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.
Phosphite dehydrogenase (PTDH) catalyzes the NAD-dependent oxidation of phosphite to phosphate, a reaction that is 15 kcal/mol exergonic. The enzyme belongs to the family of D-hydroxy acid dehydrogenases. Five other family members that were analyzed do not catalyze the oxidation of phosphite, ruling out the possibility that this is a ubiquitous activity of these proteins. PTDH does not accept any alternative substrates such as thiophosphite, hydrated aldehydes, and methylphosphinate, and potential small nucleophiles such as hydroxylamine, fluoride, methanol, and trifluoromethanol do not compete with water in the displacement of the hydride from phosphite. The pH dependence of k(cat)/K(m,phosphite) is bell-shaped with a pK(a) of 6.8 for the acidic limb and a pK(a) of 7.8 for the basic limb. The pK(a) of 6.8 is assigned to the second deprotonation of phosphite. However, whether the dianionic form of phosphite is the true substrate is not clear since a reverse protonation mechanism is also consistent with the available data. Unlike k(cat)/K(m,phosphite), k(cat) and k(cat)/K(m,NAD) are pH-independent. Sulfite is a strong inhibitor of PTDH that is competitive with respect to phosphite and uncompetitive with respect to NAD(+). Incubation of the enzyme with NAD(+) and low concentrations of sulfite results in a covalent adduct between NAD(+) and sulfite in the active site of the enzyme that binds very tightly. Fluorescent titration studies provided the apparent dissociation constants for NAD(+), NADH, sulfite, and the sulfite-NAD(+) adduct. Substrate isotope effect studies with deuterium-labeled phosphite resulted in small normal isotope effects (1.4-2.1) on both k(cat) and k(cat)/K(m,phosphite) at pH 7.25 and 8.0. Solvent isotope effects (SIEs) on k(cat) are similar in size; however, the SIE of k(cat)/K(m,phosphite) at pH 7.25 is significantly larger (4.4), whereas at pH 8.0, it is the inverse (0.6). The pH-rate profile of k(cat)/K(m,phosphite), which predicts that the observed SIEs will have a significant thermodynamic origin, can account for these effects.  相似文献   

2.
The pH dependence of the kinetic parameters and the primary deuterium isotope effects with nicotinamide adenine dinucleotide (NAD) and also thionicotinamide adenine dinucleotide (thio-NAD) as the nucleotide substrates were determined in order to obtain information about the chemical mechanism and location of rate-determining steps for the Ascaris suum NAD-malic enzyme reaction. The maximum velocity with thio-NAD as the nucleotide is pH-independent from pH 4.2 to 9.6, while with NAD, V decreases below a pK of 4.8. V/K for both nucleotides decreases below a pK of 5.6 and above a pK of 8.9. Both the tartronate pKi and V/Kmalate decrease below a pK of 4.8 and above a pK of 8.9. Oxalate is competitive vs. malate above pH 7 and noncompetitive below pH 7 with NAD as the nucleotide. The oxalate Kis increases from a constant value above a pK of 4.9 to another constant value above a pK of 6.7. The oxalate Kii also increases above a pK of 4.9, and this inhibition is enhanced by NADH. In the presence of thio-NAD the inhibition by oxalate is competitive vs. malate below pH 7. For thio-NAD, both DV and D(V/K) are pH-independent and equal to 1.7. With NAD as the nucleotide, DV decreases to 1.0 below a pK of 4.9, while D(V/KNAD) and D(V/Kmalate) are pH-independent. Above pH 7 the isotope effects on V and the V/K values for NAD and malate are equal to 1.45, the pH-independent value of DV above pH 7. From the above data, the following conclusions can be made concerning the mechanism for this enzyme. Substrates bind to only the correctly protonated form of the enzyme. Two enzyme groups are necessary for binding of substrates and catalysis. Both NAD and malate are released from the Michaelis complex at equal rates which are equal to the rate of NADH release from E-NADH above pH 7. Below pH 7 NADH release becomes more rate-determining as the pH decreases until at pH 4.0 it completely limits the overall rate of the reaction.  相似文献   

3.
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   

4.
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   

5.
Human malic enzyme was studied by steady state kinetics, deuterium isotope effects, and 13C isotope effects with both the physiological dinucleotide cofactor and several alternate cofactors. The log V vs pH profile with NAD revealed two pK(a) values too close to be separately determined, but with an average value of 7.33. The log V/K vs pH profile with NAD revealed two pK(a) values at 7.4 and 5.6. Deuterium and 13C isotope effects indicate that the mechanism of human malic enzyme is stepwise with both NAD and epsilonNAD, but that hyperconjugation in the transition state for hydride transfer is detectable only with the former. With thioNAD and APAD, the isotope effects do not clearly indicate whether the mechanism is stepwise or concerted. The intrinsic 13C isotope effect for decarboxylation was calculated to be 1.0485 by measurement of the partition ratio of oxaloacetate in the presence of NADH and human malic enzyme (decarboxylation to pyruvate/reduction to malate = 2.33). The isotope effect and partitioning data suggest that the energy barrier for decarboxylation of oxaloacetate is not as high relative to the barrier for reduction of oxaloacetate as with the chicken liver enzyme.  相似文献   

6.
Argyrou A  Blanchard JS 《Biochemistry》2004,43(14):4375-4384
1-Deoxy-D-xylulose-5-phosphate (DXP) isomeroreductase catalyzes the isomerization and reduced nicotinamide adenine dinucleotide phosphate- (NADPH-) dependent reduction of DXP to generate 2-C-methylerythritol 4-phosphate (MEP) in the first committed step of the MEP pathway of isoprenoid biosynthesis. We have cloned the gene encoding the Mycobacterium tuberculosis DXP isomeroreductase, expressed the protein in Escherichia coli, and purified the enzyme to homogeneity using conventional column chromatography methods. DXP isomeroreductase is a metal ion-activated enzyme displaying superior specificity for Co(2+), good specificity for Mn(2+), and poor specificity for Mg(2+). Although NADPH is preferred over reduced nicotinamide adenine dinucleotide (NADH) about 100-fold as evaluated by the relative k(cat)/K(m) values, the maximum turnover numbers are similar, suggesting that the 2'-phosphate of NADPH contributes predominantly to binding and not to catalysis. While k(cat) was independent of pH in the region 6.0 相似文献   

7.
The solvent kinetic isotope effects (SKIE) on the yeast alpha-glucosidase-catalyzed hydrolysis of p-nitrophenyl and methyl-d-glucopyranoside were measured at 25 degrees C. With p-nitrophenyl-D-glucopyranoside (pNPG), the dependence of k(cat)/K(m) on pH (pD) revealed an unusually large (for glycohydrolases) solvent isotope effect on the pL-independent second-order rate constant, (DOD)(k(cat)/K(m)), of 1.9 (+/-0.3). The two pK(a)s characterizing the pH profile were increased in D(2)O. The shift in pK(a2) of 0.6 units is typical of acids of comparable acidity (pK(a)=6.5), but the increase in pK(a1) (=5.7) of 0.1 unit in going from H(2)O to D(2)O is unusually small. The initial velocities show substrate inhibition (K(is)/K(m) approximately 200) with a small solvent isotope effect on the inhibition constant [(DOD)K(is)=1.1 (+/-0.2)]. The solvent equilibrium isotope effects on the K(is) for the competitive inhibitors D-glucose and alpha-methyl D-glucoside are somewhat higher [(DOD)K(i)=1.5 (+/-0.1)]. Methyl glucoside is much less reactive than pNPG, with k(cat) 230 times lower and k(cat)/K(m) 5 x 10(4) times lower. The solvent isotope effect on k(cat) for this substrate [=1.11 (+/-0. 02)] is lower than that for pNPG [=1.67 (+/-0.07)], consistent with more extensive proton transfer in the transition state for the deglucosylation step than for the glucosylation step.  相似文献   

8.
The prephenate dehydrogenase activity of the bifunctional enzyme chorismate mutase-prephenate dehydrogenase from Escherichia coli catalyzes the oxidative decarboxylation of both prephenate and deoxoprephenate, which lacks the keto group in the side chain (V 78% and V/K 18% those of prephenate). Hydride transfer is to the B side of NAD, and the acetylpyridine and pyridinecarboxaldehyde analogues of NAD have V/K values 40 and 9% and V values 107 and 13% those of NAD. Since the 13C isotope effect on the decarboxylation is 1.0103 with deuterated and 1.0033 with unlabeled deoxoprephenate (the deuterium isotope effect on V/K is 2.34), the mechanism is concerted, and if CO2 has no reverse commitment, the intrinsic 13C and deuterium isotope effects are 1.0155 (corresponding to a very early transition state for C-C bond cleavage) and 7.3, and the forward commitment is 3.7. With deoxodihydroprephenate (lacking one double bond in the ring), oxidation occurs without decarboxylation, and one enantiomer has a V/K value 23-fold higher than the other (deuterium isotope effects are 3.6 and 4.1 for fast and slow isomers; V for the fast isomer is 5% and V/K 0.7% those of prephenate). The fully saturated analogue of deoxoprephenate is a very slow substrate (V 0.07% and V/K approximately 10(-5%) those of prephenate). pH profiles show a group with pK = 8.3 that must be protonated for substrate binding and a catalytic group with pK = 6.5 that is a cationic acid (likely histidine). This group facilitates hydride transfer by beginning to accept the proton from the 4-hydroxyl group of prephenate prior to the beginning of C-C cleavage (or fully accepting it in the oxidation of the analogues with only one double bond or none in the ring). In contrast with the enzymatic reaction, the acid-catalyzed decarboxylation of prephenate and deoxoprephenate (t1/2 of 3.7 min at low pH) is a stepwise reaction with a carbonium ion intermediate, since 18O is incorporated into substrate and its epi isomer during reaction in H218O. pH profiles show that the hydroxyl group must be protonated and the carboxyl (pK approximately 4.2) ionized for carbonium ion formation. The carbonium ion formed from prephenate decarboxylates 1.75 times faster than it reacts with water (giving 1.8 times as much prephenate as epi isomer). The observed 13C isotope effect of 1.0082 thus corresponds to an intrinsic isotope effect of 1.023, indicating an early transition state for the decarboxylation step. epi-Prephenate is at least 20 times more stable to acid than prephenate because it exists largely as an internal hemiketal.(ABSTRACT TRUNCATED AT 400 WORDS)  相似文献   

9.
Beta-ketoacyl-acyl carrier protein (ACP) reductase from Mycobacterium tuberculosis (MabA) is responsible for the second step of the type-II fatty acid elongation system of bacteria, plants, and apicomplexan organisms, catalyzing the NADPH-dependent reduction of beta-ketoacyl-ACP to generate beta-hydroxyacyl-ACP and NADP(+). In the present work, the mabA-encoded MabA has been cloned, expressed, and purified to homogeneity. Initial velocity studies, product inhibition, and primary deuterium kinetic isotope effects suggested a steady-state random bi-bi kinetic mechanism for the MabA-catalyzed reaction. The magnitudes of the primary deuterium kinetic isotope effect indicated that the C(4)-proS hydrogen is transferred from the pyridine nucleotide and that this transfer contributes modestly to the rate-limiting step of the reaction. The pH-rate profiles demonstrated groups with pK values of 6.9 and 8.0, important for binding of NADPH, and with pK values of 8.8 and 9.6, important for binding of AcAcCoA and for catalysis, respectively. Temperature studies were employed to determine the activation energy of the reaction. Solvent kinetic isotope effects and proton inventory analysis established that a single proton is transferred in a partially rate-limiting step and that the mechanism of carbonyl reduction is probably concerted. The observation of an inverse (D)2(O)V/K and an increase in (D)2(O)V when [4S-(2)H]NADPH was the varied substrate obscured the distinction between stepwise and concerted mechanisms; however, the latter was further supported by the pH dependence of the primary deuterium kinetic isotope effect. Kinetic and chemical mechanisms for the MabA-catalyzed reaction are proposed on the basis of the experimental data.  相似文献   

10.
Polgár L 《Biochemistry》1999,38(47):15548-15555
Oligopeptidase B, a member of the novel prolyl oligopeptidase family of serine peptidases, is involved in cell invasion by trypanosomes. The kinetic analysis of the reactions of oligopeptidase B, which preferentially cleaves peptides at two adjacent basic residues, has revealed significant differences from the trypsin-like serine peptidases. (i) The pH dependence of k(cat)/K(m) deviates from normal bell-shaped curves due to ionization of an enzymatic group characterized by a macroscopic pK(a) of approximately 8.3. The effect of this group is abolished at high ionic strength. (ii) The second-order acylation rate constants, k(cat)/K(m), are similar with the ester and the corresponding amide substrates, suggesting that their chemical reactivity does not prevail in the rate-limiting step. The kinetic deuterium isotope effects indicate that the rate-limiting step for k(cat)/K(m) is principally governed by conformational changes. (iii) The pH-k(cat)/K(m) profile and the very low rate constant for benzoyl-citrulline ethyl ester reveal a new kinetically influential group ionizing below the pK(a) of the active site histidine and indicate that the positive charge of arginine is essential for effective catalysis. (iv) The enzyme is inhibited by high concentrations of substrate. The mechanism of inhibition markedly varies with the reaction conditions. (v) The optimum temperature for the reactions of amide substrates is unusually low, slightly below 25 degrees C, whereas with benzoyl-arginine ethyl ester a linear Eyring plot is obtained up to 39 degrees C. The positive entropies of activation point to substantial reorganization of water molecules upon substrate binding.  相似文献   

11.
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   

12.
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   

13.
Enzyme elements that are involved in the reversible cyclization of L-carbamylaspartate to L-dihdroorotate catalyzed by dihydroorotase (EC 3.5.2.3) from Clostridium oroticum (ATCC 25750) have been studied. Removal of Zn(II) from the enzyme by chelators followed by incubation of apoenzyme with Co(II) results in replacement of two to three of the four Zn(II) ions per molecule by Co(II). The catalytic properties of the Zn(II)Co(II) dihydroorotase are different from those of native enzyme. The Vmax is increased for both the synthesis and hydrolysis of L-dihydroorotate. The Km for L-dihydroorotate is unchanged, while the Km for L-carbamylaspartate is increased more than twofold. On the other hand, the kinetic properties of Zn(II)-reconstituted dihydroorotase are indistinguishable from those of native enzyme. The pH dependence of Vmax is also altered by the Co(II) substitution. For both Zn(II)- and Zn(II)Co(II)-dihydroorotase, this pH dependence is well described by a single ionization and the pK's for L-dihydroorotate synthesis and hydrolysis are different. Substitution with Co(II) increases the pK for both reaction directions to different extents. These results strongly support a role for the tightly bound metals in the catalytic mechanism. In addition, diethylpyrocarbonate rapidly inactivates the enzyme. The inactivation is prevented by L-dihydroorotate. This result is consistent with a role for at least one histidine in catalysis. The possibility that C. oroticum dihydroorotase may be useful model for the more complex mammalian enzyme is considered.  相似文献   

14.
Hardy LW  Kirsch JF 《Biochemistry》1984,23(6):1282-1287
The solvent kinetic isotope effects (SKIE's) on k(cat) (D(V)) and on k(cat/Km[D(V/K)] were determined for the Bacillus cereus beta-lactamase I catalyzed hydrolysis of five substrates that have values of k(cat)/K(m) varying over the range (0.014-46.3) X 10(6)M(-1) s(-1) and of k(cat) between 0.5 and 2019 s(-1). The variation of D(V/K) was only from 1.06 to 1.25 among these compounds and that in D(V) was from 1.50 to 2.16. These results require that Dk(1), the SKIE on the enzyme-substrate association rate constant, and D(k-1/k2), that on the partition ratio of the ES complex, both be near 1. The larger SKIE observed on D(V) requires that an exchangeable proton be in flight for either or both the acylation and the deacylation reaction. The pH dependence of the values k(cat)/K(m) for three substrates shows identical pK(a)s of 5.5. and 8.4. This identity combined with the fact that only one of these three substrates is kinetically "sticky" proves that the substrates can combine productively with only one protonic form of the enzyme. There is considerable substrate variation in the pK(a) values of k(cat) observed vs. pH profiles; the inflection points for all substrates studied are at pH values more extreme than are observed in the pH profiles for k(cat)/K(m).  相似文献   

15.
Karsten WE  Tipton PA  Cook PF 《Biochemistry》2002,41(40):12193-12199
Tartrate dehydrogenase catalyzes the divalent metal ion- and NAD-dependent oxidative decarboxylation of D-malate to yield CO(2), pyruvate, and NADH. The enzyme also catalyzes the metal ion-dependent oxidation of (+)-tartrate to yield oxaloglycolate and NADH. pH-rate profiles and isotope effects were measured to probe the mechanism of this unique enzyme. Data suggest a general base mechanism with likely general acid catalysis in the oxidative decarboxylation of D-malate. Of interest, the mechanism of oxidative decarboxylation of D-malate is stepwise with NAD(+) or the more oxidizing thio-NAD(+). The mechanism does not become concerted with the latter as observed for the malic enzyme, which catalyzes the oxidative decarboxylation of L-malate [Karsten, W. E., and Cook, P. F. (1994) Biochemistry 33, 2096-2103]. It appears the change in mechanism observed with malic enzyme is specific to its transition state structure and not a generalized trait of metal ion- and NAD(P)-dependent beta-hydroxy acid oxidative decarboxylases. The V/K(malate) pH-rate profile decreases at low and high pH and exhibits pK(a) values of about 6.3 and 8.3, while that for V/K(tartrate) (measured from pH 7.5 to pH 9) exhibits a pK(a) of 8.6 on the basic side. A single pK(a) of 6.3 is observed on the acid side of the V(max) pH profile, but the pK(a) seen on the basic side of the V/K pH profiles is not observed in the V(max) pH profiles. Data suggest the requirement for a general base that accepts a proton from the 2-hydroxyl group of either substrate to facilitate hydride transfer. A second enzymatic group is also required protonated for optimum binding of substrates and may also function as a general acid to donate a proton to the enolpyruvate intermediate to form pyruvate. The (13)C isotope effect, measured on the decarboxylation of D-malate using NAD(+) as the dinucleotide substrate, decreases from a value of 1.0096 +/- 0.0006 with D-malate to 1.00787 +/- 0.00006 with D-malate-2-d, suggesting a stepwise mechanism for the oxidative decarboxylation of D-malate. Using thio-NAD(+) as the dinucleotide substrate the (13)C isotope effects are 1.0034 +/- 0.0007 and 1.0027 +/- 0.0002 with D-malate and D-malate-2-d, respectively.  相似文献   

16.
C B Grissom  W W Cleland 《Biochemistry》1988,27(8):2934-2943
The catalytic mechanism of porcine heart NADP isocitrate dehydrogenase has been investigated by use of the variation of deuterium and 13C kinetic isotope effects with pH. The observed 13C isotope effect on V/K for isocitrate increases from 1.0028 at neutral pH to a limiting value of 1.040 at low pH. The limiting 13C isotope effect with deuteriated isocitrate at low pH is 1.016. This decrease in 13(V/KIc) upon deuteriation indicates a stepwise mechanism for the oxidation and decarboxylation of isocitrate. This predicts a deuterium isotope effect on V/K of 2.9, but D(V/K) at low pH only increases to a maximum of 1.08. It is not known why 13(V/KIc) with deuteriated isocitrate decreases more than predicted. The pK seen in the 13(V/KIc) pH profile for isocitrate is 4.5. This pK is displaced 1.2 pH units from the true pK of the acid/base functionality of 5.7 seen in the pKi profile for oxalylglycine, a competitive inhibitor for isocitrate. From this displacement, catalysis is estimated to be 16 times faster than substrate dissociation. By use of the pH-dependent partitioning ratio of the reaction intermediate oxalosuccinate between decarboxylation to 2-ketoglutarate and reduction to isocitrate, the forward commitment to catalysis for decarboxylation was determined to be 7.3 at pH 5.4 and 3.2 at pH 5.0. This gives an intrinsic 13C isotope effect for decarboxylation of 1.050. 3-Fluoroisocitrate is a new substrate oxidatively decarboxylated by NADP isocitrate dehydrogenase. At neutral pH, D(V/K3-F-Ic) = 1.45 and 13(V/K3-F-Ic) = 1.0129. At pH 5.2, 13(V/K3-F-Ic) increases to 1.0186, indicating that a finite, but diminished, external commitment remains at neutral pH. The product of oxidative decarboxylation of 3-hydroxyisocitrate by NADP isocitrate dehydrogenase is 2-hydroxy-3-ketoglutarate. This results from enzymatic protonation of the cis-enediol intermediate at C2 rather than C3 (as seen with isocitrate and 3-fluoroisocitrate). 2-Hydroxy-3-ketoglutarate further decarboxylates in solution to 2-hydroxy-3-ketobutyrate, which further decarboxylates to acetol. This makes 3-hydroxyisocitrate unsuitable for 13C isotope effect studies.  相似文献   

17.
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   

18.
In the pyrimidine biosynthetic pathway, N-carbamyl-L-aspartate (CA-asp) is converted to L-dihydroorotate (DHO) by dihydroorotase (DHOase). The mechanism of this important reaction was probed using primary and secondary 15N and 13C isotope effects on the ring opening of DHO using isotope ratio mass spectrometry (IRMS). The reaction was performed at three different temperatures (25, 37, and 45 degrees C for hamster DHOase; 37, 50, and 60 degrees C for Bacillus caldolyticus), and the product CA-asp was purified for analysis. The primary and secondary kinetic isotope effects for the ring opening of the DHO were determined from analysis of the N and C of the carbamyl group after hydrolysis. In addition, the beta-carboxyl of the residual aspartate was liberated enzymatically by transamination to oxaloacetate with aspartate aminotransferase and then decarboxylation with oxaloacetate decarboxylase. The 13C/12C ratio from the released CO2 was determined by IRMS, yielding a second primary isotope effect. The primary and secondary isotope effects for the reaction catalyzed by DHOase showed little variation between enzymes or temperatures, the primary 13C and 15N isotope effects being approximately 1% on average, while the secondary 13C isotope effect is negligible or very slightly normal (>1.0000). These data indicate that the chemistry is at least partially rate-limiting while the secondary isotope effects suggest that the transition state may have lost some bending and torsional modes leading to a slight lessening of bond stiffness at the carbonyl carbon of the amide of CA-asp. The equilibrium isotope effects for DHO --> CA-asp have also been measured (secondary 13K(eq) = 1.0028 +/- 0.0002, primary 13K(eq) = 1.0053 +/- 0.0003, primary 15K(eq) = 1.0027 +/- 0.0003). Using these equilibrium isotope effects, the kinetic isotope effects for the physiological reaction (CA-asp --> DHO) have been calculated. These values indicate that the carbon of the amide group is more stiffly bonded in DHO while the slightly lesser, but still normal, values of the primary kinetic isotope effect show that the chemistry remains at least partially rate-limiting for the physiological reaction. It appears that the ring opening and closing is the slow step of the reaction.  相似文献   

19.
Catalytic mechanism of hamster arylamine N-acetyltransferase 2   总被引:1,自引:0,他引:1  
Wang H  Liu L  Hanna PE  Wagner CR 《Biochemistry》2005,44(33):11295-11306
Arylamine N-acetyltransferases (NATs) catalyze an acetyl group transfer from AcCoA to primary arylamines, hydrazines, and hydrazides and play a very important role in the metabolism and bioactivation of drugs, carcinogens, and other xenobiotics. The reaction follows a ping-pong bi-bi mechanism. Structure analysis of bacterial NATs revealed a Cys-His-Asp catalytic triad that is strictly conserved in all known NATs. Previously, we have demonstrated by kinetic and isotope effect studies that acetylation of the hamster NAT2 is dependent on a thiolate-imidazolium ion pair (Cys-S(-)-His-ImH(+)) and not a general acid-base catalysis. In addition, we established that, after formation of the acetylated enzyme intermediate, the active-site imidazole, His-107, is likely deprotonated at physiological pH. In this paper, we report steady-state kinetic studies of NAT2 with two acetyl donors, acetyl coenzyme A (AcCoA) and p-nitrophenyl acetate (PNPA), and four arylamine substrates. The pH dependence of k(cat)/K(AcCoA) exhibited two inflection points at 5.32 +/- 0.13 and 8.48 +/- 0.24, respectively. The pK(a) at 5.32 is virtually identical with the previously reported pK(a) of 5.2 for enzyme acetylation, reaffirming that the first half of the reaction is catalyzed by a thiolate-imidazolium ion pair in the active site. The inflection point at 8.48 indicates that a pH-sensitive group on NAT2 is involved in AcCoA binding. A Br?nsted plot constructed by the correlation of log k(4) and log k(H)2(O) with the pK(a) for each arylamine substrate and water displays a linear free-energy relationship in the pK(a) range from -1.7 (H(2)O) to 4.67 (PABA), with a slope of beta(nuc) = 0.80 +/- 0.1. However, a further increase of the pK(a) from 4.67 (PABA) to 5.32 (anisidine) resulted in a 2.5-fold decrease in the k(4) value. Analysis of the pH-k(cat)/K(PABA) profile revealed a pK(a) of 5.52 +/- 0.14 and a solvent kinetic isotope effect (SKIE) of 2.01 +/- 0.04 on k(cat)/K(PABA). Normal solvent isotope effects of 4.8 +/- 0.1, 3.1 +/- 0.1, and 3.2 +/- 0.1 on the k(cat)/K(b) for anisidine, pABglu, and PNA, respectively, were also determined. These observations are consistent with a deacetylation mechanism dominated by nucleophilic attack of the thiol ester for arylamines with pK(a) values or=5.5. The general base is likely His-107 because the His-107 to Gln and Asn mutants were found to be devoid of catalytic activity. In contrast, an increase in pH-dependent hydrolysis of the acetylated enzyme was not observed over a pH range of 5.2-7.5. On the basis of these observations, a catalytic mechanism for the acetylation of arylamines by NAT2 is proposed.  相似文献   

20.
3alpha-hydroxysteroid dehydrogenase/carbonyl reductase from Comamonas testosteroni catalyzes the oxidation of androsterone with NAD(+) to form androstanedione and NADH with a concomitant releasing of protons to bulk solvent. To probe the proton transfer during the enzyme reaction, we used mutagenesis, chemical rescue, and kinetic isotope effects to investigate the release of protons. The kinetic isotope effects of (D)V and (D(2)O)V for wild-type enzyme are 1 and 2.1 at pL 10.4 (where L represents H, (2)H), respectively, and suggest a rate-limiting step in the intramolecular proton transfer. Substitution of alanine for Lys(159) changes the rate-limiting step to the hydride transfer, evidenced by an equal deuterium isotope effect of 1.8 on V(max) and V/K(androsterone) and no solvent kinetic isotope effect at saturating 3-(cyclohexylamino)propanesulfonic acid (CAPS). However, a value of 4.4 on V(max) is observed at 10 mm CAPS at pL 10.4, indicating a rate-limiting proton transfer. The rate of the proton transfer is blocked in the K159A and K159M mutants but can be rescued using exogenous proton acceptors, such as buffers, small primary amines, and azide. The Br?nsted relationship between the log(V/K(d)(-base)Et) of the external amine (corrected for molecular size effects) and pK(a) is linear for the K159A mutant-catalyzed reaction at pH 10.4 (beta = 0.85 +/- 0.09) at 5 mm CAPS. These results show that proton transfer to the external base with a late transition state occurred in a rate-limiting step. Furthermore, a proton inventory on V/Et is bowl-shaped for both the wild-type and K159A mutant enzymes and indicates a two-proton transfer in the transition state from Tyr(155) to Lys(159) via 2'-OH of ribose.  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