首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 15 毫秒
1.
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   

2.
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   

3.
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   

4.
A method has been developed for the positional 13C isotope analysis of pyruvate and acetate by stepwise quantitative degradation. On its base, the kinetic isotope effects on the pyruvate dehydrogenase reaction (enzymes from Escherichia coli and Saccharomyces cerevisiae) for both of the carbon atoms involved in the bond scission (double isotope effect determination) and on C-3 of pyruvate have been determined. The experimental k12/k13 values with the enzyme from E. coli on C-1 and C-2 of pyruvate are 1.0093 +/- 0.0007 and 1.0213 +/- 0.0017, respectively, and, with the enzyme from S. cerevisiae, the values are 1.0238 +/- 0.0013 and 1.0254 +/- 0.0016, respectively. A secondary isotope effect of 1.0031 +/- 0.0009 on C-3 (CH3-group) was found with both enzymes. The size of the isotope on C-1 indicates that decarboxylation is more rate-determining with the yeast enzyme than with the enzyme from E. coli, although it is not the entirely rate-limiting step in the overall reaction sequence. Assuming appropriate values for the intrinsic isotope effect on the decarboxylation step (k3) and the equilibrium isotope effect on the reversible substrate binding (k1, k2), one can calculate values for the partitioning factor R (k3/k2: E. coli enzyme 4.67, S. cerevisiae enzyme 1.14) and the intrinsic isotope effects related to the carbonyl-C (k1/k'1 = 1.019; k3/k'3 = 1.033). The isotope fractionation at C-2 of pyruvate gives strong evidence that the well known relative carbon-13 depletion in lipids from biological material is mainly caused by the isotope effect on the pyruvate dehydrogenase reaction. In addition, our results indicate an alternating 13C abundance in fatty acids, that has already been verified in some cases.  相似文献   

5.
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   

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.
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   

8.
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   

9.
Rishavy MA  Cleland WW 《Biochemistry》2000,39(25):7546-7551
Heavy atom isotope effects at C-2, C-3, and the amino nitrogen of aspartate were determined for the reaction of porcine heart cytosolic aspartate aminotransferase and the tyrosine-225 to phenylalanine mutant of Escherichia coli aspartate aminotransferase. The effects of deuteration at C-2 of aspartate and of D(2)O on the observed heavy atom isotope effects were determined. The multiple isotope effects support the contribution of C(alpha)-H cleavage, ketimine hydrolysis, and oxaloacetate dissociation to the rate limitation with the wild-type enzyme. The existence of a quinonoid intermediate could not be determined due to the kinetic complexity of the enzyme. For the tyrosine-225 to phenylalanine mutant, we are able to conclude that ketimine hydrolysis is the major rate-determining step.  相似文献   

10.
Isotope effects are one of the most powerful kinetic tools for determining enzyme mechanisms. There are three methods of measurement. First, one can compare reciprocal plots with labeled and unlabeled substrates. The ratio of the slopes is the isotope effect on V/K, and the ratio of the vertical intercepts is the isotope effect on V(max). This is the only way to determine V(max) isotope effects, but is limited to isotope effects of 5% or greater. The second method is internal competition, where the labeled and unlabeled substrates are present at the same time and the change in their ratio in residual substrate or in product is used to calculate an isotope effect, which is that on V/K of the labeled reactant. This is the method used for tritium or (14)C, or with the natural abundances of (13)C, (15)N, or (18)O. The third method involves perturbations from equilibrium when a labeled substrate and corresponding unlabeled product are present at chemical equilibrium. This also gives just an isotope effect on V/K for the labeled reactant. The chemistry is typically not fully rate limiting, so that the isotope effect on V/K is given by: (x)(V/K)=((x)k+c(f)+c(r)(x)K(eq))/(1+c(f)+c(r)) where x defines the isotope (D, T, 13, 15, 18 for deuterium, tritium, (13)C, (15)N, or (18)O), and (x)(V/K), (x)k, and (x)K(eq) are the observed isotope effect, the intrinsic one on the chemical step, and the isotope effect on the equilibrium constant, respectively. The constants c(f) and c(r) are commitments in forward and reverse directions, and are the ratio of the rate constant for the chemical reaction and the net rate constant for release from the enzyme of the varied substrate (direct comparison) or labeled substrate (internal competition and equilibrium perturbation) for c(f), or the first product released or the one involved in the perturbation for c(r). The intrinsic isotope effect, (x)k, can be estimated by comparing deuterium and tritium isotope effects on V/K, or by comparing the deuterium isotope effect with (13)C ones with deuterated and undeuterated substrates. Adding a secondary deuterium isotope effect and its effect on the (13)C one can give an exact solution for all intrinsic isotope effects and commitments. The effect of deuteration on a (13)C isotope effect allows one to tell if the two isotope effects are on the same or different steps. Applications of these methods to several enzyme systems will be presented.  相似文献   

11.
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   

12.
Interaction between a 70-amino acid and zinc-binding polypeptide from the regulatory chain and the catalytic (C) trimer of aspartate transcarbamoylase (ATCase) leads to dramatic changes in enzyme activity and affinity for active site ligands. The hypothesis that the complex between a C trimer and 3 polypeptide fragments (zinc domain) is an analog of R state ATCase has been examined by steady-state kinetics, heavy-atom isotope effects, and isotope trapping experiments. Inhibition by the bisubstrate ligand, N-(phosphonacetyl)-L-aspartate (PALA), or the substrate analog, succinate, at varying concentrations of substrates, aspartate, or carbamoyl phosphate indicated a compulsory ordered kinetic mechanism with carbamoyl phosphate binding prior to aspartate. In contrast, inhibition studies on C trimer were consistent with a preferred order mechanism. Similarly, 13C kinetic isotope effects in carbamoyl phosphate at infinite aspartate indicated a partially random kinetic mechanism for C trimer, whereas results for the complex of C trimer and zinc domain were consistent with a compulsory ordered mechanism of substrate binding. The dependence of isotope effect on aspartate concentration observed for the Zn domain-C trimer complex was similar to that obtained earlier for intact ATCase. Isotope trapping experiments showed that the compulsory ordered mechanism for the complex was attributable to increased "stickiness" of carbamoyl phosphate to the Zn domain-C trimer complex as compared to C trimer alone. The rate of dissociation of carbamoyl phosphate from the Zn domain-C trimer complex was about 10(-2) that from C trimer.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

13.
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   

14.
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   

15.
Formulations of an enzyme mechanism where only a single step is presumed to be isotopically sensitive can be written in terms of forward and reverse commitments to catalysis. These commitments provide a natural and intuitive way of interpreting the observed isotope effects. Unfortunately, when multiple isotopically sensitive steps are present in the mechanism, including effects associated with pre-equilibria of the unbound substrate, the observed V/K kinetic isotope effect is expressed as a complicated expression of the intrinsic rate constants for each step, the interpretation of which is not always immediately obvious. We show here that V/K isotope effects from unbranched or rapid-equilibrium random Michaelis-Menten systems containing multiple isotopically sensitive steps can be written as a weighted average of the intrinsic isotope effects on each step, where this intrinsic isotope effect from each step is the product of the equilibrium isotope effect on the formation of the reacting intermediate for that step and the intrinsic kinetic effect on the forward rate constant for that step, and the weighting factors are simply the reciprocal sum of the forward and reverse commitments for each step i plus unity, 1/(C(fi)+C(ri)+1), equivalent to the sensitivity index [Ray, W.J., 1983.  相似文献   

16.
The biosynthesis of S-adenosylmethionine occurs in a unique enzymatic reaction in which the synthesis of the sulfonium center results from displacement of the entire polyphosphate chain from MgATP. The mechanism of S-adenosylmethionine synthetase (ATP:L-methionine s-adenosyltransferase) from Escherichia coli has been characterized by kinetic isotope effect and substrate trapping measurements. Replacement of 12C by 14C at the 5' carbon of ATP yields a primary Vmax/Km isotope effect (12C/14C) of 1.128 +/- 0.003 in the absence of added monovalent cation activator (K+). At saturating K+ concentrations (10 mM) the primary isotope effect diminishes slightly to 1.108 +/- 0.003, indicating that the step in the mechanism involving bond breaking at the 5' carbon of MgATP has a small commitment to catalysis at conditions near Vmax. No alpha-secondary 3H isotope effect from [5'-3H]ATP was detected, (1H/3H) = 1.000 +/- 0.002, even in the absence of KCl. There was no significant primary sulfur isotope effect from [35S]methionine at KCl concentrations from 0 to 10 mM. Substitution of the methyl group of methionine with tritium yielded a beta-secondary isotope effect (CH3/C3H3) = 1.009 +/- 0.008 independent of KCl concentration. The reaction of selenomethionine and [5'-14C]ATP gave a primary isotope effect of 1.097 +/- 0.006, independent of KCl concentration. Substrate trapping experiments demonstrated that the step in the mechanism involving bond making to sulfur of methionine does not have a significant commitment to catalysis at 0.25 mM KCl, therefore intrinsic isotope effects were observed. Substrate trapping experiments indicated that the step involving bond breaking at carbon 5' of MgATP has a 10% commitment to catalysis at 0.25 mM KCl. The isotope effects are interpreted in terms of an Sn2-like transition state structure in which bonding of the C5' is symmetric with respect to the departing tripolyphosphate group and the incoming sulfur of methionine. With selenomethionine as substrate an earlier transition state is implicated.  相似文献   

17.
Primary amines functionally replace lysine 258 by catalyzing both the 1,3-prototropic shift and external aldimine hydrolysis reactions with the inactive aspartate aminotransferase mutant K258A. This finding allows classical Brønsted analyses of proton transfer reactions to be applied to enzyme-catalyzed reactions. An earlier study of the reaction of K258A with cysteine sulfinate (Toney, M.D. & Kirsch, J.F., 1989, Science 243, 1485) provided a beta value of 0.4 for the 1,3-prototropic shift. The beta value reported here for the transamination of oxalacetate to aspartate is 0.6. The catalytic efficacy of primary amines is largely determined by basicity and molecular volume. The dependence of the rate constants for the reactions of K258A and K258M on amine molecular volume is nearly identical. This observation argues that the alkyl groups of the added amines do not occupy the position of the lysine 258 side chain in the wild type enzyme. Large primary C alpha and insignificant solvent deuterium kinetic isotope effects with amino acid substrates demonstrate that the amine nitrogen of the exogenous catalysts directly abstracts the labile proton in the rate-determining step.  相似文献   

18.
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   

19.
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   

20.
K Sugiyama  W F Trager 《Biochemistry》1986,25(23):7336-7343
A kinetic model is presented from which steady-state equations are derived that describe the intramolecular competition for the enzymatically mediated hydroxylation of two like groupings of a prochiral substrate. The observed isotope effect in such a system if one of the groupings is isotopically labeled is shown to be a function of three parameters: the equilibrium constant for the catalytically sensitive orientations of the two prochiral groupings at the active site, the intrinsic isotope effect associated with the bond-breaking step, and the relative rates of bond breaking vs. enzyme-substrate dissociation. The expected isotope effects associated with the omega-hydroxylation of racemic, (R)-, and (S)-2-phenylpropane-1,1,1-d3 and the product stereoselectivity associated with the omega-hydroxylation of (R)- and (S)-[1-13C]-2-phenylpropane were determined with microsomal preparations (cytochrome P-450) from untreated and phenobarbital- and beta-naphthoflavone-pretreated male Sprague-Dawley rats. The data from these experiments allow the observed isotope effect to be evaluated in terms of its component parts, i.e., expected isotope effects, product stereoselectivity, and equilibrium constant. These data further suggest that the intramolecular isotope effect is consistent with a hydrogen abstraction recombination mechanism and is largely dependent upon the chemical nature of the porphyrin-Fe-oxene complex but independent of specific apoprotein structure, product stereoselectivity is primarily dependent upon apoprotein structure, and product stereoselectivity is a good measure of the equilibrium constant and both parameters are dependent upon the chirality of the active site.  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