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 792 毫秒
1.
Transient and steady-state kinetic analysis of the reaction of aromatic L-amino acid decarboxylase (AADC), a pyridoxal 5'-phosphate- (PLP-) dependent enzyme, with its substrate dopa was carried out at various pH. The association of AADC and dopa to form the Michaelis complex and the subsequent transaldimination reaction to form the dopa-PLP Schiff base (external aldimine) were followed with a stopped-flow spectrophotometer. Combined with the steady-state k(cat) value, we could present a minimum mechanism for the reaction of AADC and dopa. In the mechanism, the association of the aldimine-protonated form of the enzyme (EH(+)) and the alpha-amino-group-unprotonated form of the substrate (S) is the main route leading to the Michaelis complex. In addition, the association of EH(+) and the alpha-amino-group-protonated form of the substrate (SH(+)) to form a Michaelis complex EH(+).SH(+) was also found as a minor route. The pK(a) of the alpha-amino group of dopa was expected to be decreased in the Michaelis complex, promoting the conversion of EH(+).SH(+) to EH(+).S, the species that directly undergoes transaldimination to form the external aldimine complex. The association of EH(+) and S had been identified as a minor route for the reaction of aspartate and aspartate aminotransferase (AspAT), which has an unusually low pK(a) value of the aldimine and can use the aldimine-unprotonated form (E) of the enzyme for adsorbing the prevalent species SH(+) [Hayashi and Kagamiyama (1997) Biochemistry 36, 13558-13569]. The present study implies that, in most PLP enzymes that have a high pK(a) value of the aldimine like AADC, S preferentially binds to the enzyme (EH(+)). The minor route of EH(+) + SH(+) in AADC may be related to the flexibility of the protein in the Michaelis complex, and a simulation analysis showed that the presence of this route decreases the k(cat) value while increasing the k(cat)/K(m) value. It also suggested that AADC has evolved to suppress the minor route to the extent necessary to obtain the maximal k(cat) value at neutral pH.  相似文献   

2.
Eliot AC  Kirsch JF 《Biochemistry》2002,41(11):3836-3842
The active sites of the homologous pyridoxal phosphate- (PLP-) dependent enzymes 1-aminocyclopropane-1-carboxylate (ACC) synthase and aspartate aminotransferase (AATase) are almost entirely conserved, yet the pK(a)'s of the two internal aldimines are 9.3 and 7.0, respectively, to complement the substrate pK(a)'s (S-adenosylmethionine pK(a) = 7.8 and aspartate pK(a) = 9.9). This complementation is required for maximum enzymatic activity in the physiological pH range. The most prominent structural difference in the active site is that Ile232 of ACC synthase is replaced by alanine in AATase. The I232A mutation was introduced into ACC synthase with a resulting 1.1 unit decrease (from 9.3 to 8.2) in the aldimine pK(a), thus identifying Ile232 as a major determinant of the high pK(a) of ACC synthase. The mutation also resulted in reduced k(cat) (0.5 vs 11 s(-1)) and k(cat)/K(m) values (5.0 x 10(4) vs 1.2 x 10(6) M(-1) s(-1)). The effect of the mutation is interpreted as the result of shortening of the Tyr233-PLP hydrogen bond. Addition of the Y233F mutation to the I232A ACC synthase to generate the double mutant I232A/Y233F raised the pK(a) from 8.2 to 8.8, because the Y233F mutation eliminates the hydrogen bond between that residue and PLP. The introduction of the retro mutation A224I into AATase raised the aldimine pK(a) of that enzyme from 6.96 to 7.16 and resulted in a decrease in single-turnover k(max) (108 vs 900 s(-1) for aspartate) and k(max)/K(m)(app) (7.5 x 10(4) vs 3.8 x 10(5) M(-1) s(-1)) values. The distance from the pyridine nitrogen of the cofactor to a conserved aspartate residue is 2.6 A in AATase and 3.8 A in ACC synthase. The D230E mutation introduced into ACC synthase to close this distance increases the aldimine pK(a) from 9.3 to 10.0, as would be predicted from a shortened hydrogen bond.  相似文献   

3.
Royo M  Fitzpatrick PF 《Biochemistry》2005,44(18):7079-7084
In mammalian cells, the flavoprotein polyamine oxidase catalyzes a key step in the catabolism of polyamines, the oxidation of N1-acetylspermine and N1-acetylspermidine to spermidine and putrescine, respectively. The mechanism of the mouse enzyme has been studied with N1,N12-bisethylspermine (BESPM) as a substrate. At pH 10, the pH optimum, the limiting rate of reduction of the flavin in the absence of oxygen is comparable to the k(cat) value for turnover, establishing reduction as rate-limiting. Oxidation of the reduced enzyme is a simple second-order reaction. No intermediates are seen in the reductive or oxidative half-reactions. The k(cat) value decreases below a pK(a) of 9.0. The k(cat)/K(m) value for BESPM exhibits a bell-shaped pH profile, with pK(a) values of 9.8 and 10.8. These pK(a) values are assigned to the substrate nitrogens. The rate constant for the reaction of the reduced enzyme with oxygen is not affected by a pH between 7.5 and 10. Active site residue Tyr430 is conserved in the homologous protein monoamine oxidase. Mutation of this residue to phenylalanine results in a 6-fold decrease in the k(cat) value and the k(cat)/K(m) value for oxygen due to a comparable decrease in the rate constant for flavin reduction. This moderate change is not consistent with this residue forming a tyrosyl radical during catalysis.  相似文献   

4.
Ralph EC  Fitzpatrick PF 《Biochemistry》2005,44(8):3074-3081
N-Methyltryptophan oxidase (MTOX), a flavoenzyme from Escherichia coli, catalyzes the oxidative demethylation of secondary amino acids such as N-methyltryptophan or N-methylglycine (sarcosine). MTOX is one of several flavin-dependent amine oxidases whose chemical mechanism is still debated. The kinetic properties of MTOX with the slow substrate sarcosine were determined. Initial rate data are well-described by the equation for a ping-pong kinetic mechanism, in that the V/K(O)()2 value is independent of the sarcosine concentration at all accessible concentrations of oxygen. The k(cat)/K(sarc) pH profile is bell-shaped, with pK(a) values of 8.8 and about 10; the latter value matches the pK(a) value of the substrate nitrogen. The k(cat) pH profile exhibits a single pK(a) value of 9.1 for a group that must be unprotonated for catalysis. There is no significant solvent isotope effect on the k(cat)/K(sarc) value. With N-methyl-(2)H(3)-glycine as the substrate, there is a pH-independent kinetic isotope effect on k(cat), k(cat)/K(sarc), and the rate constant for flavin reduction, with an average value of 7.2. Stopped-flow spectroscopy with both the protiated and deuterated substrate failed to detect any intermediates between the enzyme-substrate complex and the fully reduced enzyme. These results are used to evaluate proposed chemical mechanisms.  相似文献   

5.
Spies MA  Toney MD 《Biochemistry》2003,42(17):5099-5107
Alanine racemase catalyzes the pyridoxal phosphate-dependent interconversion of the D- and L-isomers of alanine. Previous studies have shown that the enzyme employs a two-base mechanism in which Lys39 and Tyr265 are the acid/base catalysts. It is thus possible that stereoisomerization of the external aldimine intermediates occurs through a concerted double proton transfer without the existence of a distinct carbanionic intermediate. This possibility was tested by the application of multiple kinetic isotope effect (KIE) methodology to alanine racemase. The mutual dependence of primary substrate and solvent deuterium KIEs has been measured using equilibrium perturbation-type experiments. The conceptually straightforward measurement of the substrate KIE in H(2)O is complemented with a less intuitive protium washout perturbation-type measurement in D(2)O. The primary substrate KIE in the D --> L direction at 25 degrees C is reduced from 1.297 in H(2)O to 1.176 in D(2)O, while in the L --> D direction it is reduced from 1.877 in H(2)O to 1.824 in D(2)O. Similar reductions are also observed at 65 degrees C, the temperature to which the Bacillus stearothermophilus enzyme is adapted. These data strongly support a stepwise racemization of stereoisomeric aldimine intermediates in which a substrate-based carbanion is an obligatory intermediate. The ionizations observed in k(cat)/K(M) pH profiles have been definitively assigned based on the DeltaH(ion) values of the observed pK(a)'s with alanine and on the pH dependence of k(cat)/K(M) for the alternative substrate serine. The acidic pK(a) in the bell-shaped curve is due to the phenolic hydroxyl of Tyr265, which must be unprotonated for reaction with either isomer of alanine. The basic pK(a) is due to the substrate amino group, which must be protonated to react with Tyr265-unprotonated enzyme. A detailed reaction mechanism incorporating these results is proposed.  相似文献   

6.
7.
Sivaraman S  Kirsch JF 《The FEBS journal》2006,273(9):1920-1929
Human tyrosine aminotransferase (hTATase) is the pyridoxal phosphate-dependent enzyme that catalyzes the reversible transamination of tyrosine to p-hydrophenylpyruvate, an important step in tyrosine metabolism. hTATase deficiency is implicated in the rare metabolic disorder, tyrosinemia type II. This enzyme is a member of the poorly characterized Igamma subfamily of the family I aminotransferases. The full length and truncated forms of recombinant hTATase were expressed in Escherichia coli, and purified to homogeneity. The pH-dependent titration of wild-type reveals a spectrum characteristic of family I aminotransferases with an aldimine pK(a) of 7.22. I249A mutant hTATase exhibits an unusual spectrum with a similar aldimine pK(a) (6.85). hTATase has very narrow substrate specificity with the highest enzymatic activity for the Tyr/alpha-ketoglutarate substrate pair, which gives a steady state k(cat) value of 83 s(-1). In contrast there is no detectable transamination of aspartate or other cosubstrates. The present findings show that hTATase is the only known aminotransferase that discriminates significantly between Tyr and Phe: the k(cat)/K(m) value for Tyr is about four orders of magnitude greater than that for Phe. A comparison of substrate specificities of representative Ialpha and Igamma aminotransferases is described along with the physiological significance of the discrimination between Tyr and Phe by hTATase as applied to the understanding of the molecular basis of phenylketonuria.  相似文献   

8.
The steady-state kinetic parameters for epimerization of UDP-galactose by UDP-galactose 4-epimerase from Escherichia coli (GalE), Y149F-GalE, and S124A-GalE have been measured as a function of pH. The deuterium kinetic isotope effects for epimerization of UDP-galactose-C-d(7) by these enzymes have also been measured. The results show that the activity of wild-type GalE is pH-independent in the pH range of 5.5-9.3, and there is no significant deuterium kinetic isotope effect in the reaction of UDP-galactose-C-d(7). It is concluded that the rate-limiting step for epimerization by wild-type GalE is not hydride transfer and must be either a diffusional process or a conformational change. Epimerization of UDP-galactose-C-d(7) by Y149F-GalE proceeds with a pH-dependent deuterium kinetic isotope effect on k(cat) of 2.2 +/- 0.4 at pH 6.2 and 1.1 +/- 0.5 at pH 8.3. Moreover, the plot of log k(cat)/K(m) breaks downward on the acid side with a fitted value of 7.1 for the pK(a). It is concluded that the break in the pH-rate profile arises from a change in the rate-limiting step from hydride transfer at low pH to a conformational change at high pH. Epimerization of UDP-galactose-C-d(7) by S124A-GalE proceeds with a pH-independent deuterium kinetic isotope effect on k(cat) of 2.0 +/- 0.2 between pH 6 and 9. Both plots of log k(cat) and log k(cat)/K(m) display pH dependence. The plot of log k(cat) versus pH breaks downward with a pK(a) of 6.35 +/- 0.10. The plot of log k(cat)/K(m) versus pH is bell-shaped, with fitted pK(a) values of 6.76 +/- 0.09 and 9.32 +/- 0.21. It is concluded that hydride transfer is rate-limiting, and the pK(a) of 6.7 for free S124A-GalE is assigned to Tyr 149, which displays the same value of pK(a) when measured spectrophotometrically in this variant. Acid-base catalysis by Y149F-GalE is attributed to Ser 124, which is postulated to rescue catalysis of proton transfer in the absence of Tyr 149. The kinetic pK(a) of 7.1 for free Y149F-GalE is lower than that expected for Ser 124, as proven by the pH-dependent kinetic isotope effect. Epimerization by the doubly mutated Y149F/S124A-GalE proceeds at a k(cat) that is lower by a factor of 10(7) than that of wild-type GalE. This low rate is attributed to the synergistic actions of Tyr 149 and Ser 124 in wild-type GalE and to the absence of any internal catalysis of hydride transfer in the doubly mutated enzyme.  相似文献   

9.
The pyridoxal 5'-phosphate-dependent beta-subunit of the tryptophan synthase alpha(2)beta(2) complex catalyzes the condensation of L-serine with indole to form L-tryptophan. The first stage of the reaction is a beta-elimination that involves a very fast interconversion of the internal aldimine in a highly fluorescent L-serine external aldimine that decays, via the alpha-carbon proton removal and beta-hydroxyl group release, to the alpha-aminoacrylate Schiff base. This reaction is influenced by protons, monovalent cations, and alpha-subunit ligands that modulate the distribution between open and closed conformations. In order to identify the ionizable residues that might assist catalysis, we have investigated the pH dependence of the rate of the external aldimine decay by rapid scanning UV-visible absorption and single wavelength fluorescence stopped flow. In the pH range 6-9, the reaction was found to be biphasic with the first phase (rate constants k(1)) accounting for more than 70% of the signal change. In the absence of monovalent cations or in the presence of sodium and potassium ions, the pH dependence of k(1) exhibits a bell shaped profile characterized by a pK(a1) of about 6 and a pK(a2) of about 9, whereas in the presence of cesium ions, the pH dependence exhibits a saturation profile characterized by a single pK(a) of 9. The presence of the allosteric effector indole acetylglycine increases the rate of reaction without altering the pH profile and pK(a) values. By combining structural information for the internal aldimine, the external aldimine, and the alpha-aminoacrylate with kinetic data on the wild type enzyme and beta-active site mutants, we have tentatively assigned pK(a1) to betaAsp-305 and pK(a2) to betaLys-87. The loss of pK(a1) in the presence of cesium ions might be due to a shift to lower values, caused by the selective stabilization of a closed form of the beta-subunit.  相似文献   

10.
The six mutations, referred to as the Hex mutations, that together have been shown to convert Escherichia coli aspartate aminotransferase (AATase) specificity to be substantially like that of E. coli tyrosine aminotransferase (TATase) are dissected into two groups, (T109S/N297S) and (V39L/K41Y/T47I/N69L). The letters on the left and right of the numbers designate AATase and TATase residues, respectively. The T109S/N297S pair has been investigated previously. The latter group, the "Grease" set, is now placed in the AATase framework, and the retroGrease set (L39V/Y41K/I47T/L69N) is substituted into TATase. The Grease mutations in the AATase framework were found primarily to lower K(M)s for both aromatic and dicarboxylic substrates. In contrast, retroGrease TATase exhibits lowered k(cat)s for both substrates. The six retroHex mutations, combining retroGrease and S109T/S297N, were found to invert the substrate specificity of TATase, creating an enzyme with a nearly ninefold preference (k(cat)/K(M)) for aspartate over phenylalanine. The retroHex mutations perturb the electrostatic environment of the pyridoxal phosphate cofactor, as evidenced by a spectrophotometric titration of the internal aldimine, which uniquely shows two pK(a)s, 6.1 and 9.1. RetroHex was also found to have impaired dimer stability, with a K(D) for dimer dissociation of 350 nM compared with the wild type K(D) of 4 nM. Context dependence and additivity analyses demonstrate the importance of interactions of the Grease residues with the surrounding protein framework in both the AATase and TATase contexts, and with residues 109 and 297 in particular. Context dependence and cooperativity are particularly evident in the effects of mutations on k(cat)/K(M)(Asp). Effects on k(cat)/K(M)(Phe) are more nearly additive and context independent.  相似文献   

11.
Aspartate aminotransferase has been known to undergo a significant conformational change, in which the small domain approaches the large domain, and the residues at the entrance of the active site pack together, on binding of substrates. Accompanying this conformational change is a two-unit increase in the pK(a) of the pyridoxal 5'-phosphate-Lys(258) aldimine, which has been proposed to enhance catalysis. To elucidate how the conformational change is coupled to the shift in the aldimine pK(a) and how these changes are involved in catalysis, we analyzed structurally and kinetically an enzyme in which Val(39) located at both the domain interface and the entrance of the active site was replaced with a bulkier residue, Phe. The V39F mutant enzyme showed a more open conformation, and the aldimine pK(a) was lowered by 0.7 unit compared with the wild-type enzyme. When Asn(194) had been replaced by Ala in advance, the V39F mutation did not decrease the aldimine pK(a), showing that the domain rotation controls the aldimine pK(a) via the Arg(386)-Asn(194)-pyridoxal 5'-phosphate linkage system. The maleate-bound V39F enzyme showed the aldimine pK(a) 0.9 unit lower than that of the maleate-bound wild-type enzyme. However, the positions of maleate, Asn(194), and Arg(386) were superimposable between the mutant and the wild-type enzymes; therefore, the domain rotation was not the cause of the lowered aldimine pK(a) value. The maleate-bound V39F enzyme showed an altered side-chain packing pattern in the 37-39 region, and the lack of repulsion between Gly(38) carbonyl O and Tyr(225) Oeta seemed to be the cause of the reduced pK(a) value. Kinetic analysis suggested that the repulsion increases the free energy level of the Michaelis complex and promotes the catalytic reaction.  相似文献   

12.
Badarau A  Page MI 《Biochemistry》2006,45(35):10654-10666
The kinetics and mechanism of hydrolysis of the native zinc and metal substituted Bacillus cereus (BcII) metallo-beta-lactamase have been investigated. The pH and metal ion dependence of k(cat) and k(cat)/K(m), determined under steady-state conditions, for the cobalt substituted BcII catalyzed hydrolysis of cefoxitin, cephaloridine, and cephalexin indicate that an enzyme residue of apparent pK(a) 6.3 +/- 0.1 is required in its deprotonated form for metal ion binding and catalysis. The k(cat)/K(m) for cefoxitin and cephalexin with cadmium substituted BcII is dependent on two ionizing groups on the enzyme: one of pK(a1) = 8.7 +/- 0.1 required in its deprotonated form and the other of pK(a2) = 9.3 +/- 0.1 required in its protonated form for activity. The pH dependence of the competitive inhibition constant, K(i), for CdBcII with l-captopril indicates that pK(a1) = 8.7 +/- 0.1 corresponds to the cadmium-bound water. For the manganese substituted BcII, the pH dependence of k(cat)/K(m) for benzylpenicillin, cephalexin, and cefoxitin similarly indicated the importance of two catalytic groups: one of pK(a1) = 8.5 +/- 0.1 which needs to be deprotonated and the other of pK(a2) = 9.4 +/- 0.1 which needs to be protonated for catalysis; the pK(a1) was assigned to the manganese-bound water. The rate was metal ion concentration dependent at the highest manganese concentrations used (10(-)(3) M). The metal substituted species have similar or higher catalytic activities compared with the zinc enzyme, albeit at pHs above 7. Interestingly, with cefoxitin, a very poor substrate for ZnBcII, both k(cat) and k(cat)/K(m) increase with increasing pK(a) of the metal-bound water, in the order Zn < Co < Mn < Cd. A higher pK(a) for the metal-bound water for cadmium and manganese BCII leads to more reactive enzymes than the native zinc BcII, suggesting that the role of the metal ion is predominantly to provide the nucleophilic hydroxide, rather than to act as a Lewis acid to polarize the carbonyl group and stabilize the oxyanion tetrahedral intermediate.  相似文献   

13.
The nature of the proton donor to the C-3 of the enolate of pyruvate, the intermediate in the reaction catalyzed by yeast pyruvate kinase, was investigated by site-directed mutagenesis and physical and kinetic analyses. Thr-298 is correctly located to function as the proton donor. T298S and T298A were constructed and purified. Both mutants are catalytically active with a decrease in k(cat) and k(cat)/K(m)(,PEP). Mn(2+)-activated T298S and T298A do not exhibit homotropic kinetic cooperativity with phosphoenolpyruvate (PEP) in the absence of fructose 1,6-bisphosphate, although PEP binding to enzyme-Mn(2+) is cooperative. The pH dependence of k(cat) for T298A indicates the loss of pK(a)(,2) = 6.4-6.9. Thr-298 affects the ionization (pK(a) approximately 6.5) responsible for modulation of k(cat). Fluorescence studies show altered dissociation constants of ligands to each enzyme complex upon Thr-298 mutations. The rates of the phosphoryl transfer and proton transfer steps in the pyruvate kinase-catalyzed reaction are altered; pyruvate enolization is affected to a greater extent. Proton inventory studies demonstrate solvent isotope effects on k(cat) and k(cat)/K(m)(,PEP). Fractionation factors are metal-dependent and significantly <1. The data suggest that a water molecule in a water channel is the direct proton donor to enolpyruvate and that Thr-298 affects a late step in catalysis.  相似文献   

14.
In previous kinetic studies of Escherichia coli aspartate aminotransferase, it was determined that some substitutions of conserved cysteine 191, which is located outside of the active site, altered the kinetic parameters of the enzyme (Gloss,L.M., Spencer,D. E. and Kirsch,J.F., 1996, Protein Struct. Funct. Genet., 24, 195-208). The mutations resulted in an alkaline shift of 0.6-0.8 pH units for the pK(a) of the internal aldimine between the PLP cofactor and Lys258. The change in the pK(a) affected the pH dependence of the k(cat)/K(m) (aspartate) values for the mutant enzymes. To help to understand these observations, crystal structures of five mutant forms of E.coli aspartate aminotransferase (the maleate complexes of C191S, C191F, C191Y and C191W, and C191S without maleate) were determined at about 2 A resolution in the presence of the pyridoxal phosphate cofactor. The overall three-dimensional fold of each mutant enzyme is the same as that of the wild-type protein, but there is a rotation of the mutated side chain around its C(alpha)-C(beta) bond. This side chain rotation results in a change in the pattern of hydrogen bonding connecting the mutant residue and the protonated Schiff base of the cofactor, which could account for the altered pK(a) of the Schiff base imine nitrogen that was reported previously. These results demonstrate how residues outside the active site can be important in helping determine the subtleties of the active site amino acid geometries and interactions and how mutations outside the active site can have effects on catalysis. In addition, these results help explain the surprising result previously reported that, for some mutant proteins, replacement of a buried cysteine with an aromatic side chain did not destabilize the protein fold. Instead, rotation around the C(alpha)-C(beta) bond allowed each large aromatic side chain to become buried in a nearby pocket without large changes in the enzyme's backbone geometry.  相似文献   

15.
Park C  Raines RT 《Biochemistry》2003,42(12):3509-3518
The value of k(cat)/K(M) for catalysis of RNA cleavage by ribonuclease (RNase) A can exceed 10(9) M(-1) s(-1) in a solution of low salt concentration. This value approaches that expected for the diffusional encounter of the enzyme and its substrate. To reveal the physicochemical constraints upon catalysis by RNase A, the effects of salt concentration, pH, solvent isotope, and solvent viscosity on catalysis were determined with synthetic substrates that bind to all of the enzymic subsites and thereby enable a meaningful analysis. The pK(a) values determined from pH-k(cat)/K(M) profiles at 0.010, 0.20, and 1.0 M NaCl are inconsistent with the known macroscopic pK(a) values of RNase A. This incongruity indicates that catalysis of RNA cleavage by RNase A is limited by the rate of substrate association, even at 1.0 M NaCl. The effect of solvent isotope and solvent viscosity on catalysis support this conclusion. The data are consistent with a mechanism in which RNase A associates with RNA in an intermediate complex, which is stabilized by Coulombic interactions, prior to the formation of a Michaelis complex. Thus, RNase A has evolved to become an enzyme limited by physics rather than chemistry, a requisite attribute of a perfect catalyst.  相似文献   

16.
Catalytic mechanism of hamster arylamine N-acetyltransferase 2   总被引:1,自引:0,他引:1  
Wang H  Liu L  Hanna PE  Wagner CR 《Biochemistry》2005,44(33):11295-11306
Arylamine N-acetyltransferases (NATs) catalyze an acetyl group transfer from AcCoA to primary arylamines, hydrazines, and hydrazides and play a very important role in the metabolism and bioactivation of drugs, carcinogens, and other xenobiotics. The reaction follows a ping-pong bi-bi mechanism. Structure analysis of bacterial NATs revealed a Cys-His-Asp catalytic triad that is strictly conserved in all known NATs. Previously, we have demonstrated by kinetic and isotope effect studies that acetylation of the hamster NAT2 is dependent on a thiolate-imidazolium ion pair (Cys-S(-)-His-ImH(+)) and not a general acid-base catalysis. In addition, we established that, after formation of the acetylated enzyme intermediate, the active-site imidazole, His-107, is likely deprotonated at physiological pH. In this paper, we report steady-state kinetic studies of NAT2 with two acetyl donors, acetyl coenzyme A (AcCoA) and p-nitrophenyl acetate (PNPA), and four arylamine substrates. The pH dependence of k(cat)/K(AcCoA) exhibited two inflection points at 5.32 +/- 0.13 and 8.48 +/- 0.24, respectively. The pK(a) at 5.32 is virtually identical with the previously reported pK(a) of 5.2 for enzyme acetylation, reaffirming that the first half of the reaction is catalyzed by a thiolate-imidazolium ion pair in the active site. The inflection point at 8.48 indicates that a pH-sensitive group on NAT2 is involved in AcCoA binding. A Br?nsted plot constructed by the correlation of log k(4) and log k(H)2(O) with the pK(a) for each arylamine substrate and water displays a linear free-energy relationship in the pK(a) range from -1.7 (H(2)O) to 4.67 (PABA), with a slope of beta(nuc) = 0.80 +/- 0.1. However, a further increase of the pK(a) from 4.67 (PABA) to 5.32 (anisidine) resulted in a 2.5-fold decrease in the k(4) value. Analysis of the pH-k(cat)/K(PABA) profile revealed a pK(a) of 5.52 +/- 0.14 and a solvent kinetic isotope effect (SKIE) of 2.01 +/- 0.04 on k(cat)/K(PABA). Normal solvent isotope effects of 4.8 +/- 0.1, 3.1 +/- 0.1, and 3.2 +/- 0.1 on the k(cat)/K(b) for anisidine, pABglu, and PNA, respectively, were also determined. These observations are consistent with a deacetylation mechanism dominated by nucleophilic attack of the thiol ester for arylamines with pK(a) values or=5.5. The general base is likely His-107 because the His-107 to Gln and Asn mutants were found to be devoid of catalytic activity. In contrast, an increase in pH-dependent hydrolysis of the acetylated enzyme was not observed over a pH range of 5.2-7.5. On the basis of these observations, a catalytic mechanism for the acetylation of arylamines by NAT2 is proposed.  相似文献   

17.
JM Robbins  HR Ellis 《Biochemistry》2012,51(32):6378-6387
The alkanesulfonate monooxygenase enzyme (SsuD) catalyzes the oxygenolytic cleavage of a carbon-sulfur bond from sulfonated substrates. A mechanism involving acid-base catalysis has been proposed for the desulfonation mechanism by SsuD. In the proposed mechanism, base catalysis is involved in abstracting a proton from the alkane peroxyflavin intermediate, while acid catalysis is needed for the protonation of the FMNO(-) intermediate. The pH profiles of k(cat) indicate that catalysis by SsuD requires a group with a pK(a) of 6.6 ± 0.2 to be deprotonated and a second group with a pK(a) of 9.5 ± 0.1 to be protonated. The upper pK(a) value was not present in the pH profiles of k(cat)/K(m). Several conserved amino acid residues (His228, His11, His333, Cys54, and Arg226) have been identified as having potential catalytic importance due to the similar spatial arrangements with close structural and functional relatives of SsuD. Substitutions to these amino acid residues were generated, and the pH dependencies were evaluated and compared to wild-type SsuD. Although a histidine residue was previously proposed to be the active site base, the His variants possessed similar steady-state kinetic parameters as wild-type SsuD. Interestingly, R226A and R226K SsuD variants possessed undetectable activity, and there was no detectable formation of the C4a-(hydro)peroxyflavin intermediate for the Arg226 SsuD variants. Guanidinium rescue with the R226A SsuD variant resulted in the recovery of 1.5% of the wild-type SsuD k(cat) value. These results implicate Arg226 playing a critical role in catalysis and provide essential insights into the mechanistic steps that guide the SsuD desulfonation process.  相似文献   

18.
Aitken SM  Kirsch JF 《Biochemistry》2003,42(2):571-578
Cystathionine beta-synthase (CBS) is a pyridoxal-phosphate-dependent enzyme that catalyzes a beta-replacement reaction in which the hydroxyl group of serine (L-Ser) is displaced by the thiol of homocysteine (L-Hcys) to form cystathionine (L-Cth) in the first step of the trans-sulfuration pathway. A new continuous assay for the forward reaction, employing cystathionine beta-lyase and L-lactate dehydrogenase as coupling enzymes, is described. It alleviates product inhibition by L-Cth and revealed that the values for (1.2 mM) and for substrate inhibition by L-Hcys ( = 2.0 mM) are lower than those previously reported. A continuous, 5,5'-dithio-bis-(2-nitrobenzoic acid) (DTNB)-based assay for the CBS-catalyzed hydrolysis of L-Cth to L-Ser and L-Hcys provides a tool for investigation of the reverse reaction (k(catR) = 0.56 s(-)(1), = 0.083 mM). The (catR)/ versus pH profile of ytCBS is bell-shaped with a pH optimum of 8.3, and the pK(a) values for the acidic and basic limbs are 8.05 and 8.63, respectively. The latter is assigned to the alpha-amino group of L-Cth (pK(a) = 8.54). The internal aldimine of ytCBS remains protonated at pH < 11; therefore, the acidic pK(a) is assigned to an enzyme functionality that is not associated with the internal aldimine. K(eq) was determined directly and from the kinetic parameters, and the values are 0.61 and 1.2 microM, respectively.  相似文献   

19.
The first committed step of lipid A biosynthesis in Gram-negative bacteria is catalyzed by the zinc-dependent hydrolase LpxC that removes an acetate from the nitrogen at the 2' '-position of UDP-3-O-acyl-N-acetylglucosamine. Recent structural characterization by both NMR and X-ray crystallography provides many important details about the active site environment of LpxC from Aquifex aeolicus, a heat-stable orthologue that displays 32% sequence identity to LpxC from Escherichia coli. The detailed reaction mechanism and specific roles of active site residues for LpxC from A. aeolicus are further analyzed here. The pH dependencies of k(cat)/K(M) and k(cat) for the deacetylation of the substrate UDP-3-O-[(R)-3-hydroxymyristoyl]-GlcNAc are both bell-shaped. The ascending acidic limb (pK(1)) was fitted to 6.1 +/- 0.2 for k(cat) and 5.7 +/- 0.2 for k(cat)/K(M). The descending basic limb (pK(2)) was fitted to 8.0 +/- 0.2 for k(cat) and 8.4 +/- 0.2 for k(cat)/K(M). The pH dependence of the E73A mutant exhibits loss of the acidic limb, and the mutant retains only 0.15% activity versus the wild type. The pH dependencies of the other active site mutants H253A, K227A, H253A/K227A, and D234N remain bell-shaped, although their significantly lower activities (0.25%, 0.05%, 0.007%, and 0.57%, respectively) suggest that they contribute significantly to catalysis. Our cumulative data support a mechanism for LpxC wherein Glu73 serves as the general base for deprotonation and activation of the zinc-bound water.  相似文献   

20.
PhnP is a phosphodiesterase that plays an important role within the bacterial carbon-phosphorus lyase (CP-lyase) pathway by recycling a "dead-end" intermediate, 5-phospho-α-d-ribosyl 1,2-cyclic phosphate, that is formed during organophosphonate catabolism. As a member of the metallo-β-lactamase superfamily, PhnP is most homologous in sequence and structure to tRNase Z phosphodiesterases. X-ray structural analysis of PhnP complexed with orthovanadate to 1.5 ? resolution revealed this inhibitor bound in a tetrahedral geometry by the two catalytic manganese ions and the putative general acid residue H200. Guided by this structure, we probed the contributions of first- and second-sphere active site residues to catalysis and metal ion binding by site-directed mutagenesis, kinetic analysis, and ICP-MS. Alteration of H200 to alanine resulted in a 6-33-fold decrease in k(cat)/K(M) with substituted methyl phenylphosphate diesters with leaving group pK(a) values ranging from 4 to 8.4. With bis(p-nitrophenyl)phosphate as a substrate, there was a 10-fold decrease in k(cat)/K(M), primarily the result of a large increase in K(M). Moreover, the nickel ion-activated H200A PhnP displayed a bell-shaped pH dependence for k(cat)/K(M) with pK(a) values (pK(a1) = 6.3; pK(a2) = 7.8) that were comparable to those of the wild-type enzyme (pK(a1) = 6.5; pK(a2) = 7.8). Such modest effects are counter to what is expected for a general acid catalyst and suggest an alternate role for H200 in this enzyme. A Br?nsted analysis of the PhnP reaction with a series of substituted phenyl methyl phosphate esters yielded a linear correlation, a β(lg) of -1.06 ± 0.1, and a Leffler α value of 0.61, consistent with a synchronous transition state for phosphoryl transfer. On the basis of these data, we propose a mechanism for PhnP.  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