dissociation constant for E X NAD-CN is about 1 X 10(-6) M and is independent of pH over the range of 6-8. The equilibrium constant for the reaction of cyanide with E X NAD is about 400-fold larger than that for the nonenzymic process after a statistical correction. This increment in Ke is accounted for by a 220-fold increase in the rate of the forward enzymic reaction (20 M-1 s-1) as compared with an approximately 2-fold decrease for the reverse process (9 X 10(-5) s-1). Thus, the increased value of the rate constant for bond formation in the enzymic reaction is attributed to an equilibrium binding effect that is translated almost entirely into a rate effect on that step (bond formation). Since the nonenzymic reaction is sensitive to solvent composition, this equilibrium binding effect likely is produced by environmental effects at the nicotinamide/dehydronicotinamide part of the coenzyme binding site on the enzyme.  相似文献   

4.
S-Adenosylhomocysteine hydrolase (SAHase) was resolved into apoenzyme and NAD+ by acidic ammonium sulfate treatment. The apoenzyme was catalytically inactive, but could be reconstituted to active enzyme with NAD+. Reduced SAHase (ENADH) that was prepared by reconstitution of the apoenzyme with NADH was catalytically inactive. ENADH was oxidized by 3'-ketoadenosine to active SAHase. The recovery of activity paralleled the oxidation of enzyme-bound NADH. The association rate constant for ENADH and 3'-ketoadenosine was 6.1 x 10(2) M-1 s-1, and the dissociation rate constant was calculated to be 4 x 10(-7) s-1. This association rate constant was considerably smaller than the association rate constant for adenosine and SAHase (greater than 10(7) M-1 s-1). However, the observed pseudo first-order rate constant for reaction of 3'-ketoadenosine with ENADH (0.6 s-1 with 1 mM 3'-ketoadenosine) approached kcat for the hydrolytic reaction (1.2 s-1). Thus, bound 3'-ketoadenosine probably reacted sufficiently rapidly with ENADH to be considered a kinetically competent intermediate. The dissociation constants of SAHase for adenosine and 4',5'-dehydroadenosine, substrates for the enzyme, were 9 and 14 microM, respectively. In contrast, the dissociation constants of ENADH for 3'-ketoadenosine and 4',5'-dehydro-3'-ketoadenosine, intermediates of the catalytic reaction, were significantly lower with values of 600 and 300 pM, respectively. The equilibrium constant for reduction of enzyme-bound NAD+ in the absence of an adenosine analogue, as estimated from cyanide binding studies, was 10-fold more favorable than that for free NAD+. ENADH was highly fluorescent (emission maximum 428 nm, excitation 340 nm) with a quantum yield that was six times that of free NADH. Since SAHase reduced by adenosine was not highly fluorescent, enzyme-bound intermediates quenched the fluorescence of enzyme-bound NADH. Adenosine and adenine quenched the fluorescence of ENADH. Cyanide formed a complex with SAHase that was analogous to ENADH. Adenine stabilized this complex sufficiently that addition of 65 microM adenine and 25 mM cyanide to SAHase caused total complex formation with loss of over 95% of the catalytic activity.  相似文献   

5.
Kinetic studies were carried out on mitochondrial aldehyde dehydrogenase (EC 1.2.1.3) isolated from sheep liver. Steady-state studies over a wide range of acetaldehyde concentrations gave a non-linear double-reciprocal plot. The dissociation of NADH from the enzyme was a biphasic process with decay constants 0.6s-1 and 0.09s-1. Pre-steady-state kinetic data with propionaldehyde as substrate could be fitted by using the same burst rate constant (12 +/- 3s-1) over a wide range of propionaldehyde concentrations. The quenching of protein fluorescence on the binding of NAD+ to the enzyme was used to estimate apparent rate constants for binding (2 X 10(4) litre.mol-1.s-1) and dissociation (4s-1). The kinetic properties of the mitochondrial enzyme, compared with those reported for the cytoplasmic aldehyde dehydrogenase from sheep liver, show significant differences, which may be important in the oxidation of aldehydes in vivo.  相似文献   

6.
The kcat value for the oxidation of propionaldehyde by sheep liver cytosolic aldehyde dehydrogenase increased 3-fold, from 0.16 s-1 at pH 7.6 to 0.49 s-1 at pH 5.2, in parallel with the increase in the rate of displacement of NADH from binary enzyme.NADH complexes. A burst in nucleotide fluorescence was observed at all pH values consistent with the rate of isomerization of binary enzyme.NADH complexes constituting the rate-limiting step in the steady state. No substrate activation by propionaldehyde was observed at pH 5.2, but the enzyme exhibited dissociation/association behavior. The inactive dissociated form of the enzyme was favored by low enzyme concentration, low pH, and low ionic strength. Propionaldehyde protected the enzyme against dissociation.  相似文献   

7.
The biophysical properties of a tryptophan-shifted mutant of phosphofructokinase from Bacillus stearothermophilus (BsPFK) have been examined. The mutant, designated W179Y/Y164W, has kinetic and thermodynamic properties similar to the wild-type enzyme. A 2-fold decrease in kcat is observed, and the mutant displays a 3-fold smaller K(0.5) for the substrate, fructose-6-phosphate (Fru-6-P), as compared to the wild-type enzyme. The dissociation constant for the inhibitor, phospho(enol)pyruvate (PEP), increases 2-fold, and the coupling parameter, Q(ay), decreases 2-fold. This suggests that while the mutant displays a slightly decreased affinity for PEP, PEP is still an effective inhibitor once bound. The new position of the tryptophan in W179Y/Y164W is approximately 6 A from the Fru-6-P portion of the active site. A 25% decrease in fluorescence intensity is observed upon Fru-6-P binding, and an 80% decrease in fluorescence intensity is observed with PEP binding. In addition, the intrinsic fluorescence polarization increases from 0.327 +/- 0.001 to 0.353 +/- 0.001 upon Fru-6-P binding, but decreases to 0.290 +/- 0.001 when PEP binds. Most notably, the presence of PEP induces dissociation of the tetramer. Dissociation of the tetramer into dimers occurs along the active site interface and can be monitored by the loss in activity or the loss in tryptophan fluorescence that is observed when the enzyme is titrated with PEP. Activity can be protected or recovered by incubating the enzyme with Fru-6-P. Recovery of activity is enzyme concentration dependent, and the rate constant for association is 6.2 +/- 0.3 M(-1) x s(-1). Ultracentrifugation experiments revealed that in the absence of PEP the mutant enzyme exists in an equilibrium between the dimer and tetramer forms with a dissociation constant of 11.8 +/- 0.5 microM, while in the presence of PEP the enzyme exists in equilibrium between the dimer and monomer forms with a dissociation constant of 7.5 +/- 0.02 microM. A 3.1 A crystal structure of the mutant enzyme suggests that the amino acid substitutions have not dramatically altered the tertiary structure of the enzyme. While it is clear that wild-type BsPFK exists as a tetramer under these same conditions, these results suggest that quaternary structural changes probably play an important role in allosteric communication.  相似文献   

8.
Cyanide binding to prostaglandin H (PGH) synthase results in a spectral shift in the Soret region. This shift was exploited to determine equilibrium and kinetic parameters of the cyanide binding process. At pH 8.0, ionic strength 0.22 M, 4 degrees C, the cyanide dissociation constant, determined from equilibrium experiments, is (65 +/- 10) microM. The binding rate constant is (2.8 +/- 0.2) x 10(3) M-1 s-1, and the dissociation rate constant is zero within experimental error. Through a kinetic study of the binding process as a function of pH, from pH 3.96 to 8.00, it was possible to determine the pKa of a heme-linked acid group on the enzyme of 4.15 +/- 0.10 with citrate buffer. An apparent pKa of 4.75 +/- 0.03 was determined with acetate buffer; this different value is attributed to complexation of the enzyme with one of the components of the acetate buffer.  相似文献   

9.
Human estrogenic 17beta-hydroxysteroid dehydrogenase (17beta-HSD1) plays a crucial role in the last step of the synthesis of estrogens. A detailed kinetic study demonstrated that the enzyme shows about 240 fold higher specificity towards estrone reduction than estradiol oxidation at physiological pH using tri-phosphate cofactors. The kcat/Km values are 96 +/- 10 and 0.4 +/- 0.1 s-1 (microM)-1 respectively for the above two reactions. However, it has been shown that this difference is closely linked to the use of NADPH and NADP cofactors. A binding study using equilibrium dialysis indicated similar KD (equilibrium dissociation constant) of 11 +/- 1 and 4.7 +/- 0.9 microM for estrone and estradiol, respectively. The binding affinity of 17beta-HSD1 to estrone was significantly increased with a KD of 1.6 +/- 0.2 microM in the presence of NADP, the latter used as an analogue of the NADPH. The results of binding studies agree with the steady-state kinetics, which showed that the Km of estrone is 12-fold lower when using NADPH as a cofactor than when using NADH. These results strongly suggest that the cofactor plays a crucial role in the stimulation of the specificity for estrogen reduction.  相似文献   

10.
The binding of substrates and the herbicide N-(phosphonomethyl)glycine (glyphosate) to enolpyruvoylshikimate-3-phosphate (EPSP) synthase was evaluated by stopped-flow and equilibrium fluorescence measurements. Changes in protein fluorescence were observed upon the binding of EPSP and upon the formation of the enzyme-shikimate 3-phosphate-glyphosate ternary complex; no change was seen with either shikimate 3-phosphate (S3P) or glyphosate alone. By fluorescence titrations, the dissociation constants were determined for the formation of the enzyme binary complexes with S3P (Kd,S = 7 +/- 1.2 microM) and EPSP (Kd,EPSP = 1 +/- 0.01 microM). The dissociation constant for S3P was determined by competition with EPSP or by measurements in the presence of a low glyphosate concentration. At saturating concentrations of S3P, glyphosate bound to the enzyme--S3P binary complex with a dissociation constant of 0.16 +/- 0.02 microM. Glyphosate did not bind significantly to free enzyme, so the binding is ordered with S3P binding first: (formula; see text) where S refers to S3P, G refers to glyphosate, and E.S.G. represents the complex with altered fluorescence. The kinetics of binding were measured by stopped-flow fluorescence methods. The rate of glyphosate binding to the enzyme--S3P complex was k2 = (7.8 +/- 0.2) X 10(5) M-1 s-1, from which we calculated the dissociation rate k-2 = 0.12 +/- 0.02 s-1.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

11.
L J Maher  P B Dervan  B J Wold 《Biochemistry》1990,29(37):8820-8826
Pyrimidine oligonucleotides recognize extended purine sequences in the major groove of double-helical DNA by triple-helix formation. The resulting local triple helices are relatively stable and can block DNA recognition by sequence-specific DNA binding proteins such as restriction endonucleases. Association and dissociation kinetics for the oligodeoxyribonucleotide 5'-CTCTTTCCTCTCTTTTTCCCC (bold C's indicate 5-methylcytosine residues) are now measured with a restriction endonuclease protection assay. When oligonucleotides are present in greater than 10-fold excess over the DNA target site, the binding reaction kinetics are pseudo first order in oligonucleotide concentration. Under our standard conditions (37 degrees C, 25 mM Tris-acetate, pH 6.8, 70 mM sodium chloride, 20 mM magnesium chloride, 0.4 mM spermine tetrahydrochloride, 10 mM beta-mercaptoethanol, 0.1 mg/mL bovine serum albumin) the value of the observed pseudo-first-order association rate constant, k2obs, is 1.8 x 10(3) +/- 1.9 x 10(2) L.(mol of oligomer-1.s-1. Measurement of the dissociation rate constant yields an equilibrium dissociation constant of approximately 10 nM. Increasing sodium ion concentration slightly decreased the association rate, substantially increased the dissociation rate, and thereby reduced the equilibrium binding constant. This effect was reversible by increasing multivalent cation concentration, confirming the significant role of multivalent cations in oligonucleotide-directed triple-helix formation under these conditions. Finally, a small reduction in association rate, a large increase in dissociation rate, and a resulting reduction in the equilibrium binding constant were observed upon increasing the pH between 6.8 and 7.2.  相似文献   

12.
Purified NAD-malic enzyme from Ascaris suum is rapidly inactivated by the arginine reagent, 2,3-butanedione, and this inactivation is facilitated by 30 mM borate. Determination of the inactivation rate as a function of butanedione concentration suggests a second-order process overall, which is first order in butanedione. A second-order rate constant of 0.6 M-1 s-1 at pH 9 is obtained for the butanedione reaction. The inactivation is reversed by removal of the excess reagent upon dialysis. The enzyme is protected against inactivation by saturating amounts of malate in the presence and absence of borate. The divalent metal Mg2+ affords protection in the presence of borate but has no effect in its absence. The nucleotide reactant NAD+ has no effect on the inactivation rate in either the presence or absence of borate. A dissociation constant of 24 mM is obtained for E:malate from the decrease in the inactivation rate as a function of malate concentration. An apparent Ki of 0.5 mM is obtained for oxalate (an inhibitor competitive vs malate) from E:Mg:oxalate while no significant binding is observed for oxalate using the butanedione modified enzyme. The pH dependence of the first-order rate of inactivation by butanedione gives a pKa of 9.4 +/- 0.1 for the residue(s) modified, and this pK is increased when NAD is bound. The arginine(s) modified is implicated in the binding of malate.  相似文献   

13.
J A Cognet  B G Cox  G G Hammes 《Biochemistry》1983,22(26):6281-6287
The kinetics of reduced nicotinamide adenine dinucleotide phosphate (NADPH) binding to fatty acid synthase from chicken liver and of the reduction of enzyme-bound acetoacetyl by NADPH (beta-ketoacyl reductase) and the steps leading to formation of the acetoacetyl-enzyme have been studied in 0.1 M potassium phosphate-1 mM ethylenediaminetetraacetic acid (EDTA), pH 7.0, at 25 degrees C by monitoring changes in NADPH fluorescence with a stopped-flow apparatus. Improved fluorescence detection has permitted the use of NADPH concentrations as low as 20 nM. The kinetics of the binding of NADPH to the enzyme is consistent with a simple bimolecular binding mechanism and four equivalent sites on the enzyme (presumably two beta-ketoacyl reductase sites and two enoyl reductase sites). The bimolecular rate constant is 12.7 X 10(6) M-1 s-1, and the dissociation rate constant is 76.7 s-1, which gives an equilibrium dissociation constant of 6.0 microM. The formation of the acetoacetyl-enzyme and its subsequent reduction by NADPH could be analyzed as two consecutive pseudo-first-order reactions by mixing enzyme-NADPH with acetyl-CoA and malonyl-CoA under conditions where [acetyl-CoA], [malonyl-CoA] much greater than [enzyme] much greater than [NADPH]. From the dependence of the rate of reduction of aceto-acetyl-enzyme by NADPH on enzyme concentration, an independent estimate of the equilibrium dissociation constant for NADPH binding to the enzyme of 5.9 microM is obtained, and the rate constant for the reduction is 17.5 s-1.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

14.
1. Supernatant pig heart malate dehydrogenase is completely inhibited by reaction with diethyl pyrocarbonate at pH6.5, when 0.58+/-0.1 residue of ethoxycarbonylhistidine is formed per NADH-binding site. 2. Oxaloacetate and hydroxymalonate protect the enzyme from inhibition in the absence of coenzyme. 3. Limited ethoxycarbonylation does not alter the binding of NADH to the enzyme but prevents the enzyme-NADH complex from interacting with hydroxymalonate in a ternary complex.  相似文献   

15.
Phalloidin enhances actin assembly by preventing monomer dissociation   总被引:20,自引:11,他引:9       下载免费PDF全文
Incubation of the isolated acrosomal bundles of Limulus sperm with skeletal muscle actin results in assembly of actin onto both ends of the bundles. These cross-linked bundles of actin filaments taper, thus allowing one to distinguish directly the preferred end for actin assembly from the nonpreferred end; the preferred end is thinner. Incubation with actin in the presence of equimolar phalloidin in 100 mM KCl, 1 mM MgCl2 and 0.5 mM ATP at pH 7.5 resulted in a slightly smaller association rate constant at the preferred end than in the absence of the drug (3.36 +/- 0.14 X 10(6) M-1 s-1 vs. 2.63 +/- 0.22 X 10(6) M-1 s- 1, control vs. experimental). In the presence of phalloidin, the dissociation rate constant at the preferred end was reduced from 0.317 +/- 0.097 s-1 to essentially zero. Consequently, the critical concentration at the preferred end dropped from 0.10 microM to zero in the presence of the drug. There was no detectable change in the rate constant of association at the nonpreferred end in the presence of phalloidin (0.256 +/- 0.015 X 10(6) M-1 s-1 vs. 0.256 +/- 0.043 X 10(6) M-1 s-1, control vs. experimental); however, the dissociation rate constant was reduced from 0.269 +/- 0.043 s-1 to essentially zero. Thus, the critical concentration at the nonpreferred end changed from 1.02 microM to zero in the presence of phalloidin. Dilution-induced depolymerization at both the preferred and nonpreferred ends was prevented in the presence of phalloidin. Thus, phalloidin enhances actin assembly by lowering the critical concentration at both ends of actin filaments, a consequence of reducing the dissociation rate constants at each end.  相似文献   

16.
Kinetics and mechanism of bilirubin binding to human serum albumin   总被引:3,自引:0,他引:3  
The kinetics of bilirubin binding to human serum albumin at pH 7.40, 4 degrees C, was studied by monitoring changes in bilirubin absorbance. The time course of the absorbance change at 380 nm was complex: at least three kinetic events were detected including the bimolecular association (k1 = 3.8 +/- 2.0 X 10(7) M-1 S-1) and two relaxation steps (52 = 40.2 +/- 9.4 s-1 and k3 = 3.8 +/- 0.5 s-1). The presence of the two slow relaxations was confirmed under pseudo-first order conditions with excess albumin. Curve-fitting procedures allowed the assignment of absorption coefficients to the intermediate species. When the bilirubin-albumin binding kinetics was observed at 420 nm, only the two relaxations were seen; apparently the second order association step was isosbestic at this wavelength. The rate of albumin-bound bilirubin dissociation was measured by mixing the pre-equilibrated human albumin-bilirubin complex with bovine albumin. The rate constant for bilirubin dissociation measured at 485 nm was k-3 = 0.01 s-1 at 4 degrees C. A minimum value of the equilibrium constant for bilirubin binding to human albumin determined from the ratio k1/k-3 is therefore approximately 4 X 10(9) M-1.  相似文献   

17.
The kinetic mechanism of homoisocitrate dehydrogenase from Saccharomyces cerevisiae was determined using initial velocity studies in the absence and presence of product and dead end inhibitors in both reaction directions. Data suggest a steady state random kinetic mechanism. The dissociation constant of the Mg-homoisocitrate complex (MgHIc) was estimated to be 11 +/- 2 mM as measured using Mg2+ as a shift reagent. Initial velocity data indicate the MgHIc complex is the reactant in the direction of oxidative decarboxylation, while in the reverse reaction direction, the enzyme likely binds uncomplexed Mg2+ and alpha-ketoadipate. Curvature is observed in the double-reciprocal plots for product inhibition by NADH and the dead-end inhibition by 3-acetylpyridine adenine dinucleotide phosphate when MgHIc is the varied substrate. At low concentrations of MgHIc, the inhibition by both nucleotides is competitive, but as the MgHIc concentration increases, the inhibition changes to uncompetitive, consistent with a steady state random mechanism with preferred binding of MgHIc before NAD. Release of product is preferred and ordered with respect to CO2, alpha-ketoadipate, and NADH. Isocitrate is a slow substrate with a rate (V/E(t)) 216-fold slower than that measured with HIc. In contrast to HIc, the uncomplexed form of isocitrate and Mg2+ bind to the enzyme. The kinetic mechanism in the direction of oxidative decarboxylation of isocitrate, on the basis of initial velocity studies in the absence and presence of dead-end inhibitors, suggests random addition of NAD and isocitrate with Mg2+ binding before isocitrate in rapid equilibrium, and the mechanism approximates rapid equilibrium random. The Keq for the overall reaction measured directly using the change in NADH as a probe is 0.45 M.  相似文献   

18.
Sheep liver cytoplasmic aldehyde dehydrogenase was purified to homogeneity to give a sample with a specific activity of 380 nmol NADH min(-1) mg(-1). An amino acid analysis of the enzyme gave results similar to those reported for aldehyde dehydrogenases from other sources. The isoelectric point was at pH 5.25 and the enzyme contained no significant amounts of metal ions. On the binding of NADH to the enzyme there is a shift in absorption maximum of NADH to 344 nm, and a 5.6-fold enhancement of nucleotide fluorescence. The protein fluorescence (lambdaexcit = 290 nm, lambdaemisson = 340 nm) is quenched on the binding of NAD+ and NADH. The enhancement of nucleotide fluorescence on the binding of NADH has been utilised to determine the dissociation constant for the enzyme . NADH complex (Kd = 1.2 +/- 0.2 muM). A Hill plot of the data gave a straight line with a slope of 1.0 +/- 0.3 indicating the absence of co-operative effects. Ellman's reagent reacted only slowly with the enzyme but in the presence of sodium dodecylsulphate complete reaction occurred within a few minutes to an extent corresponding to 36 thiol groups/enzyme. Molecular weights were determined for both cytoplasmic and mitochondrial aldehyde dehydrogenases and were 212 000 +/- 8 000 and 205 000 respectively. Each enzyme consisted of four subunits with molecular weight of 53 000 +/- 2 000. Properties of the cytoplasmic and mitochondrial aldehyde dehydrogenases from sheep liver were compared with other mammalian liver aldehyde dehydrogenases.  相似文献   

19.
1. The rate constants for NADH binding and dissociation for carboxymethylated alcohol dehydrogenase have been determined and compared to those for the native enzyme. 2. Steady-state and transient kinetic experiments have shown that the hydrogen transfer step is rate-determining for oxidation of ethanol by carboxymethylated alcohol dehydrogenase. The rate constant of 0.19 s-1 is considerably slower than that for the native enzyme. 3. The steady-state parameter, V/[E], was obtained for each of a series of alcohols and correlated with the Taft sigma parameter. The linear relationship obtained indicates that the same step, hydrogen transfer, is rate-determining for all the alcohols. The sigma value obtained is the same as for the native enzyme; the implications of this for the mechanism of hydrogen transfer are discussed.  相似文献   

20.
Equilibria and kinetics of cyanide binding to canine myeloperoxidase were studied. Spectral results support the presence of two heme binding sites; an isosbestic point at 444 nm and a linear Scatchard plot suggest that the binding affinity of cyanide to the two subunits of the enzyme is the same. The dissociation constant is 0.53 microM. The pH dependence of the apparent second order rate constant indicates the presence of an acid-base group on the enzyme with a pKa of 3.8 +/- 0.1. The protonated form of cyanide binds to the basic enzyme with a rate constant of (4.3 +/- 0.3) x 10(6) M-1 s-1.  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