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.
Decarboxylation of oxalacetate by pyruvate carboxylase   总被引:1,自引:0,他引:1  
P V Attwood  W W Cleland 《Biochemistry》1986,25(25):8191-8196
The decarboxylation of oxalacetate by pyruvate carboxylase in the absence of ADP and Pi is stimulated 400-fold by the presence of oxamate, which is an inhibitory analogue of pyruvate. The observation of substrate inhibition when either oxamate or oxalacetate is varied at a fixed concentration of the other indicates that both molecules bind at the same site on the enzyme. The pH profiles for this reaction show no evidence of the involvement of an enzymic acid-base catalyst, suggesting that the proton and CO2 units may be exchanged directly between the reactants (although CO2 sequestered in the active site may be an intermediate in the process). The pH profiles of the full reverse reaction of pyruvate carboxylase in which oxalacetate decarboxylation is coupled to ATP formation and where Pi is the variable substrate do, however, indicate that such an acid-base catalyst is involved in the other partial reaction of the enzyme in proton transfer to and from biotin. The enzyme also displays two oxamate-independent oxalacetate decarboxylating activities, one of which is biotin-dependent and the other is independent of biotin.  相似文献   

2.
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   

3.
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   

4.
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   

5.
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   

6.
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   

7.
Deuterium isotope effects and 13C isotope effects with deuterium- and protium-labeled malate have been obtained for both NAD- and NADP-malic enzymes by using a variety of alternative dinucleotide substrates. With nicotinamide-containing dinucleotides as the oxidizing substrate, the 13C effect decreases when deuterated malate is the substrate compared to the value obtained with protium-labeled malate. These data are consistent with a stepwise chemical mechanism in which hydride transfer precedes decarboxylation of the oxalacetate intermediate as previously proposed [Hermes, J. D., Roeske, C. A., O'Leary, M. H., & Cleland, W. W. (1982) Biochemistry 21, 5106]. When dinucleotide substrates such as thio-NAD, 3-acetylpyridine adenine dinucleotide, and 3-pyridinealdehyde adenine dinucleotide that contain modified nicotinamide rings are used, the 13C effect increases when deuterated malate is the substrate compared to the value obtained with protium-labeled malate. These data, at face value, are consistent with a change in mechanism from stepwise to concerted for the oxidative decarboxylation portion of the mechanism. However, the increase in the deuterium isotope effect from 1.5 to 3 with a concomitant decrease in the 13C isotope effect from 1.034 to 1.003 as the dinucleotide substrate is changed suggests that the reaction may still be stepwise with the non-nicotinamide dinucleotides. A more likely explanation is that a beta-secondary 13C isotope effect accompanies hydride transfer as a result of hyperconjugation of the beta-carboxyl of malate as the transition state for the hydride transfer step is approached.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

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

9.
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   

10.
Macrophomate synthase from the fungus Macrophoma commelinae IFO 9570 is a Mg(II)-dependent dimeric enzyme that catalyzes an extraordinary, complex five-step chemical transformation from 2-pyrone and oxalacetate to benzoate involving decarboxylation, C-C bond formation, and dehydration. The catalytic mechanism of the whole pathway was investigated in three separate chemical steps. In the first decarboxylation step, the enzyme loses oxalacetate decarboxylation activity upon incubation with EDTA. Activity is fully restored by addition of Mg(II) and is not restored with other divalent metal cations. The dissociation constant of 0.93 x 10(-)(7) for Mg(II) and atomic absorption analysis established a 1:1 stoichiometric complex. Inhibition of pyruvate formation with 2-pyrone revealed that the actual product in the first step is a pyruvate enolate, which undergoes C-C bond formation in the presence of 2-pyrone. Incubation of substrate analogs provided aberrant adducts that were produced via C-C bond formation and rearrangement. This strongly indicates that the second step is two C-C bond formations, affording a bicyclic intermediate. Based on the stereospecificity, involvement of a Diels-Alder reaction at the second step is proposed. Incubation of the stereospecifically deuterium-labeled malate with 2-pyrones in the presence of malate dehydrogenase provided information for the stereochemical course of the reaction catalyzed by macrophomate synthase, indicating that the first decarboxylation provides pyruvate (Z)-[3-(2)H]enolate and that dehydration at the final step occurs with anti-elimination accompanied by concomitant decarboxylation. Examination of kinetic parameters in the individual steps suggests that the third step is the rate-determining step of the overall transformation.  相似文献   

11.
Phosphoenolpyruvate carboxykinase, which has been isolated from chicken liver mitochondria in essentially homogenous form, carries out the irreversible decarboxylation of oxalacetate to pyruvate in the presence of catalytic amounts of GDP or IDP, as well as the reversible decarboxylation of oxalacetate to phosphoenolpyruvate in the presence of substrate amounts of GTP or ITP. The pyruvate- and phosphoenolpyruvate-forming reactions are similar in their nucleoside specificity and appear to be carried out by the same protein. However, the two activities vary markedly in their response to added metal ions and sulfhydryl reagents. Phosphoenolpyruvate formation is completely dependent on the presence of a divalent metal ion, with Mn2+ the most effective species. This reaction is also stimulated by sulfhydryl reagents such as 2-mercaptoethanol. In contrast, the pyruvate-forming reaction is strongly inhibited by divalent metal ions, including Mn2+, and also by moderate concentrations of sulfhydryl reagents. These observations and the demonstration that pyruvate kinase-like activity is very low or absent make it unlikely that pyruvate formation proceeds via phosphoenolpyruvate as an intermediate. Although the pyruvate-forming reaction is inhibited by added metal ions, the reaction is also inhibited by metal-chelating agents such as 8-hydroxyquinoline and o-phenanthroline, suggesting that the reaction is dependent on the presence of a metal ion. It has not been possible, however, to demonstrate that the enzyme is a metalloprotein.  相似文献   

12.
Karsten WE  Hwang CC  Cook PF 《Biochemistry》1999,38(14):4398-4402
The NAD-malic enzyme from Ascaris suum catalyzes the divalent metal ion-dependent oxidative decarboxylation of L-malate to give pyruvate and CO2, with NAD+ as the oxidant. Alpha-secondary tritium kinetic isotope effects were measured with NAD+ or APAD+ and L-malate-2-H(D) and several different divalent metal ions. The alpha-secondary tritium kinetic isotope effects are slightly higher than 1 with NAD+ and L-malate as substrates, much larger than the expected inverse isotope effect for a hybridization change from sp2 to sp3. The alpha-secondary tritium kinetic isotope effects are reduced to values near 1 with L-malate-2-D as the substrate, regardless of the metal ion that is used. Data suggest the presence of quantum mechanical tunneling and coupled motion in the malic enzyme reaction when NAD+ and malate are used as substrates. Isotope effects were also measured using the D/T method with NAD+ and Mn2+ as the substrate pair. A Swain-Schaad exponent of 2.2 (less than the value of 3.26 expected for strictly semiclassical behavior) is estimated, suggesting the presence of other slow steps along the reaction pathway. With APAD+ and Mn2+ as the substrate pair, inverse alpha-secondary tritium kinetic isotope effects are observed, and a Swain-Schaad exponent of 3.3 is estimated, consistent with rate-limiting hydride transfer and no quantum mechanical tunneling or coupled motion. Data are discussed in terms of the malic enzyme mechanism and the theory developed by Huskey for D/T isotope effects as an indicator of tunneling [Huskey, W. P. (1991) J. Phys. Org. Chem. 4, 361-366].  相似文献   

13.
The pyruvate dehydrogenase from Escherichia coli showed a primary kinetic isotope effect when its overall reaction or the partial reaction of the pyruvate dehydrogenase component were tested in deuterium oxide. The Michaelis constants for pyruvate were nearly unchanged, but the maximum velocities in water and deuterium oxide differed, their ratio being DV = 1.7 for the overall reaction and DV = 2.1 for the E1p reaction. The pH profile and, accordingly, the delta pK1 and delta pK2 values were shifted by 0.6 units to higher pL values. A linear proton inventory curve was obtained when varying the atom fractions of protons relative to deuterons from 100 to 0%. This is an indication for a single proton transfer. It is proposed that this relatively weak primary isotope effect may be caused by the protonation of the N1' nitrogen at the pyrimidine ring of the cofactor by an adjacent glutamate residue. The proton of its carboxylic group exchanges very fast with deuterons of the solvent.  相似文献   

14.
Patel MP  Liu WS  West J  Tew D  Meek TD  Thrall SH 《Biochemistry》2005,44(50):16753-16765
Beta-ketoacyl-acyl carrier protein reductase (KACPR) catalyzes the NADPH-dependent reduction of beta-ketoacyl-acyl carrier protein (AcAc-ACP) to generate (3S)-beta-hydroxyacyl-ACP during the chain-elongation reaction of bacterial fatty acid biosynthesis. We report the evaluation of the kinetic and chemical mechanisms of KACPR using acetoacetyl-CoA (AcAc-CoA) as a substrate. Initial velocity, product inhibition, and deuterium kinetic isotope effect studies were consistent with a random bi-bi rapid-equilibrium kinetic mechanism of KACPR with formation of an enzyme-NADP(+)-AcAc-CoA dead-end complex. Plots of log V/K(NADPH) and log V/K(AcAc)(-)(CoA) indicated the presence of a single basic group (pK = 5.0-5.8) and a single acidic group (pK = 8.0-8.8) involved in catalysis, while the plot of log V vs pH indicated that at high pH an unprotonated form of the ternary enzyme complex was able to undergo catalysis. Significant and identical primary deuterium kinetic isotope effects were observed for V (2.6 +/- 0.4), V/K(NADPH) (2.6 +/- 0.1), and V/K(AcAc)(-)(CoA) (2.6 +/- 0.1) at pH 7.6, but all three values attenuated to values of near unity (1.1 +/- 0.03 or 0.91 +/- 0.02) at pH 10. Similarly, the large alpha-secondary deuterium kinetic isotope effect of 1.15 +/- 0.02 observed for [4R-(2)H]NADPH on V/K(AcAc)(-)(CoA) at pH 7.6 was reduced to a value of unity (1.00 +/- 0.04) at high pH. The complete analysis of the pH profiles and the solvent, primary, secondary, and multiple deuterium isotope effects were most consistent with a chemical mechanism of KACPR that is stepwise, wherein the hydride-transfer step is followed by protonation of the enolate intermediate. Estimations of the intrinsic primary and secondary deuterium isotope effects ((D)k = 2.7, (alpha)(-D)k = 1.16) and the correspondingly negligible commitment factors suggest a nearly full expression of the intrinsic isotope effects on (D)V/K and (alpha)(-D)V/K, and are consistent with a late transition state for the hydride transfer step. Conversely, the estimated intrinsic solvent effect ((D)2(O)k) of 5.3 was poorly expressed in the experimentally derived parameters (D)2(O)V/K and (D)2(O)V (both = 1.2 +/- 0.1), in agreement with the estimation that the catalytic commitment factor for proton transfer to the enolate intermediate is large. Such detailed knowledge of the chemical mechanism of KAPCR may now help guide the rational design of, or inform screening assay-design strategies for, potent inhibitors of this and related enzymes of the short chain dehydrogenase enzyme class.  相似文献   

15.
We have determined 15N isotope effects and solvent deuterium isotope effects for adenosine deaminase using both adenosine and the slow alternate substrate 7,8-dihydro-8-oxoadenosine. With adenosine, 15N isotope effects were 1.0040 in H2O and 1.0023 in D2O, and the solvent deuterium isotope effect was 0.77. With 7,8-dihydro-8-oxoadenosine, 15N isotope effects were 1.015 in H2O and 1.0131 in D2O, and the solvent deuterium isotope effect was 0.45. The inverse solvent deuterium isotope effect shows that the fractionation factor of a proton, which is originally less than 0.6, increases to near unity during formation of the tetrahedral intermediate from which ammonia is released. Proton inventories for 1/V and 1/(V/K) vs percent D2O are linear, indicating that a single proton has its fractionation factor altered during the reaction. We conclude that a sulfhydryl group on the enzyme donates its proton to oxygen or nitrogen during this step. pH profiles with 7,8-dihydro-8-oxoadenosine suggest that the pK of this sulfhydryl group is 8.45. The inhibition of adenosine deaminase by cadmium also shows a pK of approximately 9 from the pKi profile. Quantitative analysis of the isotope effects suggests an intrinsic 15N isotope effect for the release of ammonia from the tetrahedral intermediate of approximately 1.03 for both substrates; however, the partition ratio of this intermediate for release of ammonia as opposed to back-reaction is 14 times greater for adenosine (1.4) than for 7,8-dihydro-8-oxoadenosine (0.1).(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

16.
The mechanism of the galactosyltransferase-catalyzed reaction was probed using positional isotope exchange, alpha-secondary deuterium isotope effects, and inhibition studies with potential transition state analogs. Incubation of [beta-18O2, alpha beta-18O]UDP-galactose and alpha-lactalbumin with galactosyltransferase from bovine milk did not result in any positional isotope exchange. The addition of 4-deoxy-4-fluoroglucose as a dead-end inhibitor did not induce any detectable positional isotope exchange. alpha-Secondary deuterium isotope effects of 1.21 +/- 0.04 on Vmax and 1.05 +/- 0.04 on Vmax/KM were observed for [1-2H]-UDP-galactose. D-Glucono-1,5-lactone, D-galactono-1,4-lactone, D-galactono-1,5-lactone, nojirimycin, and deoxynojirimycin, did not inhibit the galactosyl transfer reaction at concentrations less than 1.0 mM. The magnitude of the secondary deuterium isotope effect supports a mechanism in which the anomeric carbon of the galactosyl moiety has substantial sp2 character in the transition state. Therefore, the cleavage of the bond between the galactose and UDP moieties in the transition state has proceeded to a much greater extent than the formation of the bond between the galactose and the incoming glucose. The lack of a positional isotope exchange reaction indicates that the beta-phosphoryl group of the UDP is not free to rotate in the absence of an acceptor substrate.  相似文献   

17.
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   

18.
The yeast Pyc1 isoform of pyruvate carboxylase has been further characterized and shown to differ from the Pyc2 isoform in its K(a) for K(+) activation. Pyc1 differs from chicken liver pyruvate carboxylase in the lack of effect of acetyl-CoA on ADP phosphorylation by carbamoyl phosphate, which may be a result of differences in the loci of action of the effector between the two enzymes. Solvent D(2)O isotope effects have been measured with Pyc1 on the full pyruvate carboxylation reaction, the ATPase reaction in the absence of pyruvate, and the carbamoyl phosphate-ADP phosphorylation reaction for the first time for pyruvate carboxylase. Proton inventories indicate that the measured isotope effects are due to a single proton transfer step in the reaction. The inverse isotope effects observed in all reactions suggest that the proton transfer step converts the enzyme from an inactive to an active form. Kinetic measurements on the C249A mutant enzyme suggest that C249 is involved in the binding and action of enzyme activators K(+) and acetyl-CoA. C249 is not involved in ATP binding as was observed for the corresponding residue in the biotin carboxylase subunit of Escherichia coli acetyl-CoA carboxylase, nor is it directly responsible for the measured inverse (D)(k(cat)/K(m)) isotope effects. The size of the inverse isotope effects indicates that they may result from formation of a low-barrier hydrogen bond. Modification of the wild type and C249A mutant with o-phthalaldehyde suggests that C249 is involved in isoindole formation but that the modification of this residue is not directly responsible for the accompanying major loss of enzyme activity.  相似文献   

19.
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   

20.
Site-directed mutagenesis was used to change Lys 240 of yeast pyruvate kinase (Lys 269 in muscle PK) to Met. K240M has an absolute requirement for FBP for catalysis. K240M is 100- and 1000-fold less active than wild-type YPK in the presence of Mn(2+) and Mg(2+), respectively. Steady-state fluorescence titration data suggest that the substrate PEP binds to K240M with the same affinity as it does to wild-type YPK. The rate of phosphoryl transfer in K240M has been decreased >1000-fold compared to wild-type YPK. The detritiation of 3-[(3)H]pyruvate catalyzed by YPK occurs at a rate significantly greater than the spontaneous rate. Detritiation of pyruvate by wild-type YPK occurs as a divalent metal- and FBP-dependent process requiring ATP. There is no detectable detritiation of pyruvate catalyzed by K240M. The solvent deuterium isotope effect on k(cat) is 2.7 +/- 0.2 and 1.6 +/- 0.1 for the wild type and for K240M YPK, respectively. This suggests that the isotope sensitive step in the PK reaction does not involve Lys 240 and that the enolpyruvate intermediate is still protonated by K240M. Isotope trapping was used to characterize enolpyruvate protonation by K240M. While there was enrichment of the methyl protons of pyruvate from labeled solvent formed by catalysis with muscle PK and wild-type YPK, only background levels of tritium were trapped with K240M. In K240M, the proton donor exchanges protons with the solvent at a higher rate relative to turnover than does the proton donor in wild-type YPK. The pH-rate profile of K240M exhibits the loss of a pK(a) value of 8. 8 observed with wild-type YPK. The above data and recent crystal structure data suggest that Lys 240 interacts with the phosphoryl group of phosphoenolpyruvate and helps to stabilize the pentavalent phosphate transition state during phosphoryl transfer. Phosphoryl transfer is highly coupled to proton transfer, or Lys 240 also affects enolate protonation.  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