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 multiple isotope effect method of Hermes et al. [Hermes, J. D., Roeske, C. A., O'Leary, M. H., & Cleland, W. W. (1982) Biochemistry 21, 5106-5114] has been used to study the mechanism of the oxidative decarboxylation catalyzed by 6-phosphogluconate dehydrogenase from yeast. 13C kinetic isotope effects of 1.0096 and 1.0081 with unlabeled or 3-deuterated 6-phosphogluconate, plus a 13C equilibrium isotope effect of 0.996 and a deuterium isotope effect on V/K of 1.54, show that the chemical reaction after the substrates have bound is stepwise, with hydride transfer preceding decarboxylation. The kinetic mechanism of substrate addition is random at pH 8, since the deuterium isotope effect is the same when either NADP or 6-phosphogluconate or 6-phosphogluconate-3-d is varied at fixed saturating levels of the other substrate. Deuterium isotope effects on V and V/K decrease toward unity at high pH at the same time that V and V/K are decreasing, suggesting that proton removal from the 3-hydroxyl may precede dehydrogenation. Comparison of the tritium effect of 2.05 with the other measured isotope effects gives limits of 3-4 on the intrinsic deuterium and of 1.01-1.05 for the intrinsic 13C isotope effect for C-C bond breakage in the forward direction and suggests that reverse hydride transfer is 1-4 times faster than decarboxylation.  相似文献   

2.
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   

3.
Although aldolase-catalyzed condensations proceed by stepwise mechanisms via the intermediacy of nucleophilic enol(ate)s or enamines, the mechanisms of those enzymes that catalyze Claisen-type condensations are unclear. The reaction pathway followed by an enzyme from this second group, malate synthase, has been studied by the double-isotope fractionation method to determine whether the reaction is stepwise or concerted. In agreement with earlier work, a deuterium kinetic isotope effect D(V/K) of 1.3 +/- 0.1 has been found when [2H3]acetyl-CoA is the substrate. The 13C isotope effect at the aldehydic carbon of glyoxylate has also been measured. For this determination, the malate product (containing the carbon of interest at C-2) was quantitatively transformed into a new sample of malate having the carbon of interest at C-4. This material was decarboxylated by malic enzyme to produce the appropriate CO2 for isotope ratio mass spectrometric analysis. The 13C isotope effect with [1H3]acetyl-CoA [that is, 13(V/K)H] is 1.0037 +/- 0.0004. By use of the known values of the intermolecular and intramolecular deuterium effects and of 13(V/K)H, the value of the 13C isotope effect when deuteriated [2H3]acetyl-CoA is the substrate [that is, 13(V/K)D] can be predicted for three possible mechanisms. If 13(V/K)H is a kinetic isotope effect and the reaction is concerted, the value of the 13C effect on deuteriation of acetyl-CoA will rise to 1.011; if 13(V/K)H is a kinetic isotope effect and the reaction is stepwise, the value of the 13C effect will fall to 1.0025; and if the 13C effect is an equilibrium isotope effect deriving from glyoxylate dehydration, the reaction is necessarily stepwise, and the value of 13(V/K)D will be 1.0037, unchanged from that of 13(V/K)H. Experimentally, the value of 13(V/K)D is 1.0037 +/- 0.0007, which requires that malate synthase follow a stepwise path. It is therefore clear that the two salient characteristics of enzymes that catalyze Claisen-like condensations, namely, the absence of enzyme-catalyzed proton exchange with solvent and the inversion of the configuration at the nucleophilic center, which had been suggestive of a concerted pathway, are not mechanistically diagnostic.  相似文献   

4.
W L Sweet  J S Blanchard 《Biochemistry》1991,30(35):8702-8709
Kinetic parameters and primary deuterium kinetic isotope effects for NADH and five pyridine nucleotide substrates have been determined at pH 8.1 for human erythrocyte glutathione reductase. DV/KNADH and DV are equal to 1.4 and are pH independent below pH 8.1, but DV decreases to 1.0 at high pH as a group exhibiting a pK of 8.6 is deprotonated. This result suggests that as His-467' is deprotonated, the rate of the isotopically insensitive oxidative half-reaction is specifically decreased and becomes rate-limiting. For all substrates, equivalent V and V/K primary deuterium kinetic isotope effects are observed at pH values below 8.1. The primary deuterium kinetic isotope effect on V, but not V/K, is sensitive to solvent isotopic composition. The primary tritium kinetic isotope effects agree well with the corresponding value calculated from the primary deuterium kinetic isotope effects by using the Swain-Schaad relationship. This suggests that the primary deuterium kinetic isotope effects observed in these steady-state experiments are the intrinsic primary deuterium kinetic isotope effects for hydride transfer. The magnitude of the primary deuterium kinetic isotope effect is dependent on the redox potential of the pyridine nucleotide substrate used, varying from approximately 1.4 for NADH and -320 mV reductants to 2.7 for thioNADH to 4.2-4.8 for 3-acetylpyridine adenine dinucleotide (3APADH). The alpha-secondary tritium kinetic isotope effects also increase as the redox potential of the pyridine nucleotide substrate becomes more positive. Together, these data indicate that the transition state for hydride transfer is very early for NADH and becomes later for thioNADH and 3APADH, as predicted by Hammond's postulate.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

5.
P F Canellas  W W Cleland 《Biochemistry》1991,30(36):8871-8876
Carbon-13 and deuterium isotope effects have been measured on the reaction catalyzed by rabbit muscle glyceraldehyde-3-phosphate dehydrogenase in an effort to locate the rate-limiting steps. With D-glyceraldehyde 3-phosphate as substrate, hydride transfer is a major, but not the only, slow step prior to release of the first product, and the intrinsic primary deuterium and 13C isotope effects on this step are 5-5.5 and 1.034-1.040, and the sum of the commitments to catalysis is approximately 3. The 13C isotope effects on thiohemiacetal formation and thioester phosphorolysis are 1.005 or less. The intrinsic alpha-secondary deuterium isotope effect at C-4 of the nicotinamide ring of NAD is approximately 1.4; this large normal value (the equilibrium isotope effect is 0.89) shows tight coupling of hydrogen motions in the transition state accompanied by tunneling. With D-glyceraldehyde as substrate, the isotope effects are similar, but the sum of commitments is approximately 1.5, so that hydride transfer is more, but still not solely, rate limiting for this slow substrate. The observed 13C and deuterium equilibrium isotope effects on the overall reaction from the hydrated aldehyde are 0.995 and 1.145, while the 13C equilibrium isotope effect for conversion of a thiohemiacetal to a thioester is 0.994, and that for conversion of a thioester to an acyl phosphate is 0.997. Somewhat uncertain values for the 13C equilibrium isotope effects on aldehyde dehydration and formation of a thiohemiacetal are 1.003 and 1.004.  相似文献   

6.
Primary intrinsic deuterium and 13C isotope effects have been determined for liver (LADH) and yeast (YADH) alcohol dehydrogenases with benzyl alcohol as substrate and for yeast aldehyde dehydrogenase (ALDH) with benzaldehyde as substrate. These values have also been determined for LADH as a function of changing nucleotide substrate. As the redox potential of the nucleotide changes from -0.320 V with NAD to -0.258 V with acetylpyridine-NAD, the product of primary and secondary deuterium isotope effects rises from 4 toward 6.5, while the primary 13C isotope effect drops from 1.025 to 1.012, suggesting a trend from a late transition state with NAD to one that is more symmetrical. The values of Dk (again the product of primary and secondary isotope effects) and 13k for YADH with NAD are 7 and 1.023, suggesting for this very slow reaction a more stretched, and thus symmetrical, transition state. With ALDH and NAD, the primary 13C isotope effect on the hydride transfer step lies in the range 1.3-1.6%, and the alpha-secondary deuterium isotope effect on the same step is at least 1.22, but 13C isotope effects on formation of the thiohemiacetal intermediate and on the addition of water to the thio ester intermediate are less than 1%. On the basis of the relatively large 13C isotope effects, we conclude that carbon motion is involved in the hydride transfer steps of dehydrogenase reactions.  相似文献   

7.
Homoisocitrate dehydrogenase (HIcDH, 3-carboxy-2-hydroxyadipate dehydrogenase) catalyzes the fourth reaction of the alpha-aminoadipate pathway for lysine biosynthesis, the conversion of homoisocitrate to alpha-ketoadipate using NAD as an oxidizing agent. A chemical mechanism for HIcDH is proposed on the basis of the pH dependence of kinetic parameters, dissociation constants for competitive inhibitors, and isotope effects. According to the pH-rate profiles, two enzyme groups act as acid-base catalysts in the reaction. A group with a p K a of approximately 6.5-7 acts as a general base accepting a proton as the beta-hydroxy acid is oxidized to the beta-keto acid, and this residue participates in all three of the chemical steps, acting to shuttle a proton between the C2 hydroxyl and itself. The second group acts as a general acid with a p K a of 9.5 and likely catalyzes the tautomerization step by donating a proton to the enol to give the final product. The general acid is observed in only the V pH-rate profile with homoisocitrate as a substrate, but not with isocitrate as a substrate, because the oxidative decarboxylation portion of the isocitrate reaction is limiting overall. With isocitrate as the substrate, the observed primary deuterium and (13)C isotope effects indicate that hydride transfer and decarboxylation steps contribute to rate limitation, and that the decarboxylation step is the more rate-limiting of the two. The multiple-substrate deuterium/ (13)C isotope effects suggest a stepwise mechanism with hydride transfer preceding decarboxylation. With homoisocitrate as the substrate, no primary deuterium isotope effect was observed, and a small (13)C kinetic isotope effect (1.0057) indicates that the decarboxylation step contributes only slightly to rate limitation. Thus, the chemical steps do not contribute significantly to rate limitation with the native substrate. On the basis of data from solvent deuterium kinetic isotope effects, viscosity effects, and multiple-solvent deuterium/ (13)C kinetic isotope effects, the proton transfer step(s) is slow and likely reflects a conformational change prior to catalysis.  相似文献   

8.
Benzoylformate decarboxylase (benzoylformate carboxy-lyase, BFD; EC 4.1.1.7) from Pseudomonas putida is a thiamine pyrophosphate (TPP) dependent enzyme which converts benzoylformate to benzaldehyde and carbon dioxide. The kinetics and mechanism of the benzoylformate decarboxylase reaction were studied by solvent deuterium and 13C kinetic isotope effects with benzoylformate and a series of substituted benzoylformates (pCH3O, pCH3, pCl, and mF). The reaction was found to have two partially rate-determining steps: initial tetrahedral adduct formation (D2O sensitive) and decarboxylation (13C sensitive). Solvent deuterium and 13C isotope effects indicate that electron-withdrawing substituents (pCl and mF) reduce the rate dependence upon decarboxylation such that decreased 13(V/K) effects are observed. Conversely, electron-donating substituents increase the rate dependence upon decarboxylation such that a larger 13(V/K) is seen while the D2O effects on V and V/K are not dramatically different from those for benzoylformate. All of the data are consistent with substituent stabilization or destabilization of the carbanionic intermediate (or carbanion-like transition state) formed during decarboxylation. Additional information regarding the mechanism of the enzymic reaction was obtained from pH studies on the reaction of benzoylformate and the binding of competitive inhibitors. These studies suggest that two enzymic bases are required to be in the correct protonation state (one protonated and one unprotonated) for optimal binding of substrate (or inhibitors).  相似文献   

9.
Karsten WE  Ohshiro T  Izumi Y  Cook PF 《Biochemistry》2005,44(48):15930-15936
Serine-glyoxylate aminotransferase (SGAT) from Hyphomicrobium methylovorum is a pyridoxal 5'-phosphate (PLP) enzyme that catalyzes the interconversion of L-serine and glyoxylate to hydroxypyruvate and glycine. The primary deuterium isotope effect using L-serine 2-D is one on (V/K)serine and V in the steady state. Pre-steady-state experiments also indicate that there is no primary deuterium isotope effect with L-serine 2-D. The results suggest there is no rate limitation by abstraction of the alpha proton of L-serine in the SGAT reaction. In the steady-state a solvent deuterium isotope effect of about 2 was measured on (V/K)L-serine and (V/K)ketomalonate and about 5.5 on V. Similar solvent isotope effects were observed in the pre-steady-state for the natural substrates and the alternative substrate ketomalonate. In the pre-steady-state, no reaction intermediates typical of PLP enzymes were observed with the substrates L-serine, glyoxylate, and hydroxypyruvate. The data suggest that breakdown and formation of the ketimine intermediate is the primary rate-limiting step with the natural substrates. In contrast, using the alternative substrate ketomalonate, pre-steady-state experiments display the transient formation of a 490 nm absorbing species typical of a quinonoid intermediate. The solvent isotope effect results also suggest that with ketomalonate as substrate protonation at C(alpha) is the slowest step in the SGAT reaction. This is the first report of a rate-limiting protonation of a quinonoid at C(alpha) of the external Schiff base in an aminotransferase reaction.  相似文献   

10.
S J O'Keefe  J R Knowles 《Biochemistry》1986,25(20):6077-6084
To investigate the mechanism of the carboxylation of pyruvate to oxalacetate catalyzed by the enzyme transcarboxylase, we have measured the D(V/K) and 13(V/K) isotope effects. Comparison of the double-reciprocal plots of the initial velocities with [1H3]pyruvate and with [2H3]pyruvate as substrate yields a deuterium isotope effect on Vmax/Km of 1.39 +/- 0.04. The 13C kinetic isotope effect on the carboxylation of pyruvate to oxalacetate has been measured by the competitive method and is 1.0227 +/- 0.0008. To determine whether the removal of the proton from pyruvate and the addition of the carboxyl group occur in the same or in different steps, the double-isotope fractionation test has been used. When [2H3]pyruvate replaces [1H3]pyruvate as the substrate, the observed 13(V/K) isotope effect falls from 1.0227 to 1.0141 +/- 0.001. The latter value is in excellent agreement with the value of 1.0136, predicted for a stepwise pathway. We may conclude, therefore, that the carboxylation of pyruvate catalyzed by transcarboxylase proceeds by a stepwise mechanism involving the intermediate formation of the substrate carbanion.  相似文献   

11.
J D Hermes  P M Weiss  W W Cleland 《Biochemistry》1985,24(12):2959-2967
Phenylalanine ammonia-lyase has been shown to catalyze the elimination of ammonia from the slow alternate substrate 3-(1,4-cyclohexadienyl)alanine by an E1 cb mechanism with a carbanion intermediate. This conclusion resulted from comparison of 15N isotope effects with deuterated (0.9921) and unlabeled substrates (1.0047), and a deuterium isotope effect of 2.0 from dideuteration at C-3, with the equations for concerted, carbanion, and carbonium ion mechanisms. The 15N equilibrium isotope effect on the addition of the substrate to the dehydroalanine prosthetic group on the enzyme is 0.979, while the kinetic 15N isotope effect on the reverse of this step is 1.03-1.04 and the intrinsic deuterium isotope effect on proton removal is in the range 4-6. Isotope effects with phenylalanine itself are small (15N ones of 1.0021 and 1.0010 when unlabeled or 3-dideuterated and a deuterium isotope effect of 1.15) but are consistent with the same mechanism with drastically increased commitments, including a sizable external one (i.e., phenylalanine is sticky). pH profiles show that the amino group of the substrate must be unprotonated to react but that a group on the enzyme with a pK of 9 must be protonated, possibly to catalyze addition of the substrate to dehydroalanine. Incorrectly protonated enzyme-substrate complexes do not form. Equilibrium 15N isotope effects are 1.016 for the deprotonation of phenylalanine or its cyclohexadienyl analogue, 1.0192 for deprotonation of NH4+, 1.0163 for the conversion of the monoanion of phenylalanine to NH3, and 1.0138 for the conversion of the monoanion of aspartate to NH4+.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

12.
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   

13.
S M Miller  J P Klinman 《Biochemistry》1983,22(13):3091-3096
Intrinsic primary hydrogen isotope effects (kH/kD) have been obtained for the carbon-hydrogen bond cleavage step catalyzed by dopamine beta-monooxygenase. Irreversibility of this step is inferred from the failure to observe back-exchange of tritium from TOH into substrate under conditions of dopamine turnover; this result cannot be due to solvent inaccessibility at the enzyme active site, since we will demonstrate [Ahn, N., & Klinman, J. P. (1983) Biochemistry (following paper in this issue)] that a solvent-derived proton or triton must be at the enzyme active site prior to substrate activation. As shown by Northrop [Northrop, D. B. (1975) Biochemistry 14, 2644], for enzymatic reactions in which the carbon-hydrogen bond cleavage step is irreversible, comparison of D(V/K) to T(V/K) allows an explicit solution for kH/kD. Employing a double-label tracer method, we have been able to measure deuterium isotope effects on Vmax/Km with high precision, D(V/K) = 2.756 +/- 0.054 at pH 6.0. The magnitude of the tritium isotope effect under comparable experimental conditions is T(V/K) = 6.079 +/- 0.220, yielding kH/kD = 9.4 +/- 1.3. This result was obtained in the presence of saturating concentrations of the anion activator fumarate. Elimination of fumarate from the reaction mixture leads to high observed values for isotope effects on Vmax/Km, together with an essentially invariant value for kH/kD = 10.9 +/- 1.9. Thus, the large disparity between isotope effects, plus or minus fumarate, cannot be accounted for by a change in kH/kD, and we conclude a role for fumarate in the modulation of the partitioning of enzyme-substrate complex between catalysis and substrate dissociation. On the basis of literature correlations of primary hydrogen isotope effects and the thermodynamic properties of hydrogen transfer reactions, the very large magnitude of kH/kD = 9.4-10.9 for dopamine beta-monooxygenase suggests an equilibrium constant not very far from unity for the carbon-hydrogen bond cleavage step. This feature, together with the failure to observe re-formation of dopamine from enzyme-bound intermediate or product and overall rate limitation of enzyme turnover by product release, leads us to propose a stepwise mechanism for norepinephrine formation from dopamine in which carbon-hydrogen bond cleavage is uncoupled from the oxygen insertion step.  相似文献   

14.
Zheng R  Blanchard JS 《Biochemistry》2003,42(38):11289-11296
Ketopantoate reductase (EC 1.1.1.169), an enzyme in the pantothenate biosynthetic pathway, catalyzes the NADPH-dependent reduction of alpha-ketopantoate to form D-(-)-pantoate. The enzyme exhibits high specificity for ketopantoate, with V and V/K for ketopantoate being 5- and 365-fold higher than those values for alpha-ketoisovalerate and 20- and 648-fold higher than those values for alpha-keto-beta-methyl-n-valerate, respectively. For pyridine nucleotides, V/K for beta-NADPH is 3-500-fold higher than that for other nucleotide substrates. The magnitude of the primary deuterium kinetic isotope effects on V and V/K varied substantially when different ketoacid and pyridine nucleotide substrates were used. The small primary deuterium kinetic isotope effects observed using NADPH and NHDPH suggest that the chemical step is not rate-limiting, while larger primary deuterium isotope effects were observed for poor ketoacid and pyridine nucleotide substrates, indicating that the chemical reaction has become partially or completely rate-limiting. The pH dependence of (D)V using ketopantoate was observed to vary from a value of 1.1 at low pH to a value of 2.5 at high pH, while the magnitude of (D)V/K(NADPH) and (D)V/K(KP) were pH-independent. The value of (D)V is large and pH-independent when alpha-keto-beta-methyl-n-valerate was used as the ketoacid substrate. Solvent kinetic isotope effects of 2.2 and 1.2 on V and V/K, respectively, were observed with alpha-keto-beta-methyl-n-valerate. Rapid reaction analysis of NADPH oxidation using ketopantoate showed no "burst" phase, suggesting that product-release steps are not rate-limiting and the cause of the small observed kinetic isotope effects with this substrate pair. Large primary deuterium isotope effects on V and V/K using 3-APADPH in steady-state experiments, equivalent to the isotope effect observed in single turnover studies, suggests that chemistry is rate-limiting for this poorer reductant. These results are discussed in terms of a kinetic and chemical mechanism for the enzyme.  相似文献   

15.
Since hydride transfer is completely rate limiting for yeast formate dehydrogenase [Blanchard, J.S., & Cleland, W. W. (1980) Biochemistry 19, 3543], the intrinsic isotope effects on this reaction are fully expressed. Primary deuterium, 13C, and 18O isotope effects in formate and the alpha-secondary deuterium isotope effect at C-4 of the nucleotide have been measured for nucleotide substrates with redox potentials varying from -0.320 (NAD) to -0.258 V (acetylpyridine-NAD). As the redox potential gets more positive, the primary deuterium isotope effect increases from 2.2 to 3.1, the primary 13C isotope effect decreases from 1.042 to 1.036, the alpha-secondary deuterium isotope effect drops from 1.23 to 1.06, and Vmax decreases. The 18O isotope effects increase from 1.005 to 1.008 per single 18O substitution in formate (these values are dominated by the normal isotope effect on the dehydration of formate during binding; pyridinealdehyde-NAD gives an inverse value, possibly because it is not fully dehydrated during binding). These isotope effects suggest a progression toward earlier transition states as the redox potential of the nucleotide becomes more positive, with NAD having a late and acetyl-pyridine-NAD a nearly symmetrical transition state. By contrast, the I2 oxidation of formate in dimethyl sulfoxide has a very early transition state (13k = 1.0154; Dk = 2.2; 18k = 0.9938), which becomes later as the proportion of water in the solvent increases (13k = 1.0265 in 40% dimethyl sulfoxide and 1.0362 in water). alpha-secondary deuterium isotope effects with formate dehydrogenase are decreased halfway to the equilibrium isotope effect when deuterated formate is the substrate, showing that the bending motion of the secondary hydrogen is coupled to hydride transfer in the transition state and that tunneling of the two hydrogens is involved. The 15N isotope effect of 1.07 for NAD labeled at N-1 of the nicotinamide ring suggests that N-1 becomes pyramidal during the reaction. 18O fractionation factors for formate ion relative to aqueous solution are 1.0016 in sodium formate crystal, 1.0042 bound to Dowex-1, and 1.0040 as an ion pair (probably hydrated) in CHCl3. The CO2 analogue azide binds about 10(4) times better than the formate analogue nitrate to enzyme-nucleotide complexes (even though the Ki values for both and the affinity for formate vary by 2 orders of magnitude among the various nucleotides), but the ratio is not sensitive to the redox potential of the nucleotide. Thus, not the nature of the transition state but rather the shape of the initial binding pocket for formate is determining the relative affinity.(ABSTRACT TRUNCATED AT 400 WORDS)  相似文献   

16.
Use of isotope effects to elucidate enzyme mechanisms   总被引:5,自引:0,他引:5  
The chemical bond breaking steps are normally not rate limiting for enzymatic reactions. However, comparison of deuterium and tritium isotope effects on the same reaction, especially when coupled with 13C isotope effects for the same step measured with deuterated as well as unlabeled substrates, allows calculation of the intrinsic isotope effects on the bond breaking steps and thus a determination of the commitments to catalysis for the reactants. The variation in observed isotope effects as a function of reactant concentration can be used to determine kinetic mechanisms, while the pH variation of isotope effects can determine the stickiness of the reactants and which portions of the reactant mechanism are pH dependent. Finally the size of primary and secondary intrinsic isotope effects can be used to determine transition state structure.  相似文献   

17.
Isotope effects have been measured with Mg2+ as the activator and L-malate labeled with deuterium or tritium at carbon 2 as the substrate over the pH range 4-10. Comparison of the nearly pH-independent deuterium-isotope effect on V/Kmalate of 1.5 with the tritium effect of 2.0 by the method of Northrop (Northrop, D.B. (1975), Biochemistry 14, 2644) gives limits on the true effect of deuterium substitution on the bond-breaking step of 5-8 in the forward reaction and 4-6.5 in the reverse direction. Comparison of the deuterium effect on V/K with the 13C-isotope effect of 1.031 reported by Schimerlik et al. (Schimerlik, M.I., Rife, J.E., and Cleland, W.W. (1975), Biochemistry 14, 5347) allows the deduction that at pH 8 reverse hydride transfer is six to eight times faster than decarboxylation, which is thus largely rate limiting for the catalytic reaction. The absence of a deuterium-isotope effect on V at pH 7-8 and comparison of the Ki of pyruvate as an uncompetitive inhibitor of the forward reaction and a substrate for the reverse reaction indicate that at neutral pH the release of TPNH from enzyme-reduced triphosphopyridine nucleotide (E-TPNH) is the rate-limiting step in the forward direction. The observation of a deuterium effect on V that approaches 3 at pH 4 and 10 shows, however, that, at very low and very high pH, hydride transfer may become partly rate limiting. In the reverse reaction, the probable rate-limiting step at pH 7 is the isomerization of E-TPNH, while at pH 8.5 and above V becomes too large to measure and appears infinite. Substitution of Co2+, Ni2+, or low levels of Mn2+ for Mg2+ gives similar deuterium-isotope effects, although the values of V and Kmalate vary considerably with metal. The kinetics of Mn2+ show pronounced negative cooperativity, with Km values of 7 muM and 5 mM for concentration ranges from 4 to 100 muM and over 1 mM.  相似文献   

18.
B J Bahnson  V E Anderson 《Biochemistry》1989,28(10):4173-4181
The primary, alpha-secondary, beta-secondary, and beta'-secondary deuterium and primary 18O kinetic isotope effects on V/K for the dehydration of [(3S)-3-hydroxybutyryl]pantetheine by bovine liver crotonase (enoyl-CoA hydratase, EC 4.2.1.17) have been determined by the equilibrium perturbation method. The primary deuterium and 18O kinetic isotope effects are 1.61 and 1.051, respectively. The secondary deuterium effects at C-2, C-3, and C-4 are 1.12, 1.13, and 1.00 per H, respectively. The large 18O isotope effect suggests C-O bond cleavage is largely rate determining but is consistent with either an E1cb or E2 mechanism with a large amount of carbanion character. The beta-secondary effect is a factor of 1.05 greater than the equilibrium isotope effect, indicating that this C-H bond is less stiff in the affected transition state or that its motion is coupled to the reaction coordinate motion. Analytical solutions to the differential equations describing uni-uni equilibrium perturbations are presented.  相似文献   

19.
D M Kiick  R S Phillips 《Biochemistry》1988,27(19):7333-7338
The pH dependence of the kinetic parameters and primary deuterium isotope effects have been determined for tyrosine phenol-lyase from both Erwinia herbicola and Citrobacter freundii. The primary deuterium isotope effects indicate that proton abstraction from the 2-position of the substrate is partially rate-limiting for both enzymes. The C. freundii enzyme primary deuterium isotope effects [DV = 3.5 and D(V/Ktyr) = 2.5] are pH independent, indicating that tyrosine is not sticky (i.e., does not dissociate slower than it reacts to give products). Since Vmax for both tyrosine and the alternate substrate S-methyl-L-cysteine is also pH independent, substrate binds only to the correctly protonated form of the enzyme. For the E. herbicola enzyme, both Vmax and V/K for tyrosine or S-methyl-L-cysteine are pH dependent, as well as both DV and D(V/Ktyr). Thus, while both the protonated and unprotonated enzyme can bind substrate, and may be interconverted directly, only the unprotonated Michaelis complex is catalytically competent. At pH 9.5, DV = 2.5 and D(V/Ktyr) = 1.5. However, at pH 6.4 the isotope effect on both parameters is equal to 4.1. From these data, the forward commitment factor (cf = 5.2) and catalytic ratio (cvf = 1.1) for tyrosine and S-methyl-L-cysteine (cf = 2.2, cvf = 24) are calculated. Also, the Michaelis complex partition ratio (cf/cvf) for substrate and products is calculated to be 4.7 for tyrosine and 0.1 for S-methyl-L-cysteine.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

20.
The isotope effects for the interconversion of L-proline and D-proline, catalyzed by proline racemase, have been determined in the saturated region with both [2-2H]proline and [2-3H]proline. The deuterium fractionation factors for each of the protons in flight have been obtained from two kinds of experiment: by measuring the rate of racemization of one [2-2H]proline enantiomer as it racemizes into an equilibrated pool of unlabeled proline and by measuring the deuterium content of a proline sample at the optical rotation maximum that occurs when an equimolar mixture of one deuterium-labeled enantiomer and the other unlabeled enantiomer runs to equilibrium. The tritium fractionation factors for each of the protons in flight have been determined from measurements of the rate of loss of tritium to the solvent as one [2-3H]proline enantiomer runs to equilibrium. Good agreement is found among the fractionation factors determined by each method. The deuterium fractionation factors for the two protons are not identical: that for the proton derived from L-proline is 0.375 and that for the proton derived from D-proline is 0.44. This difference has been confirmed by a double-competition experiment in which the optical rotation of a mixture of DL-[2-2H]proline and unlabeled DL-proline is followed with time. The rotation (initially zero) passes through a maximum, from which the ratio of the two fractionation factors (0.86) is obtained. These data, coupled with the equilibrium fractionation factor for the 2-position of proline (which has been determined to be 1.17), provide the transition-state factors for each of the in-flight protons, and delineate the nature of the transition state(s) for the enzyme-catalyzed racemization.  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