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.
Nidetzky B  Klimacek M  Mayr P 《Biochemistry》2001,40(34):10371-10381
Microbial xylose reductase, a representative aldo-keto reductase of primary sugar metabolism, catalyzes the NAD(P)H-dependent reduction of D-xylose with a turnover number approximately 100 times that of human aldose reductase for the same reaction. To determine the mechanistic basis for that physiologically relevant difference and pinpoint features that are unique to the microbial enzyme among other aldo/keto reductases, we carried out stopped-flow studies with wild-type xylose reductase from the yeast Candida tenuis. Analysis of transient kinetic data for binding of NAD(+) and NADH, and reduction of D-xylose and oxidation of xylitol at pH 7.0 and 25 degrees C provided estimates of rate constants for the following mechanism: E + NADH right arrow over left arrow E.NADH right arrow over left arrow E.NADH + D-xylose right arrow over left arrow E.NADH.D-xylose right arrow over left arrow E.NAD(+).xylitol right arrow over left arrow E.NAD(+) right arrow over left arrow E.NAD(+) right arrow over left arrow E + NAD(+). The net rate constant of dissociation of NAD(+) is approximately 90% rate limiting for k(cat) of D-xylose reduction. It is controlled by the conformational change which precedes nucleotide release and whose rate constant of 40 s(-)(1) is 200 times that of completely rate-limiting E.NADP(+) --> E.NADP(+) step in aldehyde reduction catalyzed by human aldose reductase [Grimshaw, C. E., et al. (1995) Biochemistry 34, 14356-14365]. Hydride transfer from NADH occurs with a rate constant of approximately 170 s(-1). In reverse reaction, the E.NADH --> E.NADH step takes place with a rate constant of 15 s(-1), and the rate constant of ternary-complex interconversion (3.8 s(-1)) largely determines xylitol turnover (0.9 s(-1)). The bound-state equilibrium constant for C. tenuis xylose reductase is estimated to be approximately 45 (=170/3.8), thus greatly favoring aldehyde reduction. Formation of productive complexes, E.NAD(+) and E.NADH, leads to a 7- and 9-fold decrease of dissociation constants of initial binary complexes, respectively, demonstrating that 12-fold differential binding of NADH (K(i) = 16 microM) vs NAD(+) (K(i) = 195 microM) chiefly reflects difference in stabilities of E.NADH and E.NAD(+). Primary deuterium isotope effects on k(cat) and k(cat)/K(xylose) were, respectively, 1.55 +/- 0.09 and 2.09 +/- 0.31 in H(2)O, and 1.26 +/- 0.06 and 1.58 +/- 0.17 in D(2)O. No deuterium solvent isotope effect on k(cat)/K(xylose) was observed. When deuteration of coenzyme selectively slowed the hydride transfer step, (D)()2(O)(k(cat)/K(xylose)) was inverse (0.89 +/- 0.14). The isotope effect data suggest a chemical mechanism of carbonyl reduction by xylose reductase in which transfer of hydride ion is a partially rate-limiting step and precedes the proton-transfer step.  相似文献   

2.
Digits JA  Hedstrom L 《Biochemistry》1999,38(8):2295-2306
IMP dehydrogenase (IMPDH) catalyzes the oxidation of IMP to XMP with conversion of NAD+ to NADH. This reaction is the rate-limiting step in de novo guanine nucleotide biosynthesis. IMPDH is a target for antitumor, antiviral, and immunosuppressive chemotherapy. We have determined the complete kinetic mechanism for IMPDH from Tritrichomonas foetus using ligand binding, isotope effect, pre-steady-state kinetic, and rapid quench kinetic experiments. Both substrates bind to the free enzyme, which suggests a random mechanism. IMP binds to the enzyme in two steps. Two steps are also involved when IMP binds to a mutant IMPDH in which the active site Cys is substituted with a Ser. This observation suggests that this second step may be a conformational change of the enzyme. No Vm isotope effect is observed when [2-2H]IMP is the substrate which indicates that hydride transfer is not rate-limiting. This result is confirmed by the observation of a pre-steady-state burst of NADH production when monitored by absorbance. However, when NADH production was monitored by fluorescence, the rate constant for the exponential phase is 5-10-fold lower than when measured by absorbance. This observation suggests that the fluorescence of enzyme-bound NADH is quenched and that this transient represents NADH release from the enzyme. The time-dependent formation and decay of [14C]E-XMP intermediates was monitored using rapid quench kinetics. These experiments indicate that both NADH release and E-XMP hydrolysis are rate-limiting and suggest that NADH release precedes hydrolysis of E-XMP.  相似文献   

3.
M Slatner  B Nidetzky  K D Kulbe 《Biochemistry》1999,38(32):10489-10498
To characterize catalysis by NAD-dependent long-chain mannitol 2-dehydrogenases (MDHs), the recombinant wild-type MDH from Pseudomonas fluorescens was overexpressed in Escherichia coli and purified. The enzyme is a functional monomer of 54 kDa, which does not contain Zn(2+) and has B-type stereospecificity with respect to hydride transfer from NADH. Analysis of initial velocity patterns together with product and substrate inhibition patterns and comparison of primary deuterium isotope effects on the apparent kinetic parameters, (D)k(cat), (D)(k(cat)/K(NADH)), and (D)(k(cat)/K(fructose)), show that MDH has an ordered kinetic mechanism at pH 8.2 in which NADH adds before D-fructose, and D-mannitol and NAD are released in that order. Isomerization of E-NAD to a form which interacts with D-mannitol nonproductively or dissociation of NAD from the binary complex after isomerization is the slowest step (>/=110 s(-)(1)) in D-fructose reduction at pH 8.2. Release of NADH from E-NADH (32 s(-)(1)) is the major rate-limiting step in mannitol oxidation at this pH. At the pH optimum for D-fructose reduction (pH 7.0), the rate of hydride transfer contributes significantly to rate limitation of the catalytic cascade and the overall reaction. (D)(k(cat)/K(fructose)) decreases from 2.57 at pH 7.0 to a value of 相似文献   

4.
Heredia VV  Penning TM 《Biochemistry》2004,43(38):12028-12037
3Alpha-hydroxysteroid dehydrogenases (3alpha-HSDs) catalyze the interconversion between 5alpha-dihydrotestosterone (5alpha-DHT), the most potent androgen, and 3alpha-androstanediol (3alpha-diol), a weak androgen metabolite. To identify the rate-determining step in this physiologically important reaction, rat liver 3alpha-HSD (AKR1C9) was used as the protein model for the human homologues in fluorescence stopped-flow transient kinetic and kinetic isotope effect studies. Using single and multiple turnover experiments to monitor the NADPH-dependent reduction of 5alpha-DHT, it was found that k(lim) and k(max) values were identical to k(cat), indicating that chemistry is rate-limiting overall. Kinetic isotope effect measurements, which gave (D)k(cat) = 2.4 and (D)2(O)k(cat) = 3.0 at pL 6.0, suggest that the slow chemical transformation is significantly rate-limiting. When the NADP(+)-dependent oxidation of 3alpha-diol was monitored, single and multiple turnover experiments showed a k(lim) and burst kinetics consistent with product release as being rate-limiting overall. When NAD(+) was substituted for NADP(+), burst phase kinetics was eliminated, and k(max) was identical to k(cat). Thus with the physiologically relevant substrates 5alpha-DHT plus NADPH and 3alpha-diol plus NAD(+), the slowest event is chemistry. R276 forms a salt-linkage with the phosphate of 2'-AMP, and when it is mutated, tight binding of NAD(P)H is no longer observed [Ratnam, K., et al. (1999) Biochemistry 38, 7856-7864]. The R276M mutant also eliminated the burst phase kinetics observed for the NADP(+)-dependent oxidation of 3alpha-diol. The data with the R276M mutant confirms that the release of the NADPH product is the slow event; and in its absence, chemistry becomes rate-limiting. W227 is a critical hydrophobic residue at the steroid binding site, and when it is mutated to alanine, k(cat)/K(m) for oxidation is significantly depressed. Burst phase kinetics for the NADP(+)-dependent turnover of 3alpha-diol by W227A was also abolished. In the W227A mutant, the slow release of NADPH is no longer observed since the chemical transformation is now even slower. Thus, residues in the cofactor and steroid-binding site can alter the rate-determining step in the NADP(+)-dependent oxidation of 3alpha-diol to make chemistry rate-limiting overall.  相似文献   

5.
Site-directed mutagenesis of Thr(66) in porcine liver NADH-cytochrome b(5) reductase demonstrated that this residue modulates the semiquinone form of FAD and the rate-limiting step in the catalytic sequence of electron transfer. The absorption spectrum of the T66V mutant showed a typical neutral blue semiquinone intermediate during turnover in the electron transfer from NADH to ferricyanide but showed an anionic red semiquinone form during anaerobic photoreduction. The apparent k(cat) values of this mutant were approximately 10% of that of the wild type enzyme (WT). These data suggest that the T66V mutation stabilizes the neutral blue semiquinone and that the conversion of the neutral blue to the anionic red semiquinone form is the rate-limiting step. In the WT, the value of the rate constant of FAD reduction (k(red)) was consistent with the k(cat) values, and the oxidized enzyme-NADH complex was observed during the turnover with ferricyanide. This indicates that the reduction of FAD by NADH in the WT-NADH complex is the rate-limiting step. In the T66A mutant, the k(red) value was larger than the k(cat) values, but the k(red) value in the presence of NAD(+) was consistent with the k(cat) values. The spectral shape of this mutant observed during turnover was similar to that during the reduction with NADH in the presence of NAD(+). These data suggest that the oxidized T66A-NADH-NAD(+) ternary complex is a major intermediate in the turnover and that the release of NAD(+) from this complex is the rate-limiting step. These results substantiate the important role of Thr(66) in the one-electron transfer reaction catalyzed by this enzyme. On the basis of these data, we present a new kinetic scheme to explain the mechanism of electron transfer from NADH to one-electron acceptors including cytochrome b(5).  相似文献   

6.
Dihydrofolate reductase from Mycobacterium tuberculosis (MtDHFR) catalyzes the NAD(P)-dependent reduction of dihydrofolate, yielding NAD(P)(+) and tetrahydrofolate, the primary one-carbon unit carrier in biology. Tetrahydrofolate needs to be recycled so that reactions involved in dTMP synthesis and purine metabolism are maintained. In this work, we report the kinetic characterization of the MtDHFR. This enzyme has a sequential steady-state random kinetic mechanism, probably with a preferred pathway with NADPH binding first. A pK(a) value for an enzymic acid of approximately 7.0 was identified from the pH dependence of V, and the analysis of the primary kinetic isotope effects revealed that the hydride transfer step is at least partly rate-limiting throughout the pH range analyzed. Additionally, solvent and multiple kinetic isotope effects were determined and analyzed, and equilibrium isotope effects were measured on the equilibrium constant. (D(2)O)V and (D(2)O)V/K([4R-4-(2)H]NADH) were slightly inverse at pH 6.0, and inverse values for (D(2)O)V([4R-4-(2)H]NADH) and (D(2)O)V/K([4R-4-(2)H]NADH) suggested that a pre-equilibrium protonation is occurring before the hydride transfer step, indicating a stepwise mechanism for proton and hydride transfer. The same value was obtained for (D)k(H) at pH 5.5 and 7.5, reaffirming the rate-limiting nature of the hydride transfer step. A chemical mechanism is proposed on the basis of the results obtained here.  相似文献   

7.
Aldehyde dehydrogenases are general detoxifying enzymes, but there are also isoenzymes that are involved in specific metabolic pathways in different organisms. Two of these enzymes are Escherichia coli lactaldehyde (ALD) and phenylacetaldehyde dehydrogenases (PAD), which participate in the metabolism of fucose and phenylalanine, respectively. These isozymes share some properties with the better characterized mammalian enzymes but have kinetic properties that are unique. It was possible to thread the sequences into the known ones for the mammalian isozymes to better understand some structural differences. Both isozymes were homotetramers, but PAD used both NAD+ and NADP+ but with a clear preference for NAD, while ALD used only NAD+. The rate-limiting step for PAD was hydride transfer as indicated by the primary isotopic effect and the absence of a pre-steady-state burst, something not previously found for tetrameric enzymes from other organisms where the rate-limiting step is related to both deacylation and coenzyme dissociation. In contrast, ALD had a pre-steady-state burst indicating that the rate-limiting step was located after the NADH formation, but the rate-limiting step was a combination of deacylation and coenzyme dissociation. Both enzymes possessed esterase activity that was stimulated by NADH; NAD+ stimulated the esterase activity of PAD but not of ALD. Finding enzymes that structurally are similar to the well-characterized mammalian enzymes but have a different rate-limiting step might serve as models to allow us to determine what regulates the rate-limiting step.  相似文献   

8.
The steady-state kinetic parameters for epimerization of UDP-galactose by UDP-galactose 4-epimerase from Escherichia coli (GalE), Y149F-GalE, and S124A-GalE have been measured as a function of pH. The deuterium kinetic isotope effects for epimerization of UDP-galactose-C-d(7) by these enzymes have also been measured. The results show that the activity of wild-type GalE is pH-independent in the pH range of 5.5-9.3, and there is no significant deuterium kinetic isotope effect in the reaction of UDP-galactose-C-d(7). It is concluded that the rate-limiting step for epimerization by wild-type GalE is not hydride transfer and must be either a diffusional process or a conformational change. Epimerization of UDP-galactose-C-d(7) by Y149F-GalE proceeds with a pH-dependent deuterium kinetic isotope effect on k(cat) of 2.2 +/- 0.4 at pH 6.2 and 1.1 +/- 0.5 at pH 8.3. Moreover, the plot of log k(cat)/K(m) breaks downward on the acid side with a fitted value of 7.1 for the pK(a). It is concluded that the break in the pH-rate profile arises from a change in the rate-limiting step from hydride transfer at low pH to a conformational change at high pH. Epimerization of UDP-galactose-C-d(7) by S124A-GalE proceeds with a pH-independent deuterium kinetic isotope effect on k(cat) of 2.0 +/- 0.2 between pH 6 and 9. Both plots of log k(cat) and log k(cat)/K(m) display pH dependence. The plot of log k(cat) versus pH breaks downward with a pK(a) of 6.35 +/- 0.10. The plot of log k(cat)/K(m) versus pH is bell-shaped, with fitted pK(a) values of 6.76 +/- 0.09 and 9.32 +/- 0.21. It is concluded that hydride transfer is rate-limiting, and the pK(a) of 6.7 for free S124A-GalE is assigned to Tyr 149, which displays the same value of pK(a) when measured spectrophotometrically in this variant. Acid-base catalysis by Y149F-GalE is attributed to Ser 124, which is postulated to rescue catalysis of proton transfer in the absence of Tyr 149. The kinetic pK(a) of 7.1 for free Y149F-GalE is lower than that expected for Ser 124, as proven by the pH-dependent kinetic isotope effect. Epimerization by the doubly mutated Y149F/S124A-GalE proceeds at a k(cat) that is lower by a factor of 10(7) than that of wild-type GalE. This low rate is attributed to the synergistic actions of Tyr 149 and Ser 124 in wild-type GalE and to the absence of any internal catalysis of hydride transfer in the doubly mutated enzyme.  相似文献   

9.
Steady state kinetic analysis at pH 7.0 of the reduction of DL-glyceraldehyde by pig muscle aldose reductase showed that the enzyme follows a sequential ordered mechanism with NADPH binding first. However, the "off constant" for NADP+ in the forward direction was 1 order of magnitude less than the kcat. Analysis of this anomaly by pre-steady state kinetics using stopped-flow fluorescence spectroscopy showed that this could be accounted for by isomerization of the enzyme-NADP+ complex and that the rate of isomerization is the rate-limiting step. The rate constant for this step was of the same order of magnitude as the kcat for the forward reaction. Fluorescence emission spectra of free and NADP(H)-bound enzyme suggested a conformational change upon binding of coenzyme. In the reverse direction (oxidation of glycerol) pre-steady state and steady state kinetic analyses were consistent with the rate-limiting step occurring before isomerization of the enzyme-NADPH complex. We conclude, therefore, that during the kinetic mechanism of the reduction of aldehydes by aldose reductase, a slow (kinetically detectable) conformational change in the enzyme occurs upon coenzyme binding. Since NADPH and NADP+ bind to the enzyme very tightly, this has implications for the targeting and binding of drugs that are aldose reductase inhibitors.  相似文献   

10.
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   

11.
Knapp MJ  Klinman JP 《Biochemistry》2003,42(39):11466-11475
The reactivity of O(2) with soybean lipoxygenase-1 (SLO) has been examined using a range of kinetic probes. We are able to rule out diffusional encounter of O(2) with protein, an outer-sphere electron transfer to O(2), and proton transfer as rate-limiting steps in k(cat)/K(M)(O(2)) for wild-type enzyme (WT SLO); this restricts the rate-limiting step to either the combination of O(2) with L(*) or a subsequent conformational change. In the Ile(553) --> Phe mutant, which constricts the putative O(2) binding channel [Knapp et al. (2001) J. Am. Chem. Soc. 123, 2931-2932], k(cat)/K(M)(O(2)) decreases by over a factor of 20; yet, this mutant appears to have the same rate-limiting step as WT SLO. It is argued that the slow step on k(cat)/K(M)(O(2)) is the combination of O(2) with L(*), with proximal protein effects determining the rate of reaction. The available data for SLO support the view that enzymes can affect O(2) reactivity without a direct involvement of metal cofactors. The primary role of the Fe(3+) cofactor is to generate an enzyme-bound radical, while the protein is concluded to control the stereo- and regiochemistry of O(2) encounter with this radical.  相似文献   

12.
Rapid mixing, kinetic experiments were performed on native and modified [Tyr(3NO2)237)] porcine H4 lactate dehydrogenase at low temperatures in a medium containing 30% dimethyl sulphoxide. In the temperature range -16 to +8 degrees C, the modified enzyme-NADH complex, when mixed with 1 mM pyruvate, is converted to enzyme, NAD+ and lactate at two distinctly different rates. At -16 degrees C the more rapid process occurs at a rate of 40 s-1 and the slower at 3 s-1. The slower rate is identical to that assigned to the steady-state turnover of the enzyme in these conditions and therefore reflects the slow, rate-limiting rearrangement of protein structure which has been inferred from previous kinetic experiments. The fast phase of NADH oxidation, however, proceeds at a rate which coincides with that of the closure of a loop of polypeptide over the active site of the enzyme (sensed by the nitrotyrosine group, which protonates in response to the approach of glutamate 107, a residue situated on this mobile loop). We explain these results by proposing that: (i) both the slow and fast changes in protein structure must occur before the enzyme can accomplish the redox step, (ii) the enzyme-NADH (binary) complex exists in two, slowly interconverting forms, (iii) the structural change giving rise to this slow conformational equilibrium can also occur in the ternary (enzyme-NADH-pyruvate) complex and (iv) it is this step which limits the rate of the steady-state reaction. Both of the binary forms are able to bind pyruvate, but the rate of NADH oxidation in one of the forms is rapid, since it has already undergone this slow rearrangement. In this rapidly reacting form, it is the closure of the loop (not transfer of the hydride ion) which limits the rate at which the coenzyme is oxidized, while the slowly reacting form must undergo both loop-closure and the slow structural conversion before the redox reaction can occur.  相似文献   

13.
Stopped-flow experiments in which sheep liver cytoplasmic aldehyde dehydrogenase (EC 1.2.1.3) was rapidly mixed with NAD(+) and aldehyde showed a burst of NADH formation, followed by a slower steady-state turnover. The kinetic data obtained when the relative concentrations and orders of mixing of NAD(+) and propionaldehyde with the enzyme were varied were fitted to the following mechanism: [Formula: see text] where the release of NADH is slow. By monitoring the quenching of protein fluorescence on the binding of NAD(+), estimates of 2x10(5) litre.mol(-1).s(-1) and 2s(-1) were obtained for k(+1) and k(-1) respectively. Although k(+3) could be determined from the dependence of the burst rate constant on the concentration of propionaldehyde to be 11s(-1), k(+2) and k(-2) could not be determined uniquely, but could be related by the equation: (k(-2)+k(+3))/k(+2) =50x10(-6)mol.litre(-1). No significant isotope effect was observed when [1-(2)H]propionaldehyde was used as substrate. The burst rate constant was pH-dependent, with the greatest rate constants occurring at high pH. Similar data were obtained by using acetaldehyde, where for this substrate (k(-2)+k(+3))/k(+2)=2.3x10 (-3)mol.litre(-1) and k(+3) is 23s(-1). When [1,2,2,2-(2)H]acetaldehyde was used, no isotope effect was observed on k(+3), but there was a significant effect on k(+2) and k(-2). A burst of NADH production has also been observed with furfuraldehyde, trans-4-(NN-dimethylamino)cinnamaldehyde, formaldehyde, benzaldehyde, 4-(imidazol-2-ylazo)benzaldehyde, p-methoxybenzaldehyde and p-methylbenzaldehyde as substrates, but not with p-nitrobenzaldehyde.  相似文献   

14.
Ma H  Ratnam K  Penning TM 《Biochemistry》2000,39(1):102-109
Rat liver 3alpha-hydroxysteroid dehydrogenase (3alpha-HSD), an aldo-keto reductase, binds NADP(+) in an extended anti-conformation across an (alpha/beta)(8)-barrel. The orientation of the nicotinamide ring, which permits stereospecific transfer of the 4-pro-R hydride from NAD(P)H to substrate, is achieved by hydrogen bonds formed between the C3-carboxamide of the nicotinamide ring and Ser 166, Asn 167, and Gln 190 and by pi-stacking between this ring and Tyr 216. These residues were mutated to yield S166A, N167A, Q190A, and Y216S. In these mutants, K(d)(NADP(H)) increased by 2-11-fold but without a significant change in K(d)(NAD(H)). Steady-state kinetic parameters showed that K(m)(NADP)()+ increased 13-151-fold, and this was accompanied by comparable decreases in k(cat)/K(m)(NADP)()+. By contrast, K(m)(NAD)()+ increased 4-8-fold, but changes in k(cat)/K(m)(NAD)()+ were more dramatic and ranged from 23- to 930-fold. Corresponding changes in binding energies indicated that each residue contributed equally to the binding of NADP(H) in the ground and transition states. However, the same residues stabilized the binding of NAD(H) only in the transition state. These observations suggest that different modes of binding exist for NADP(H) and NAD(H). Importantly, these modes were revealed by mutating residues in the nicotinamide pocket indicating that direct interactions with the 2'-phosphate in the adenine mononucleotide is not the sole determinant of cofactor preference. The single mutations were unable to invert or racemize the stereochemistry of hydride transfer even though the nicotinamide pocket can accommodate both anti- and syn-conformers once the necessary hydrogen bonds are eliminated. When 4-pro-R-[(3)H]NADH was used to monitor incorporation into [(14)C]-5alpha-dihydrotestosterone, a decrease in the (3)H:(14)C ratio was observed in the mutants relative to wild-type enzyme reflecting a pronounced primary kinetic isotope effect. This observation coupled with the change in the binding energy for NAD(P)(H) in the transition state suggests that these mutants have altered the reaction trajectory for hydride transfer.  相似文献   

15.
The kinetics of the enzymatic step of the peroxidatic reaction between NAD and hydrogen peroxide, catalysed by horse liver alcohol dehydrogenase (alcohol:NAD+ oxidoreductase, EC 1.1.1.1), has been investigated at pH 7 at high enzyme concentration. Under such conditions no burst phase has been observed, thus indicating that the rate-limiting step in the process, which converts NAD into Compound I, either precedes or coincides with the chemical step responsible for the observed spectroscopic change. Kinetic analysis of the data, performed according to a simplified reaction scheme suggests that the rate-limiting step is coincident with the spectroscopic (i.e., chemical) step itself. Furthermore, the absence of a proton burst phase indicates the proton release step does not precede the chemical step, in contrast with the case of ethanol oxidation. A kinetic effect of different premixing conditions on the reaction rate has been observed and attributed to the presence of NADH formed in the 'blank reaction' between NAD and residual ethanol tightly bound to alcohol dehydrogenase. A molecular mechanism for the enzymatic peroxidation step is finally proposed, exploiting the knowledge of the much better known reaction of ethanol oxidation. Inhibition of this reaction by NADH has been investigated with respect to H2O2 (noncompetitive, Ki about 10 microM) and to NAD (competitive, Ki about 0.7 microM). The effect of temperature on the steady-state reaction state (about 65 kJ/mol activation energy) has also been studied.  相似文献   

16.
The A245K mutant of Bacillus stearothermophilus L-lactate dehydrogenase has been expressed in Escherichia coli and purified. A qualitative change in the reaction mechanism prior to the hydride transfer step in the reverse direction in the mutant is revealed. Both transient and steady state characteristics of the mutant are presented and show in contrast to the wild-type enzyme where a rearrangement of an enzyme-NADH-pyruvate complex is rate-limiting that in the mutant the rearrangement is much faster and hydride transfer is the first slow step. The steady state is limited by a new second slower conformation change involving an NAD+ complex. The mutation may provide a valuable framework for inhibitor and drug design research.  相似文献   

17.
Pan JJ  Chiou ST  Liang PH 《Biochemistry》2000,39(35):10936-10942
Undecaprenyl pyrophosphate synthase (UPPs) catalyzes the condensation of eight molecules of isopentenyl pyrophosphate (IPP) with farnesyl pyrophosphate (FPP) to generate C(55) undecaprenyl pyrophosphate. We investigated the kinetics and mechanism of this reaction pathway using Escherichia coli UPPs. With a variety of different ratios of enzyme to substrate and FPP to IPP in the presence or absence of Triton, different product distributions were found. In the presence of excess FPP, the intermediates (C(25)-C(50)) accumulated. Under a condition with enzyme and FPP in excess of IPP, instead of C(20)-geranylgeranyl pyrophosphate, C(20), C(25), and C(30) were the major products. The UPPs steady-state k(cat) value (2.5 s(-1)) in the presence of 0.1% Triton was 190-fold larger than in the absence of Triton (0.013 s(-1)). The k(cat) value matched the rate constant of each IPP condensation obtained from the enzyme single-turnover experiments. This suggested that the IPP condensation rather than product release was the rate-limiting step in the presence of Triton. In the absence of Triton, the intermediates formed and disappeared in a similar manner under enzyme single turnover in contrast to the slow steady-state rate, which indicated a step after product generation was rate limiting. This was further supported by a burst product formation. Judging from the accumulation level of C(55), C(60), and C(65), their dissociation from the enzyme cannot be too slow and an even slower enzyme conformational change with a rate of 0.001 s(-1) might govern the UPPs reaction rate under the steady-state condition in the absence of Triton.  相似文献   

18.
J Ovádi  I R Osman  J Batke 《Biochemistry》1982,21(25):6375-6382
Covalent binding of FITC up to 2 mol/mol of tetrameric enzyme does not affect the enzymatic activity and dissociation properties of pig muscle D-glyceraldehyde-3-phosphate dehydrogenase (GAPD). The binding of NAD to dehydrogenase-FITC complex partially reverts the quenching caused by the binding of dye to apo-GAPD. This phenomenon, as well as the formation of a characteristic absorption difference spectrum caused by the binding of NAD, makes it possible to follow the NAD-induced local conformational changes near the dye-binding region. The time course of NAD-induced spectral changes shows biphasic kinetics: a burst and a slow phase. The amplitude of burst phase as a function of NAD equivalents has sigmoidal shape due to the cooperative interaction between subunits. The same conclusion could be drawn from fluorescence anisotropy measurements. In the presence of excess NAD a slow conformational change can be detected, the amplitude of which is a function of NAD concentration. This phenomenon can be attributed to the binding of further NAD molecules to the holoenzyme. The slow phase follows first-order kinetics, and the rate constant depends on enzyme concentration. The specific fluorescence intensity and the fluorescence anisotropy of fluorescent dye labeled apo-GAPD and GAPD saturated with NAD are also dependent on enzyme concentration. We suggest that NAD binding induces major changes in the steric structure of tetrameric enzyme without influencing remarkably the interacting forces between the contact surfaces of subunits. Data are quantitatively interpreted in terms of a two-step dissociation model.  相似文献   

19.
Although the structures of mammalian cytosolic and mitochondrial ALDH have been determined, several differences, mainly functional, between these two 70% identical isozymes remain unexplained. A major difference is the differential effect of Mg(2+) ions that inhibits the cytosolic and activates the mitochondrial isozyme. Here, we have investigated the effect of Mg(2+) ions on each individual kinetic step of ALDH1 and ALDH2. The metal ions were found not to affect either acylation or hydride transfer for either isozyme. The lack of a Mg(2+) ion effect on hydride transfer was further demonstrated with an E399Q mutant of ALDH1 whose rate-limiting step had been changed from NADH dissociation to hydride transfer. The other steps, however, were affected by Mg(2+) ions for both isozymes. The metal ions inhibited NADH dissociation, the rate-limiting step for ALDH1, and enhanced deacylation, the rate-limiting step for ALDH2. Our results indicated that, with both isozymes, Mg(2+) ions tightened the binding of NADH, and by binding to the coenzyme, they increased the nucleophilicity of the nucleophile Cys302. The inhibition of ALDH1 and activation of ALDH2 at pH 7.4 are due to their different rate-limiting steps. Mg(2+) ions affected similarly the NADH activation of the esterase reaction for both isozymes. In contrast, the metal ions affected only the NAD(+) activation of ALDH1. This latter finding and other features described here can be rationalized on the basis of the known three-dimensional structures of the isozymes.  相似文献   

20.
Klimacek M  Nidetzky B 《Biochemistry》2002,41(31):10158-10165
Mannitol dehydrogenases (MDH) are a family of Zn(2+)-independent long-chain alcohol dehydrogenases that catalyze the regiospecific NAD(+)-dependent oxidation of a secondary alcohol group in polyol substrates. pH and primary deuterium kinetic isotope effects on kinetic parameters for reaction of recombinant MDH from Pseudomonas fluorescens with D-mannitol have been measured in H(2)O and D(2)O at 25 degrees C and used to determine the relative timing of C-H and O-H bond cleavage steps during alcohol conversion. The enzymatic rates decreased at low pH; apparent pK values for log(k(cat)/K(mannitol)) and log k(cat) were 9.2 and 7.7 in H(2)O, respectively, and both were shifted by +0.4 pH units in D(2)O. Proton inventory plots for k(cat) and k(cat)/K(mannitol) were determined at pL 10.0 using protio or deuterio alcohol and were linear at the 95% confidence level. They revealed the independence of primary deuterium isotope effects on the atom fraction of deuterium in a mixed H(2)O-D(2)O solvent and yielded single-site transition-state fractionation factors of 0.43 +/- 0.05 and 0.47 +/- 0.01 for k(cat)/K(mannitol) and k(cat), respectively. (D)(k(cat)/K(mannitol)) was constant (1.80 +/- 0.20) in the pH range 6.0-9.5 and decreased at high pH to a limiting value of approximately 1. Measurement of (D)(k(cat)/K(fructose)) at pH 10.0 and 10.5 using NADH deuterium-labeled in the 4-pro-S position gave a value of 0.83, the equilibrium isotope effect on carbonyl group reduction. A mechanism of D-mannitol oxidation by MDH is supported by the data in which the partly rate-limiting transition state of hydride transfer is stabilized by a single solvation catalytic proton bridge. The chemical reaction involves a pH-dependent internal equilibrium which takes place prior to C-H bond cleavage and in which proton transfer from the reactive OH to the enzyme catalytic base may occur. Loss of a proton from the enzyme at high pH irreversibly locks the ternary complex with either alcohol or alkoxide bound in a conformation committed of undergoing NAD(+) reduction at a rate about 2.3-fold slower than the corresponding reaction rate of the protonated complex. Transient kinetic studies for D-mannitol oxidation at pH(D) 10.0 showed that the solvent isotope effect on steady-state turnover originates from a net rate constant of NADH release that is approximately 85% rate-limiting for k(cat) and 2-fold smaller in D(2)O than in H(2)O.  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