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.
Steady-state kinetic studies for the reaction of the flavocytochrome p-cresol methylhydroxylase with the reducing substrates (S) p-cresol, 4-ethylphenol, and their corresponding alpha-deuteriated analogues are presented. The results from these experiments and those from studies involving various reoxidizing substrates support the proposed apparent ping-pong mechanism. With phenazine methosulfate (PMS) as the reoxidant for studies at pH 7.6 and 6 or 25 degrees C, the isotope effects on kcat are lower than the intrinsic isotope effect. The values for D(kcat/KS) are equal to the intrinsic effect for p-cresol at 25 degrees C and for 4-ethylphenol at both 6 and 25 degrees C. However, the value for this steady-state parameter at 6 degrees C for p-cresol is lower than the intrinsic effect. The values for D(kcat/KPMS) are nearly equal to 1.0 under all conditions. In contrast, the steady-state kinetic analysis for the isolated flavoprotein subunit of p-cresol methylhydroxylase involving p-cresol and PMS as substrates indicates that a random-binding mechanism is operating. Additionally, several of the steady-state parameters yield values for the apparent intrinsic isotope effect for the flavoprotein. The results of stopped-flow kinetic studies are also reported. At pH 7.6 the intrinsic isotope effect (Dk2) for the reduction of the enzyme by 4-ethylphenol is 4.8-5.0 at 25 degrees C and 4.0 at 6 degrees C. This technique yields a value for Dk2 of 7.05 at 6 degrees C and pH 7.6 for p-cresol.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

2.
We have measured the 13C kinetic isotope effect at pH 4.0, 5.0, 6.0, and 6.5 and in D2O at pD 5.0 and the rate of D-H exchange of the alpha and beta protons of aspartic acid in D2O at pD 5.0 for the reaction catalyzed by the enzyme aspartate beta-decarboxylase from Alcaligenes faecalis. The 13C kinetic isotope effect, with a value of 1.0099 +/- 0.0002 at pH 5.0, is less than the intrinsic isotope effect for the decarboxylation step, indicating that the decarboxylation step is not entirely rate limiting. We have been able to estimate probable values of the relative free energies of the transition states of the enzymatic reaction up to and including the decarboxylation step from the 13C kinetic isotope effect and the rate of D-H exchange of alpha-H. The pH dependence of the kinetic isotope effect reflects the pKa of the pyridine nitrogen of the coenzyme pyridoxal 5'-phosphate but not that of the imine nitrogen. A mechanism is proposed for the exchange of aspartate beta-H that is consistent with the stereochemistry suggested earlier.  相似文献   

3.
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   

4.
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   

5.
C B Grissom  W W Cleland 《Biochemistry》1988,27(8):2927-2934
The role of the metal ion in the oxidative decarboxylation of malate by chicken liver NADP malic enzyme and details of the reaction mechanism have been investigated by 13C isotope effects. With saturating NADP and the indicated metal ion at a total concentration 10-fold higher than its Km, the following primary 13C kinetic isotope effects at C4 of malate [13(V/Kmal)] were observed at pH 8.0: Mg2+, 1.0336; Mn2+, 1.0365; Cd2+, 1.0366; Zn2+, 1.0337; Co2+, 1.0283; Ni2+, 1.025. Knowing the partitioning of the intermediate oxalacetate between decarboxylation to pyuvate and reduction to malate allows calculation of the intrinsic carbon isotope effect for decarboxylation. For Mg2+ as activator, this was 1.049 with NADP and 1.046 with 3-acetylpyridine adenine dinucleotide phosphate, although the intrinsic primary deuterium isotope effects on dehydrogenation were 5.6 and 4.2, and the partition ratios of the oxalacetate intermediate for decarboxylation as opposed to hydride transfer were 0.11 and 3.96 (the result of the different redox potentials of NADP and the acetylpyridine analogue). The close agreement of the intrinsic 13C isotope effects with each other and with the 13C isotope effect for the Mg2+-catalyzed nonenzymatic decarboxylation of oxalacetate of 1.0489 [Grissom, C. B., & Cleland, W. W. (1986) J. Am. Chem. Soc. 108, 5582] indicates a similarity of transition states for these reactions. It was not possible to calculate reasonable intrinsic carbon isotope effects with the other metal ions by use of the partitioning ratio of oxalacetate because of decarboxylation by another mechanism.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

6.
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   

7.
Orotidine 5'-monophosphate decarboxylase (ODCase) has been overexpressed in yeast 15C cells transformed with a plasmid carrying the URA3 gene that encodes ODCase. Twenty g of cells having ODCase activity equal to 30 mg of pure enzyme per liter of cell culture were obtained after 9 h of galactose induction. To remove yeast proteases, a 60-90% ammonium sulfate fractionation step plus the addition of EDTA as an inhibitor of metallopeptidases was necessary. The purification protocol yielded ODCase that was protease-free and stable to storage at 4 degrees C for 16 months. The pure enzyme had a specific activity of 40 units/mg in 50 mM phosphate buffer, pH 6, and could be stored at -20 degrees C in 20% glycerol with retention of full activity for more than 2 years. The enzyme had a Km for orotidine 5'-monophosphate of 0.7 microM at pH 6 and 25 degrees C. The molecular weight of the plasmid-derived ODCase monomer determined by electrophoresis on denaturing polyacrylamide gels was 29,500. ODCase sedimented through sucrose density gradients as a monomer of about 30 kDa at low protein concentration and in the absence of ligands that bind at the catalytic site. An increase in the sedimentation rate could be induced by increasing the ODCase concentration or by adding ligands that are competitive inhibitors. ODCase sedimented in a single band typical of a protein of 46 kDa at the highest protein concentration studied or in the presence of 50 mM phosphate or 933 microM substrate (orotidine 5'-monophosphate) or product (UMP). A dimer sedimenting as a protein of about 64 kDa occurred in the presence of 50 microM 6-azauridine 5'-monophosphate or 2 microM 1-(5'-phospho-beta-D-ribofuranosyl) barbituric acid, competitive inhibitors of ODCase. These results resemble the ligand-induced subunit association of the ODCase domain of bifunctional UMP synthase and support the use of yeast ODCase as a model for ODCases from other species.  相似文献   

8.
L M Abell  M H O'Leary 《Biochemistry》1988,27(16):5933-5939
The decarboxylation of histidine by the pyruvate-dependent histidine decarboxylase of Lactobacillus 30a shows a carbon isotope effect of k12/k13 = 1.0334 +/- 0.0005 and a nitrogen isotope effect k14/k15 = 0.9799 +/- 0.0006 at pH 4.8, 37 degrees C. The carbon isotope effect is slightly increased by deuteriation of the substrate and slightly decreased in D2O. The observed nitrogen isotope effect indicates that the imine nitrogen in the substrate-Schiff base intermediate complex is ordinarily protonated, and the pH dependence of the carbon isotope effect indicates that both protonated and unprotonated forms of this intermediate are capable of undergoing decarboxylation. As with the pyridoxal 5'-phosphate dependent enzyme, Schiff base formation and decarboxylation are jointly rate-limiting, with the intermediate histidine-pyruvate Schiff base showing a decarboxylation/Schiff base hydrolysis ratio of 0.5-1.0 at pH 4.8. The decarboxylation transition state is more reactant-like for the pyruvate-dependent enzyme than for the pyridoxal 5'-phosphate dependent enzyme. These studies find no particular energetic or catalytic advantage to the use of pyridoxal 5'-phosphate over covalently bound pyruvate in catalysis of the decarboxylation of histidine.  相似文献   

9.
The isotope effect at C-1 on the H2O2-catalysed decarboxylation of pyruvate (used as a model reaction for the enzymic reaction) increases between pH 3 and 10 from 1.0007 +/- 0.0004 to 1.0283 +/- 0.0014 (25 degrees C). This result indicates a change in the rate-determining step from formation of the tetrahedral intermediate to decarboxylation of this intermediate. Practically no isotope fractionation at C-1 (1.0011 +/- 0.0002, pH 6.0, 25 degrees C) is found in the lactate oxidase-catalysed decarboxylation of lactate, which is indicative for the existence of an irreversible O2-dependent step prior to the enzyme-catalysed decarboxylation. In addition, the result provides further evidence that dissociation of pyruvate and H2O2 from the enzyme can be excluded. The isotope effect at C-2 of lactate in the enzymic reaction (1.0048 +/- 0.0004) is attributed to the hydrogen transfer step from lactate to the coenzyme.  相似文献   

10.
Tryptophan 2-monooxygenase (TMO) from Pseudomonas savastanoi catalyzes the oxidative decarboxylation of l-tryptophan during the biosynthesis of indoleacetic acid. Structurally and mechanistically, the enzyme is a member of the family of l-amino acid oxidases. Deuterium and 15N kinetic isotope effects were used to probe the chemical mechanism of l-alanine oxidation by TMO. The primary deuterium kinetic isotope effect was pH independent over the pH range 6.5-10, with an average value of 6.0 +/- 0.5, consistent with this being the intrinsic value. The deuterium isotope effect on the rate constant for flavin reduction by alanine was 6.3 +/- 0.9; no intermediate flavin species were observed during flavin reduction. The kcat/Kala value was 1.0145 +/- 0.0007 at pH 8. NMR analyses gave an equilibrium 15N isotope effect for deprotonation of the alanine amino group of 1.0233 +/- 0.0004, allowing calculation of the 15N isotope effect on the CH bond cleavage step of 0.9917 +/- 0.0006. The results are consistent with TMO oxidation of alanine occurring through a hydride transfer mechanism.  相似文献   

11.
L M Abell  M H O'Leary 《Biochemistry》1988,27(16):5927-5933
The pyridoxal 5'-phosphate dependent histidine decarboxylase from Morganella morganii shows a nitrogen isotope effect k14/k15 = 0.9770 +/- 0.0021, a carbon isotope effect k12/k13 = 1.0308 +/- 0.0006, and a carbon isotope effect for L-[alpha-2H]histidine of 1.0333 +/- 0.0001 at pH 6.3, 37 degrees C. These results indicate that the overall decarboxylation rate is limited jointly by the rate of Schiff base interchange and by the rate of decarboxylation. Although the observed isotope effects are quite different from those for the analogous glutamate decarboxylase from Escherichia coli [Abell, L. M., & O'Leary, M. H. (1988) Biochemistry 27, 3325], the intrinsic isotope effects for the two enzymes are essentially the same. The difference in observed isotope effects occurs because of a roughly twofold difference in the partitioning of the pyridoxal 5'-phosphate-substrate Schiff base between decarboxylation and Schiff base interchange. The observed nitrogen isotope effect requires that the imine nitrogen in this Schiff base is protonated. Comparison of carbon isotope effects for deuteriated and undeuteriated substrates reveals that the deuterium isotope effect on the decarboxylation step is about 1.20; thus, in the transition state for the decarboxylation step, the carbon-carbon bond is about two-thirds broken.  相似文献   

12.
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   

13.
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   

14.
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   

15.
13C kinetic isotope effects have been measured in carbamyl phosphate for the reaction catalyzed by aspartate transcarbamylase. For the holoenzyme, the value was 1.0217 at zero aspartate, but unity at infinite aspartate, with 4.8 mM aspartate eliminating half of the isotope effect. This pattern proves an ordered kinetic mechanism, with carbamyl phosphate adding before aspartate. The same parameters were observed in the presence of ATP or CTP, showing that there is only one form of active enzyme present, regardless of the presence or absence of allosteric modifiers. These data support the Monod model of allosteric behavior in which the equilibrium between fully active and inactive enzyme is perturbed by selective binding interactions of substrates and modifiers, and there are no enzyme forms having partial activity. Isolated catalytic subunits of the enzyme showed similar 13C isotope effects (1.0240 at zero aspartate, 1.0039 at infinite aspartate, 3.8 mM aspartate causing half of the change from one value to the other), but the finite isotope effect at infinite aspartate shows that the kinetic mechanism is now partly random. With the very slow and poorly bound aspartate analog cysteine sulfinate, the 13C isotope effects were 1.039 for both holoenzyme and catalytic subunits and were not decreased significantly by high levels of cysteine sulfinate. The value of 1.039 is probably close to the intrinsic isotope effect on the chemical reaction, while the kinetic mechanism with this substrate is now fully random because the chemistry is so much slower than release of either reactant from the enzyme.  相似文献   

16.
Orotidine 5'-monophosphate decarboxylase (ODCase) catalyzes the decarboxylation of orotidine 5'-monophosphate to uridine 5'-monophosphate during pyrimidine nucleotide biosynthesis. This enzyme is one of the most proficient known, exhibiting a rate enhancement of over 17 orders of magnitude over the uncatalyzed rate. An interesting question is whether the high proficiency of ODCase is associated with a highly optimized sequence of active site residues. This question was addressed by randomizing 24 residue positions in and around the active site of the E. coli ODCase (pyrF) by site-directed mutagenesis. The libraries of mutants were selected for function from a multicopy plasmid or by single-copy replacement at the pyrF locus on the E. coli chromosome. Stringent sequence requirements for function were found for the mutants expressed from the chromosomal pyrF locus. Six positions were not tolerant of substitutions and several others accepted very limited substitutions. In contrast, all positions could be substituted to some extent when the library mutants were expressed from a multicopy plasmid. For the conserved quartet of charged residues Lys44-Asp71-Lys73-Asp76, a cysteine substitution was found to provide function at positions 71 and 76. A lower pK(a) for both cysteine mutants supports a mechanism whereby the thiolate group of cysteine substitutes for the negatively charged aspartate side chain. The partial function mutants such as D71C and D76C exhibit reduced catalytic efficiency relative to wild type but nevertheless provide a rate enhancement of 15 orders of magnitude over the uncatalyzed rate indicating the catalytic proficiency of the enzyme is robust and tolerant of mutation.  相似文献   

17.
15N isotope effects and solvent deuterium isotope effects have been measured for the hydrolytic deamination of cytidine catalyzed by Escherichia coli cytidine deaminase and for the uncatalyzed reaction proceeding spontaneously in neutral solution at elevated temperatures. The primary (15)(V/K) arising from the exocyclic amino group for wild-type cytidine deaminase acting on its natural substrate, cytidine, is 1.0109 (in H(2)O, pH 7.3), 1.0123 (in H(2)O, pH 4.2), and 1.0086 (in D(2)O, pD 7.3). Increasing solvent D(2)O content has no substantial effect on k(cat) but enhances k(cat)/K(m), with a proton inventory showing that the fractionation factors of at least two protons increase markedly during the reaction. Mutant cytidine deaminases with reduced catalytic activity show more pronounced (15)N isotope effects of 1.0124 (Glu91Ala), 1.0134 (His102Ala), and 1.0158 (His102Asn) at pH 7.3 in H(2)O, as expected for processes in which the chemical transformation of the substrate becomes more rate determining. The isotope effect of mutant His102Asn is 1.033 after correcting for protonation of the -NH(2) group, and represents the intrinsic isotope effect on C-N bond cleavage. This result allows an estimation of the forward commitment of the reaction with the wild-type enzyme. The observed (15)N kinetic isotope effect of the pyrimidine N-3, for wild-type cytidine deaminase acting on cytidine, is 0.9879, which is consistent with protonation and rehybidization of N-3 with hydroxide ion attack on the adjacent carbon to create a tetrahedral intermediate. These results show that enzymatic deamination of cytidine proceeds stepwise through a tetrahedral intermediate with ammonia elimination as the major rate-determining step. The primary (15)N isotope effects observed for the uncatalyzed reaction at pH 7 (1.0021) and pH 12.5 (1.0034) were found to be insensitive to changing temperatures between 100 and 185 degrees C. These results show that the uncatalyzed and the enzymatic deaminations of cytidine proceed by similar mechanisms, although the commitment to C-N bond breaking is greater for the spontaneous reaction.  相似文献   

18.
A detailed kinetic analysis of the catalytic trimer of aspartate transcarbamoylase containing the active site substitution H134A was performed to investigate the role of His 134 in the catalytic mechanism. Replacement of histidine by alanine resulted in decreases in the affinities for the two substrates, carbamoyl phosphate and aspartate, and the inhibitor succinate, by factors of 50, 10, and 6, respectively, and yielded a maximum velocity that was 5% that of the wild-type enzyme. However, the pK values determined from the pH dependence of the kinetic parameters, log V and log (V/K) for aspartate, the pK(i) for succinate, and the pK(ia) for carbamoyl phosphate, were similar for both the mutant and the wild-type enzymes, indicating that the protonated form of His 134 does not participate in binding and catalysis between pH 6.2 and 9.2. 13C and 15N isotope effects were studied to determine which steps in the catalytic mechanism were altered by the amino acid substitutions. The 13(V/K) for carbamoyl phosphate exhibited by the catalytic trimer containing alanine at position 134 revealed an isotope effect of 4.1%, probably equal to the intrinsic value and, together with quantitative analysis of the 15N isotope effects, showed that formation of the tetrahedral intermediate is rate-determining for the mutant enzyme. Thus, His 134 plays a role in the chemistry of the reaction in addition to substrate binding. The initial velocity pattern for the reaction catalyzed by the H134A mutant intersected to the left of the vertical axis, negating an equilibrium ordered kinetic mechanism.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

19.
K Shostak  M E Jones 《Biochemistry》1992,31(48):12155-12161
Pyrimidine nucleotides were tested as substrates for pure yeast orotidylate decarboxylase in an attempt to gain insight into the nature of the catalytic mechanism of the enzyme. Substitutions of the 5-position in the pyrimidine ring of the orotidylate substrate resulted in compounds that are either excellent inhibitors or substrates of the enzyme. The 5-bromo- and 5-chloroorotidylates are potent inhibitors while the 5-fluoro derivative is a good substrate with a turnover number 30 times that observed with orotidylate. When carbon 5 of the pyrimidine ring is replaced by nitrogen in 5-azaorotidylate, the resulting compound is unstable in solution with a half-life of 25 min at pH 6. However, studies with freshly generated 5-azaorotidylate show that an enzyme-dependent reaction occurs, presumably decarboxylation. This enzyme reaction follows simple Michaelis-Menten kinetics. Because the 5-aza group is not electrophilic, an enzyme mechanism utilizing a nucleophilic addition of the enzyme at the 5-position is ruled out. We also present studies that are not compatible with a mechanism requiring the formation of a Schiff's base prior to decarboxylation. The enzyme is tolerant of modest substitution at the 4-position, for the 4-keto group can be replaced with a thioketone. However, no catalysis is observed when the same substitution is made at the 2-position. Similarities in the substrate specificity of orotate phosphoribosyltransferase and orotidylate decarboxylase led us to compare the amino acid sequences of the two enzymes; significant (20%) sequence homology was observed.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

20.
R Hille 《Biochemistry》1991,30(35):8522-8529
Solvent kinetic isotope effect studies of electron transfer within xanthine oxidase have been performed, using a stopped-flow pH-jump technique to perturb the distribution of reducing equivalents within partially reduced enzyme and follow the kinetics of reequilibration spectrophotometrically. It is found that the rate constant for electron transfer between the flavin and one of the iron-sulfur centers of the enzyme observed when the pH is jumped from 10 to 6 decreases from 173 to 25 s-1 on going from H2O to D2O, giving an observed solvent kinetic isotope effect of 6.9. An effect of comparable magnitude is observed for the pH jump in the opposite direction, the rate constant decreasing from 395 to 56 s-1. The solvent kinetic isotope effect on kobs is found to be directly proportional to the mole fraction of D2O in the reaction mix for the pH jump in each direction, consistent with the effect arising from a single exchangeable proton. Calculations of the microscopic rate constants for electron transfer between the flavin and the iron-sulfur center indicate that the intrinsic solvent kinetic isotope effect for electron transfer from the neutral flavin semiquinone to the iron-sulfur center designated Fe/S I is substantially greater than for electron transfer in the opposite direction and that the observed solvent kinetic isotope effect is a weighted averaged of the intrinsic isotope effects for the forward and reverse microscopic electron-transfer steps.(ABSTRACT TRUNCATED AT 250 WORDS)  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