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 kinetic mechanism of betaine aldehyde dehydrogenase from leaves of the plant Amaranthus hypochondriacus is ordered with NAD(+) adding first. NADH is a noncompetitive inhibitor against NAD(+), which was interpreted before as evidence of an iso mechanism, in which NAD(+) and NADH binds to different forms of free enzyme. With the aim of testing the proposed kinetic mechanism, we have now investigated the ability of NADH to form different complexes with the enzyme. By initial velocity and equilibrium binding studies, we found that the steady-state levels of E.glycine betaine are negligible, ruling out binding of NADH to this complex. However, NADH readily bind to E.betaine aldehyde, whose levels most likely are kinetically significant given its low dissociation constant. Also, NADH combined with E.NADH and E.NAD(+). Finally, NADH was not able to revert the hydride transfer step, what suggest that there is no acyl-enzyme intermediate, i.e. the release of the reduced dinucleotide takes place after the deacylation step. Although formation of the complex E.NAD(+).NADH would produce an uncompetitive effect in the inhibition of NADH against NAD(+), the iso mechanism cannot be conclusively discarded.  相似文献   

2.
The kinetic mechanism of betaine aldehyde dehydrogenase from leaves of the plant Amaranthus hypochondriacus is ordered with NAD+ adding first. NADH is a noncompetitive inhibitor against NAD+, which was interpreted before as evidence of an iso mechanism, in which NAD+ and NADH binds to different forms of free enzyme. With the aim of testing the proposed kinetic mechanism, we have now investigated the ability of NADH to form different complexes with the enzyme. By initial velocity and equilibrium binding studies, we found that the steady-state levels of E.glycine betaine are negligible, ruling out binding of NADH to this complex. However, NADH readily bind to E.betaine aldehyde, whose levels most likely are kinetically significant given its low dissociation constant. Also, NADH combined with E.NADH and E.NAD+. Finally, NADH was not able to revert the hydride transfer step, what suggest that there is no acyl-enzyme intermediate, i.e. the release of the reduced dinucleotide takes place after the deacylation step. Although formation of the complex E.NAD+.NADH would produce an uncompetitive effect in the inhibition of NADH against NAD+, the iso mechanism cannot be conclusively discarded.  相似文献   

3.
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   

4.
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   

5.
The binding of NADH to porcine mitochondrial malate dehydrogenase in phosphate buffer at pH 7.5 has been studied by equilibrium and kinetic methods. Hyperbolic binding was obtained by fluorimetric titration of enzyme with NADH, in the presence or absence of hydroxymalonate. Identical results were obtained for titrations of NADH with enzyme in the presence or absence of hydroxymalonate, measured either by fluorescence emission intensity or by the product of intensity and anisotropy. The equilibrium constant for NADH dissociation was 3.8 +/- 0.2 micrometers, over a 23-fold range of enzyme concentration, and the value in the presence of saturating hydroxymalonate was 0.33 +/- 0.02 micrometer over a 10-fold range of enzyme concentration. The rate constant for NADH binding to the enzyme in the presence of hydroxymalonate was 3.6 X 10(7) M-1 s-1, while the value for dissociation from the ternary complex was 30 +/- 1 s-1. No limiting binding rate was obtained at pseudo-first order rate constants as high as 200 s-1, and the rate curve for dissociation was a single exponential for at least 98% of the amplitude. In addition to demonstrating that the binding sites are independent and indistinguishable, the absence of effects of enzyme concentration on the KD value indicates that NADH binds with equal affinity to monomeric and dimeric enzyme forms.  相似文献   

6.
Aldehyde binding to liver alcohol dehydrogenase in the absence and presence of coenzymes has been characterized by spectrometric equilibrium methods, using auramine O and bipyridine as reporter ligands. Free enzyme shows a significant affinity for aldehydes, and equilibrium constants for dissociation of the binary complexes formed with typical aldehyde substrates are reported. Binary-complex formation does not lead to any detectable inner-sphere coordination of aldehydes to the catalytic zinc ion of the enzyme subunit. Complex formation with NAD+ or NADH increases the affinity of the enzyme for aromatic aldehydes by a factor of 1.8 - 3.5 and 6-17, respectively. Benzaldehyde and dimethylaminocinnamaldehyde binding to the enzyme . NAD+ complex is not detectably associated with inner-sphere coordination of the aldehyde to zinc. It is concluded that binding of NADH is required to induce catalytically adequate bonding interactions between enzyme and aromatic aldehydes. The effect of reduced coenzyme in this respect is attributed to hydrophobic interactions leading to dehydration of the active-site region, which allows aldehyde substrates to compete successfully with water for inner-sphere coordination to the catalytic zinc ion. Oxidized coenzyme is proposed to have a similar promoting effect on metal coordination of aldehydes which function as substrates for the dismutase activity of the enzyme.  相似文献   

7.
Cytoplasmic aldehyde dehydrogenase may be modified by reaction with p-nitrophenyl dimethylcarbamate to give a stable E-X-CO-NMe2 species that is an analogue of the usual labile acyl-enzyme involved in the enzyme's reactions. [X is derived from an enzymic nucleophilic group.] This species still contains the tightly bound NADH that is present in the native enzyme. When further NADH binds to E-X-CO-NMe2 its fluorescence is enhanced over 4 times more than when it binds to unmodified enzyme; this fluorescence is completely unaffected by high propionaldehyde concentration and only slightly affected by p-nitrobenzaldehyde. The modified species has 1.0 NADH binding site in the absence of Mg2+ and 1.67 sites in its presence. The rate of dissociation of E-X-CO-NMe2.NADH is biphasic (k 3.4 and 1.8 min-1) and is considerably lower than that of E.NADH; the presence of Mg2+ slows the process even more (k 0.47 and 0.37 min-1). The implications of these studies towards a greater understanding of the nature of aldehyde dehydrogenase-catalysed reactions are discussed.  相似文献   

8.
The binding of oxidized and reduced coenzyme (NAD+ and NADH) to 3-phosphoglyceroyl-glyceraldehyde-3-phosphate dehydrogenase has been studied spectrophotometrically and fluorimetrically. The binding of NAD+ to the acylated sturgeon enzyme is characterized by a significant quenching of the enzyme fluorescence (about 25%) and the induction of a difference spectrum in the ultraviolet absorbance region of the enzyme. Both of these spectroscopic properties are quantitatively distinguishable from those of the corresponding binary enzyme-NAD+ complex. Binding isotherms estimated by gel filtration of the acylated enzyme are in close agreement to those obtained by spectrophotometric and fluorimetric titrations. Up to four NAD+ molecules are bound to the enzyme tetramer. No anticooperativity can be detected in the binding of oxidized coenzyme, which is well described on the basis of a single class of four binding sites with a dissociation constant of 25 muM at 10 degrees C, pH 7.0. The binding of NADH to the acylenzyme has been characterized spectrophotometrically. The absorption band of the dihydronicotinamide moiety of the coenzyme is blue-shifted to 335 nm with respect to free NADH. In addition, a large hypochromicity (23%) is observed together with a significant increase of the bandwidth at half height of this absorption band. This last property is specific to the acylenzyme-DADH complex, since it disappears upon arsenolysis of the acylenzyme. The binding affinity of NADH to the acylated enzyme has been estimated by performing simultaneous spectrophotometric and fluorimetric titrations of the NADH appearance upon addition of NAD+ to a mixture of enzyme and excess glyceraldehyde 3-phosphate. In contrast to NAD+, the reduced coenzyme NADH appears to be relatively strongly bound to the acylated enzyme, the dissociation constant of the acylenzyme-NADH complex being estimated as 2.0 muM at 25 degrees C. In addition a large quenching of the NADH fluorescence (about 83%) is observed. The comparison of the dissociation constants of the coenzyme-acylenzyme complexes and the corresponding Michaelis constants suggests a reaction mechanism of the enzyme in which significant formation and dissociation of NAD+-acylenzyme and NADH-acylenzyme complexes occur. Under physiological conditions the activity of the enzyme can be regulated by the ratio of oxidized and reduced coenzymes. Possible reasons for the lack of anticooperativity in coenzyme binding to the acylated form of the enzyme are discussed.  相似文献   

9.
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   

10.
The direct transfer of NADH between A-B pairs of dehydrogenases and also the dissociation of NADH from individual E-NADH complexes have been investigated by transient stopped-flow kinetic techniques. Such A-B transfers of NADH occur without the intermediate dissociation of coenzyme into the aqueous solvent environment [Srivastava, D.K., & Bernhard, S.A. (1985) Biochemistry 24, 623-628]. The equilibrium distributions of limiting NADH among aqueous solvent and A and B dehydrogenase sites have also been determined. At sufficiently high but realizable concentrations of dehydrogenases, both the transfer rate and the equilibrium distribution of bound NADH are virtually independent of the excessive enzyme concentrations; at excessive E2 concentration, substantial NADH is bound to the E1 site. These results further substantiate earlier kinetic arguments for the preferential formation of an EA-NADH-EB complex, within which coenzyme is directly transferred between sites. The unimolecular specific rates of coenzyme transfer from site to site are nearly invariant among different A-B dehydrogenase pairs. The equilibrium constants for the distribution of coenzyme within the EA X EB complexes are near unity. At high [E2] and for [E2] greater than [E1] greater than [NADH], E1-NADH X E2 and E1 X NADH-E2 are virtually the only coenzyme-contained species. In contrast to the nearly invariant unimolecular NADH transfer rates within EA X EB complexes, unimolecular specific rates of dissociation of NADH from E-NADH into aqueous solution are highly variable.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

11.
1. The reaction of pig heart lactate dehydrogenase (EC 1.1.1.27) with NAD(+) and lactate to form pyruvate and NADH was followed by rapid spectrophotometric methods. The distinct spectrum of enzyme-bound NADH permits the measurement of the rate of dissociation of this compound. 2. The reduction of the first mole equivalent of NAD(+) per mole of enzyme sites can also be observed, and is much more rapid than the steady-state rate of NADH production. 3. At pH8 the dissociation of the enzyme-NADH complex is rate-determining for the steady-state oxidation of lactate. At lower pH some other step after the interconversion of the ternary complex and before the dissociation of NADH is rate-determining. Other evidence for a compulsory-order mechanism is provided.  相似文献   

12.
Transient-kinetic studies of pig muscle lactate dehydrogenase   总被引:9,自引:8,他引:1  
1. The very fast pre-steady-state formation of NADH catalysed by pig M(4) lactate dehydrogenase was equivalent to the enzyme-site concentration at pH values greater than 8.0 and to one-half the site concentration at pH6.8. 2. The rate of dissociation of NADH from the enzyme at pH8.0 (450s(-1)) in the absence of other substrates is faster than the steady-state oxidation of lactate (80s(-1)). The latter process is therefore controlled by a step before NADH dissociation but subsequent to the hydride transfer. 3. The oxidation of enzyme-NADH by excess of pyruvate was studied as a first-order process at pH9.0. There was no effect of NADD on this reaction and it was concluded that the ternary complex undergoes a rate-limiting change before the hydride-transfer step. 4. Some conclusions about the reactions catalysed by the M(4) isoenzyme were drawn from a comparison of these results with those obtained with the H(4) isoenzyme and liver alcohol dehydrogenase.  相似文献   

13.
We have examined aspects of the second catalytic activity of alcohol dehydrogenase from horse liver (LADH), which involves an apparent dismutation of an aldehyde substrate into alcohol and acid in the presence of LADH and NAD. Using the substrate p-trifluoromethylbenzaldehyde, we have observed various bound complexes by 19F NMR in an effort to further characterize the mechanism of the reaction. The mechanism appears to involve the catalytic activity of LADH · NAD · aldehyde complex which reacts to form an enzyme · NADH · acid complex. The affinity of the acid product for LADH · NADH is weak and the acid product readily desorbs from the ternary complex. The resulting LADH · NADH can then react with a second molecule of aldehyde to form NAD and the corresponding alcohol. The result is the conversion of two molecules of aldehyde to one each of acid and alcohol, with LADH and NAD acting catalytically. This sequence of reactions can also explain the slow formation of acid product observed when alcohol and NAD are incubated with the enzyme.  相似文献   

14.
The variation with pH of the kinetic parameters for the alcohol and acetaldehyde reactions were studied for the alleloenzyme AdhS from Drosophila melanogaster. The variation of Ki (KEO,I) with pH for two ethanol-competitive inhibitors, pyrazole and 2,2,2-trifluoroethanol, was also studied. Both alcohol oxidation and acetaldehyde reduction follow a compulsory ordered pathway, with coenzyme binding first. The rate-limiting step for ethanol oxidation is complex and involves at least hydride transfer and dissociation of the enzyme-NADH complex (ER). In contrast with this, the rate-limiting step for the back reaction, i.e. the reduction of acetaldehyde, is dissociation of the enzyme-NAD+ complex (EO). A rate-limiting ER dissociation appears in the oxidation of the secondary alcohol propan-2-ol, whereas for the back reaction, i.e. acetone reduction, hydride transfer in the ternary complexes is rate-limiting. There is one group in the free enzyme, with a pK of approx. 8.0, that regulates the kon velocity for NADH, whereas for NAD+ several groups seem to be involved. A group in the enzyme is drastically perturbed by the formation of the binary EO complex. Protonation of this group with a pK of 7.6 in the EO complex resulted in weakened alcohol and inhibitor binding, in addition to an increased dissociation rate of NAD+ from the binary EO complex. Neither the binding of acetaldehyde nor the dissociation rate of NADH from the binary ER complex varied within the pH region studied.  相似文献   

15.
The effects of quercetin and resveratrol (substances found in red wine) on the activity of cytosolic aldehyde dehydrogenase in vitro are compared with those of the synthetic hormone diethylstilbestrol. It is proposed that quercetin inhibits the enzyme by binding competitively in both the aldehyde substrate binding-pocket and the NAD(+)-binding site, whereas resveratrol and diethylstilbestrol can only bind in the aldehyde site. When inhibition is overcome by high aldehyde and NAD(+) concentrations (1 mM of each), the modifiers enhance the activity of the enzyme; we hypothesise that this occurs through binding to the enzyme-NADH complex and consequent acceleration of the rate of dissociation of NADH. The proposed ability of quercetin to bind in both enzyme sites is supported by gel filtration experiments with and without NAD(+), by studies of the esterase activity of the enzyme, and by modelling the quercetin molecule into the known three-dimensional structure of the enzyme. The possibility that interaction between aldehyde dehydrogenase and quercetin may be of physiological significance is discussed.  相似文献   

16.
Role of lecithin in D-beta-hydroxybutyrate dehydrogenase function   总被引:1,自引:0,他引:1  
Binding of NADH to D-β-hydroxybutyrate dehydrogenase, a lecithin-requiring enzyme from beef heart mitochondria, has been studied using a homogeneous enzyme which is soluble and, most important, free of lipid. The enzyme complexed with lecithin or with a phospholipid mixture containing lecithin binds NADH with a dissociation constant of 6–16 μM, while the apoenzyme or phospholipid alone or complexes formed with non-reactivating phospholipids bind no NADH. The results show that the binding of NADH to the dehydrogenase is dependent upon the formation of an enzyme-lecithin complex. This is the first demonstration of a role of lipid in a particular step of the reaction mechanism of a specific lipid-requiring enzyme.  相似文献   

17.
Glutathione reductase from Escherichia coli is inactivated when incubated with either NADPH or NADH. The process is inversely dependent on the enzyme concentration. Inactivation is rapid and monophasic with 1 microM NADPH and 1 nM enzyme FAD giving a t1/2 of 1 min. Complex formation between NADPH and the two-electron reduced enzyme (EH2) at higher levels of NADPH protects against rapid inactivation. NADP+, produced in a side reaction with oxygen, also protects by forming a complex with EH2. These complexes make analysis of the concentration dependence of the inactivation process difficult. Inactivation with NADH, where complexes do not interfere, is slower but can be analyzed more readily. With 152 microM NADH and 5.4 nM enzyme FAD, the time required for 50% inactivation is 17 min. The process is markedly biphasic, reaching the final inactivation level after 5-7 h. Analysis of the relationship between the final level of inactivation with NADH and the enzyme concentration indicates that inactivation is due to dissociation of the normally dimeric enzyme. Thus, the position of the dimer-monomer equilibrium between an active dimeric two-electron reduced species and an inactive monomeric two-electron reduced form determines the enzyme activity. An apparent equilibrium constant (Kd) for dissociation of dimer obtained from the anaerobic concentration dependent inactivation curves is 220 nM. Enzyme inactivated with NADH can be reactivated with glutathione, and the reactivation kinetics are second order, monomer-monomer over 75% of the reaction with an average apparent association rate constant (ka) of 13.1 (+/- 5.5) X 10(6) M-1 min-1.  相似文献   

18.
UDP-glucose dehydrogenase catalyzes the NAD+-dependent twofold oxidation of UDP-glucose to give UDP-glucuronic acid. A sequestered aldehyde intermediate is produced in the first oxidation step and a covalently bound thioester is produced in the second oxidation step. This work demonstrates that the Streptococcus pyogenes enzyme incorporates a single solvent-derived oxygen atom during catalysis and probably does not generate an imine intermediate. The reaction of UDP-[6",6"-di-2H]-d-glucose is not accompanied by a primary kinetic isotope effect, indicating that hydride transfer is not rate determining in this reaction. Studies with a mutant of the key active site nucleophile, Cys260Ala, show that it is capable of both reducing the aldehyde intermediate, and oxidizing the hydrated form of the aldehyde intermediate but is incapable of oxidizing UDP-glucose to UDP-glucuronic acid. In the latter case, a ternary Cys260Ala/aldehyde intermediate/NADH complex is presumably formed, but it does not proceed to product as both release and hydration of the bound aldehyde occur slowly. A washout experiment demonstrates that the NADH in this ternary complex is not exchangeable with external NADH, indicating that dissociation only occurs after the addition of a nucleophile to the aldehyde carbonyl. Studies on Thr118Ala show that the value of kcat is reduced 160-fold by this mutation, and that the reaction of UDP-D-[6",6"-di-2H]-glucose is now accompanied by a primary kinetic isotope effect. This indicates that the barriers for the hydride transfer steps have been selectively increased and supports a mechanism in which an ordered water molecule (H-bonded to Thr118) serves as the catalytic base in these steps.  相似文献   

19.
Pig heart lactate dehydrogenase was studied in the direction of pyruvate and NADH formation by recording rapid changes in extinction, proton concentration, nucleotide fluorescence and protein fluorescence. Experiments measuring extinction changes show that there is a very rapid formation of NADH within the first millisecond and that the amplitude of this phase (phase 1) increases threefold over the pH range 6-8. A second transient rate (phase 2) can also be distinguished (whose rate is pH-dependent), followed by a steady-state rate (phase 3) of NADH production. The sum of the amplitudes of the first two phases corresponds to 1mol of NADH produced/mol of active sites of lactate dehydrogenase. Experiments that measured the liberation of protons by using Phenol Red as an indicator show that no proton release occurs during the initial very rapid formation of NADH (phase 1), but protons are released during subsequent phases of NADH production. Fluorescence experiments help to characterize these phases, and show that the very rapid phase 1 corresponds to the establishment of an equilibrium between E(NAD) (Lactate) right harpoon over left harpoon H(+)E(NADH) (Pyruvate). This equilibrium can be altered by changing lactate concentration or pH, and the H(+)E(NADH) (Pyruvate) species formed has very low nucleotide fluorescence and quenched protein fluorescence. Phase 2 corresponds to the dissociation of pyruvate and a proton from the complex with a rate constant of 1150s(-1). The observed rate constant is slower than this and is proportional to the position of the preceding equilibrium. The E(NADH) formed has high nucleotide fluorescence and quenched protein fluorescence. The reaction, which is rate-limiting during steady-state turnover, must then follow this step and be involved with dissociation of NADH from the enzyme or some conformational change immediately preceding dissociation. Several inhibitory complexes have also been studied including E(NAD+) (Oxamate) and E(NADH) (Oxamate') and the abortive ternary complex E(NADH) (Lactate). The rate of NADH dissociation from the enzyme was measured and found to be the same whether measured by ligand displacement or by relaxation experiments. These results are discussed in relation to the overall mechanism of lactate dehydrogenase turnover and the independence of the four binding sites in the active tetramer.  相似文献   

20.
The transient-state kinetics of enzymic reduction of acetaldehyde and benzaldehyde by NADH, catalyzed by horse liver alcohol dehydrogenase, have been examined under single-turnover conditions, obtained by carrying out reactions either with limiting amounts of enzyme in the presence of 20 mM pyrazole or with limiting amounts of substrate. Analysis of the variation with substrate, coenzyme, and enzyme concentrations of amplitudes and time constants for the exponential transients observed at 328 nm and 300 nm shows that the kinetics of enzymic aldehyde reduction are qualitatively and quantitatively consistent with the relationships derived in the preceding paper for an ordered ternary-complex mechanism involving identical and independent catalytic sites. It is concluded that there is no evidence whatsoever for the kinetic significance of a half-of-the-sites reactivity or any other kind of subunit interaction in the liver alcohol dehydrogenase system. The biphasic transients observed at 328 nm for the reduction of aromatic aldehydes such as benzaldehyde are a normal kinetic characteristic of the ordered ternary-complex mechanism, being attributable to accumulation of the ternary enzyme-NAD-product complex when product dissociation from this complex is slow in comparison to its formation by ternary-complex interconversion.  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