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.
Yeast (Saccharomyces cerevisiae) is unusual in being the only organism thus far identified as having two genes for pyruvate carboxylase. The expression of the two isozymes Pyc1 and Pyc2 appears to be differentially regulated, and since both are expressed cytoplasmically, this suggests that they have different properties. To the present, little has been done to characterize these isozymes, and almost all of the published kinetic information on yeast pyruvate carboxylase comes from measurements of enzyme prepared from bakers' yeast which is likely to be a mixture of both isozymes. Here we have measured basic kinetic parameters for Pyc1 and found that the K(a) of this isozyme for acetyl CoA is in the order of 8-10-fold higher than previously recorded, suggesting that Pyc1 and Pyc2 may be differentially regulated by this effector. Pyc1 is highly dependent on the presence of acetyl CoA for activity and in this respect is similar to chicken liver pyruvate carboxylase. However, unlike the chicken liver enzyme, the quaternary structure of the enzyme is quite stable in the absence of acetyl CoA, and the major locus of action of this effector appears to lie outside of the stimulation of the biotin carboxylation reaction.  相似文献   

2.
Levert KL  Lloyd RB  Waldrop GL 《Biochemistry》2000,39(14):4122-4128
Biotin carboxylase from Escherichia coli catalyzes the ATP-dependent carboxylation of biotin and is one component of the multienzyme complex acetyl-CoA carboxylase, which catalyzes the committed step in long-chain fatty acid synthesis. For the carboxylation of biotin to occur, biotin must be deprotonated at its N1' position. Kinetic investigations, including solvent isotope effects and enzyme inactivation by N-ethylmaleimide, suggested a catalytic role for a cysteine residue and led to the proposal of a mechanism for the deprotonation of biotin. The proposed pathway suggests a catalytic base removes a proton from a nearby cysteine residue, forming a thiolate anion, which then abstracts the proton from biotin. Inactivation studies of pyruvate carboxylase, which has an analogous mode of action to biotin carboxylase, suggests the catalytic base in this reaction is a lysine residue. Using the crystal structure of biotin carboxylase, cysteine 230 and lysine 238 were identified as the likely active-site residues that act as this acid-base pair. To test the hypothesis that cysteine 230 and lysine 238 act as an acid-base pair to deprotonate biotin, site-directed mutagenesis was used to mutate cysteine 230 to alanine (C230A) and lysine 238 to glutamine (K238Q). Mutations at either residue resulted in a 50-fold increase in the K(m) for ATP. The C230A mutation had no effect on the formation of carboxybiotin, indicating that cysteine 230 does not play a role in the deprotonation of biotin. However, the K238Q mutation resulted in no formation of carboxybiotin, which showed that lysine 238 has a role in the carboxylation reaction. N-Ethylmaleimide was found to inactivate the C230A mutant but not the K238Q mutant, suggesting that N-ethylmaleimide is reacting with lysine 238 and not cysteine 230. The pH dependence of N-ethylmaleimide inactivation revealed that the pK value for lysine 238 was 9.4 or higher, suggesting lysine 238 is not a catalytic base. Thus, the results suggest that cysteine 230 and lysine 238 do not act as an acid-base pair in the deprotonation of biotin.  相似文献   

3.
When chicken liver pyruvate carboxylase was incubated with either H14CO3- or gamma-[32P]ATP, a labeled carboxyphospho-enzyme intermediate could be isolated. The complex was catalytically competent, as determined by its subsequent ability to transfer either 14CO2 to pyruvate or 32P to ADP. While the carboxyphospho-enzyme complex was inherently unstable and the stoichiometry of the transfer was variable depending on experimental conditions, both the [14C]carboxyphospho-enzyme and the carboxy[32P]phospho-enzyme had similar half-lives. Acetyl-CoA was shown to be involved in the conversion of the carboxyphospho-enzyme complex to the more stable carboxybiotin-enzyme species, which was consistent with the effects of acetyl-CoA on isotope exchange reactions involving ATP. We were unable to detect the formation of a phosphorylated biotin derivative during the ATP cleavage reaction. In the presence of K+ and at pH 9.5, the acetyl-CoA-independent activity of chicken liver pyruvate carboxylase approached 2% of the acetyl-CoA-stimulated rate, which represents a 30-fold increase on previously reported activity for this enzyme.  相似文献   

4.
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   

5.
1. The reaction pathway for the carboxylation of pyruvate, catalysed by pig liver pyruvate carboxylase, was studied in the presence of saturating concentrations of K(+) and acetyl-CoA. 2. Free Mg(2+) binds to the enzyme in an equilibrium fashion and remains bound during all further catalytic cycles. MgATP(2-) binds next, followed by HCO(3) (-) and then pyruvate. Oxaloacetate is released before the random release, at equilibrium, of P(i) and MgADP(-). 3. This reaction pathway is compared with the double displacement (Ping Pong) mechanisms that have previously been described for pyruvate carboxylases from other sources. The reaction pathway proposed for the pig liver enzyme is superior in that it shows no kinetic inconsistencies and satisfactorily explains the low rate of the ATP[unk][(32)P]P(i) equilibrium exchange reaction. 4. Values are presented for the stability constants of the magnesium complexes of ATP, ADP, acetyl-CoA, P(i), pyruvate and oxaloacetate.  相似文献   

6.
7.
The activity of the biotin-dependent enzyme pyruvate carboxylase from many organisms is highly regulated by the allosteric activator acetyl-CoA. A number of X-ray crystallographic structures of the native pyruvate carboxylase tetramer are now available for the enzyme from Rhizobium etli and Staphylococcus aureus. Although all of these structures show that intersubunit catalysis occurs, in the case of the R. etli enzyme, only two of the four subunits have the allosteric activator bound to them and are optimally configured for catalysis of the overall reaction. However, it is apparent that acetyl-CoA binding does not induce the observed asymmetrical tetramer conformation and it is likely that, under normal reaction conditions, all of the subunits have acetyl-CoA bound to them. Thus the activation of the enzyme by acetyl-CoA involves more subtle structural effects, one of which may be to facilitate the correct positioning of Arg353 and biotin in the biotin carboxylase domain active site, thereby promoting biotin carboxylation and, at the same time, preventing abortive decarboxylation of carboxybiotin. It is also apparent from the crystal structures that there are allosteric interactions induced by acetyl-CoA binding in the pair of subunits not optimally configured for catalysis of the overall reaction.  相似文献   

8.
While crystallographic structures of the R. etli pyruvate carboxylase (PC) holoenzyme revealed the location and probable positioning of the essential activator, Mg(2+), and nonessential activator, acetyl-CoA, an understanding of how they affect catalysis remains unclear. The current steady-state kinetic investigation indicates that both acetyl-CoA and Mg(2+) assist in coupling the MgATP-dependent carboxylation of biotin in the biotin carboxylase (BC) domain with pyruvate carboxylation in the carboxyl transferase (CT) domain. Initial velocity plots of free Mg(2+) vs pyruvate were nonlinear at low concentrations of Mg(2+) and a nearly complete loss of coupling between the BC and CT domain reactions was observed in the absence of acetyl-CoA. Increasing concentrations of free Mg(2+) also resulted in a decrease in the K(a) for acetyl-CoA. Acetyl phosphate was determined to be a suitable phosphoryl donor for the catalytic phosphorylation of MgADP, while phosphonoacetate inhibited both the phosphorylation of MgADP by carbamoyl phosphate (K(i) = 0.026 mM) and pyruvate carboxylation (K(i) = 2.5 mM). In conjunction with crystal structures of T882A R. etli PC mutant cocrystallized with phosphonoacetate and MgADP, computational docking studies suggest that phosphonoacetate could coordinate to one of two Mg(2+) metal centers in the BC domain active site. Based on the pH profiles, inhibition studies, and initial velocity patterns, possible mechanisms for the activation, regulation, and coordination of catalysis between the two spatially distinct active sites in pyruvate carboxylase from R. etli by acetyl-CoA and Mg(2+) are described.  相似文献   

9.
An investigation was made of the interaction of pyruvate carboxylase with its allosteric effector, acetyl-CoA, and the velocity profile of the deacylation of acetyl-CoA as a function of acetyl-CoA concentration indicated that this ligand does not bind to this enzyme in a positive homotropic co-operative manner. An examination was therefore made of the factors that contribute to the sigmoidicity of the rate curves obtained for pyruvate carboxylation with various concentrations of acetyl-CoA. Hill coefficients for acetyl-CoA obtained with both sheep and chicken liver pyruvate carboxylases were found to be dependent on the fixed pyruvate concentration used in the assay solution. Thus, by varying the acetyl-CoA concentration, the degree of saturation of the enzyme by pyruvate was also changed. A further consequence of non-saturating concentrations of pyruvate was that the non-productive hydrolysis of the enzyme- carboxybiotin complex increased, resulting in an under-estimate of the reaction velocity measured by oxaloacetate formation. Another factor contributing to the sigmoidicity is that, at non-saturating concentrations of acetyl-CoA, the enzyme undergoes inactivation upon dilution to low protein concentrations, again resulting in an under-estimate of the reaction velocity. Under conditions where none of the above factors was operating and the only effect of varying acetyl-CoA concentrations was to alter the proportion of the enzyme catalysing the carboxylation reaction at acetyl-CoA-dependent and -independent rates, the sigmoidicity of the acetyl-CoA velocity profile was completely eliminated.  相似文献   

10.
Earlier, it had been proposed in the laboratories at Halle that a cysteine residue is responsible for the hysteretic substrate activation behavior of yeast pyruvate decarboxylase. More recently, this idea has received support in a series of studies from Rutgers with the identification of residue C221 as the site where substrate is bound to transmit the information to H92, to E91, to W412, and finally to the active center thiamin diphosphate. According to steady-state kinetic assays, the C221A/C222A variant is no longer subject to substrate activation yet is still a well-functioning enzyme. Several further experiments are reported on this variant: (1) The variant exhibits lag phases in the product formation progress curves, which can be attributed to a unimolecular step in the pre-steady-state stage of catalysis. (2) The rate of exchange with solvent deuterium of the thiamin diphosphate C2H atom is slowed by a factor of 2 compared to the wild-type enzyme, suggesting that the reduced activity that results from the substitutions some 20 A from the active center is also seen in the first key step of the reaction. (3) The solvent (deuterium oxide) kinetic isotope effect was found to be inverse on V(max)/K(m) (0.62), and small but normal on V(max) (1.26), virtually ruling out residue C221 as being responsible for the inverse effects reported for the wild-type enzyme at low substrate concentrations. The solvent kinetic isotope effects are compared to those on two related enzymes not subject to substrate activation, Zymomonas mobilis pyruvate decarboxylase and benzoylformate decarboxylase.  相似文献   

11.
1. Pyruvate carboxylase from baker's yeast is inhibited by ADP, AMP and adenosine at pH8.0 in the presence of magnesium chloride concentrations equal to or higher than the ATP concentration. The adenine moiety is essential for the inhibitory effect. 2. In the absence of acetyl-CoA (an allosteric activator) ADP, AMP and adenosine are competitive inhibitors with respect to ATP. In the presence of acetyl-CoA, besides the effect with respect to ATP, AMP competes with acetyl-CoA, whereas ADP and adenosine are non-competitive inhibitors with respect to the activator. 3. Pyruvate carboxylase is inhibited by NADH. The inhibition is competitive with respect to acetyl-CoA and specific with respect to NADH, since NAD(+), NADP(+) and NADPH do not affect the enzyme activity. In the absence of acetyl-CoA, NAD(+), NADH, NADP(+) and NADPH do not inhibit pyruvate carboxylase. 4. Pyruvate carboxylase is inhibited by ADP, AMP and NADH at pH6.5, in the presence of 12mm-Mg(2+), 0.75mm-Mn(2+) and 0.5mm-ATP, medium conditions similar to those existing inside the yeast cell. The ADP and NADH effects are consistent with a regulation of enzyme activity by the intracellular [ATP]/[ADP] ratio and secondarily by NADH concentration. These mechanisms would supplement the already known control of yeast pyruvate carboxylase by acetyl-CoA and l-aspartate. Inhibition by AMP is less marked and its physiological role is perhaps limited.  相似文献   

12.
Although aldolase-catalyzed condensations proceed by stepwise mechanisms via the intermediacy of nucleophilic enol(ate)s or enamines, the mechanisms of those enzymes that catalyze Claisen-type condensations are unclear. The reaction pathway followed by an enzyme from this second group, malate synthase, has been studied by the double-isotope fractionation method to determine whether the reaction is stepwise or concerted. In agreement with earlier work, a deuterium kinetic isotope effect D(V/K) of 1.3 +/- 0.1 has been found when [2H3]acetyl-CoA is the substrate. The 13C isotope effect at the aldehydic carbon of glyoxylate has also been measured. For this determination, the malate product (containing the carbon of interest at C-2) was quantitatively transformed into a new sample of malate having the carbon of interest at C-4. This material was decarboxylated by malic enzyme to produce the appropriate CO2 for isotope ratio mass spectrometric analysis. The 13C isotope effect with [1H3]acetyl-CoA [that is, 13(V/K)H] is 1.0037 +/- 0.0004. By use of the known values of the intermolecular and intramolecular deuterium effects and of 13(V/K)H, the value of the 13C isotope effect when deuteriated [2H3]acetyl-CoA is the substrate [that is, 13(V/K)D] can be predicted for three possible mechanisms. If 13(V/K)H is a kinetic isotope effect and the reaction is concerted, the value of the 13C effect on deuteriation of acetyl-CoA will rise to 1.011; if 13(V/K)H is a kinetic isotope effect and the reaction is stepwise, the value of the 13C effect will fall to 1.0025; and if the 13C effect is an equilibrium isotope effect deriving from glyoxylate dehydration, the reaction is necessarily stepwise, and the value of 13(V/K)D will be 1.0037, unchanged from that of 13(V/K)H. Experimentally, the value of 13(V/K)D is 1.0037 +/- 0.0007, which requires that malate synthase follow a stepwise path. It is therefore clear that the two salient characteristics of enzymes that catalyze Claisen-like condensations, namely, the absence of enzyme-catalyzed proton exchange with solvent and the inversion of the configuration at the nucleophilic center, which had been suggestive of a concerted pathway, are not mechanistically diagnostic.  相似文献   

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

14.
1. Acetyl-CoA acts as a positive allosteric effector in the formation of active pyruvate carboxylase from its apoenzyme, ATP and (+)-biotin which is catalysed by holoenzyme synthetase; this effect is counteracted by l-aspartate. 2. The Hill coefficients (apparent n values) were approximately 2 for acetyl-CoA and 4 for l-aspartate; the n value for each effector remained constant when the concentration of the other effector was varied. 3. Active pyruvate carboxylase was formed also when the apoenzyme was incubated with holoenzyme synthetase and synthetic biotinyl-5'-AMP; acetyl-CoA and l-aspartate affected this process as they did the overall reaction from (+)-biotin and ATP. 4. When hydroxylamine replaced the apoenzyme, holoenzyme synthetase catalysed the formation of biotinylhydroxamate from (+)-biotin and ATP. This reaction was not affected by the allosteric effectors. 5. The apoenzyme was protected against thermal denaturation by acetyl-CoA and, to a lesser degree, by l-aspartate. The holoenzyme synthetase was not markedly protected by these effectors. 6. It is concluded that the allosteric effectors act on the apoenzyme and not the synthetase.  相似文献   

15.
—Previously published equations for analysis of [14C]glucose metabolism assumed that products of glycolysis enter the citric acid cycle only through acetyl-CoA (Larrabee , 1978). These equations are now extended to include entrance into the citric acid cycle through the pyruvate carboxy-lase reaction as well as via acetyl-CoA and are applied to previously reported data from dorsal root ganglia of 15-day-old chicken embryos. The rate of output of labelled CO2 in the presence of [2-14C] glucose could not be accounted for if the flux rate into the citric acid cycle through the pyruvate carboxylase reaction was assumed to be more than about 10–15% of that through acetyl-CoA. It is concluded (1) that the pyruvate carboxylase reaction is a relatively minor source of material for the citric acid cycle in these ganglia and (2) that the previous conclusions about [14C]glucose metabolism, which ignored the pyruvate carboxylase reaction, need not be modified in the light of this reanalysis.  相似文献   

16.
1. Oxaloacetate synthesis catalysed by pyruvate carboxylase from a thermophilic Bacillus in the absence of acetyl-CoA required addition of high concentrations of pyruvate, MgATP(2-) and HCO(3) (-), and at 45 degrees C occurred at a maximum rate approx. 20% of that in the presence of a saturating concentration of acetyl-CoA. The apparent K(m) for HCO(3) (-) at pH7.8 was 400mm without acetyl-CoA, and 16mm with a saturating activator concentration. The relationship between reciprocal initial rate and reciprocal MgATP(2-) concentration was non-linear (convex-down) in the absence of acetyl-CoA, but the extent of deviation decreased as the activator concentration was increased. The relationship between reciprocal initial rate and reciprocal pyruvate concentration was non-linear (convex-down) in the presence or absence of acetyl-CoA. 2. The optimum pH for catalysis of oxaloacetate synthesis was similar in the presence or absence of acetyl-CoA. The variation with pH of apparent K(m) for HCO(3) (-) implicated residue(s) with pK(a) 8.6 in catalysis of the activator-independent oxaloacetate synthesis. 3. Linear Arrhenius and van't Hoff plots were observed for the temperature-dependence of oxaloacetate synthesis in the absence of acetyl-CoA over the range 25-55 degrees C. E(a) (activation energy) was 56.3kJ/mol and DeltaH(double dagger) (HCO(3) (-)) (enthalpy of activation) was -38.6kJ/mol. In the presence of acetyl-CoA, biphasic Arrhenius and van't Hoff plots are observed with a change of slope at 30 degrees C in each case. E(a) was 43.7 and 106.3kJ/mol above and below 30 degrees C respectively. 4. Incubation of Bacillus pyruvate carboxylase with trinitrobenzenesulphonate caused specific inactivation of acetyl-CoA-dependent catalytic activity associated with the incorporation of 1.3+/-0.2 trinitrophenyl residues per subunit. Activator-independent catalysis and regulatory inhibition by l-aspartate were unaffected. The rate of inactivation of acetyl-CoA-dependent catalysis by trinitrobenzenesulphonate was specifically decreased by addition of acetyl-CoA and other acetyl-CoA and other acyl-CoA species, but complete protection was not obtained. 5. All alkylacyl derivatives of CoA tested activated Bacillus pyruvate carboxylase; acetyl-CoA was the most effective. The apparent K(a) exhibited a biphasic relationship with acyl-chain length for the straight-chain homologues. Certain long-chain acyl-CoA species showed additional activation at a high concentration. Weak activation occurred on addition of CoA or adenosine 3',5'-bisphosphate, but carboxyacyl-CoA species and derivatives containing a modified phosphoadenosyl group were inhibitory. Thioesters of CoA with non-carboxylic acids, e.g. methanesulphonyl-CoA, serve as activators of the thermophilic Bacillus and Saccharomyces cerevisiae pyruvate carboxylases, but as inhibitors of pyruvate carboxylases obtained from chicken and rat liver. 6. alpha-Oxoglutarate mimics the effect of l-aspartate as a regulatory inhibitor of the pyruvate carboxylases from both the thermophilic Bacillus and Saccharomyces cerevisiae. l-Glutamate was ineffective in both cases.  相似文献   

17.
The role of biotin-dependent enzymes in the fatty liver and kidney syndrome of young chicks was studied. Under conditions of a marginal deficiency of dietary biotin, the level of biotin in the liver has differing effects on the activities of two biotin-dependent enzymes, pyruvate carboxylase and acetyl-CoA carboxylase. The activity of acetyl-CoA carboxylase is increased, but when the dietary deficiency of biotin produces biotin levels which are below 0-8 mug/g of liver, the activity of pyruvate carboxylase may be insufficient to completely metabolize pyruvate via gluconeogenesis. There is an increase in liver size and in the activities of enzymes involved in alternate pathways for the removal of pyruvate. Blood lactate accumulates and there is increased synthesis of fatty acids, and an accumulation of palmitoleic acid; these steps are accomplished by increased activities of at least the following enzymes: acetyl-CoA carboxylase, malate dehydrogenase (decarboxylating) (NADP+) and the desaturase enzyme. When the biotin level is below 0-35 mug/g of liver and the chick is subjected to a stress, physiological defence mechanisms of the chick may be inadequate to maintain homeostasis and they finally collapse, resulting in accumulation of triacylglycerol in the liver and blood; the chick is unable to maintain blood glucose levels and death occurs, often only a few hours after the imposition of the stress.  相似文献   

18.
19.
Johnson CM  Huang B  Roderick SL  Cook PF 《Biochemistry》2004,43(49):15534-15539
The pH dependence of kinetic parameters was determined in both reaction directions to obtain information about the acid-base chemical mechanism of serine acetyltransferase from Haemophilus influenzae (HiSAT). The maximum rates in both reaction directions, as well as the V/K(serine) and V/K(OAS), decrease at low pH, exhibiting a pK of approximately 7 for a single enzyme residue that must be unprotonated for optimum activity. The pH-independent values of V(1)/E(t), V(1)/K(serine)E(t), V/K(AcCoA)E(t), V(2)/E(t), V(2)/K(OAS)E(t), and V/K(CoA)E(t) are 3300 +/- 180 s(-1), (9.6 +/- 0.4) x 10(5) M(-1) s(-1), 3.3 x 10(6) M(-1) s(-1), 420 +/- 50 s(-1), (2.1 +/- 0.5) x 10(4) M(-1) s(-1), and (4.2 +/- 0.7) x 10(5) M(-1) s(-1), respectively. The K(i) values for the competitive inhibitors glycine and l-cysteine are pH-independent. The solvent deuterium kinetic isotope effects on V and V/K in the direction of serine acetylation are 1.9 +/- 0.2 and 2.5 +/- 0.4, respectively, and the proton inventories are linear for both parameters. Data are consistent with a single proton in flight in the rate-limiting transition state. A general base catalytic mechanism is proposed for the serine acetyltransferase. Once acetyl-CoA and l-serine are bound, an enzymic general base accepts a proton from the l-serine side chain hydroxyl as it undergoes a nucleophilic attack on the carbonyl of acetyl-CoA. The same enzyme residue then functions as a general acid, donating a proton to the sulfur atom of CoASH as the tetrahedral intermediate collapses, generating the products OAS and CoASH. The rate-limiting step in the reaction at limiting l-serine levels is likely formation of the tetrahedral intermediate between serine and acetyl-CoA.  相似文献   

20.
Lin Y  West AH  Cook PF 《Biochemistry》2008,47(40):10809-10815
Potassium is an activator of the reaction catalyzed by homoisocitrate (HIc) dehydrogenase (HIcDH) from Saccharomyces cerevisiae with either the natural substrate, homoisocitrate, or the slow substrate isocitrate. On the basis of initial velocity studies, the selectivity of the activator site for monovalent ions was determined. Potassium is the best activator, and NH 4 (+) and Rb (+) are also activators of the reaction, while Cs (+), Li (+), and Na (+) are not. Chloride inhibits the reaction, while acetate is much less effective. Substitution of potassium acetate for KCl changes the kinetic mechanism of HIcDH from a steady state random to a fully ordered mechanism with the binding of MgHIc followed by K (+) and NAD. The change in mechanism likely reflects an apparent increase in the affinity of enzyme for MgHIc as a result of elimination of the inhibitory effect of Cl (-). The V/K NAD pH-rate profile in the absence of K (+) exhibits a >10-fold decrease in the affinity of enzyme for NAD upon deprotonation of an enzyme side chain with a p K a of about 5.5-6. On the other hand, the affinity for NAD is relatively constant at high pH in the presence of 200 mM KCl. Since the affinity of the dinucleotide decreases as the enzyme group is protonated and the effect is overcome by a monovalent cation, the enzyme residue may be a neutral acid, aspartate or glutamate. Data suggest that K (+) replaces the proton, and likely binds to the enzyme residue, the pyrophosphoryl moiety of NAD, or both. Viscosity and solvent deuterium isotope effects studies suggest the isomerization of E-MgHIc binary complex limits the rate in the absence of K (+).  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