首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 15 毫秒
1.
Nagar M  Narmandakh A  Khalak Y  Bearne SL 《Biochemistry》2011,50(41):8846-8852
Mandelate racemase (EC 5.1.2.2) from Pseudomonas putida catalyzes the interconversion of the enantiomers of mandelic acid and a variety of aryl- and heteroaryl-substituted mandelate derivatives, suggesting that β,γ-unsaturation is a requisite feature of substrates for the enzyme. We show that β,γ-unsaturation is not an absolute requirement for catalysis and that mandelate racemase can bind and catalyze the racemization of (S)-trifluorolactate (k(cat) = 2.5 ± 0.3 s(-1), K(m) = 1.74 ± 0.08 mM) and (R)-trifluorolactate (k(cat) = 2.0 ± 0.2 s(-1), K(m) = 1.2 ± 0.2 mM). The enzyme was shown to catalyze hydrogen-deuterium exchange at the α-postion of trifluorolactate using (1)H NMR spectrocsopy. β-Elimination of fluoride was not detected using (19)F NMR spectroscopy. Although mandelate racemase bound trifluorolactate with an affinity similar to that exhibited for mandelate, the turnover numbers (k(cat)) were markedly reduced by ~318-fold, resulting in catalytic efficiencies (k(cat)/K(m)) that were ~400-fold lower than those observed for mandelate. These observations suggested that chemical steps on the enzyme were likely rate-determining, which was confirmed by demonstrating that the rates of mandelate racemase-catalyzed racemization of (S)-trifluorolactate were not dependent upon the solvent microviscosity. Circular dichroism spectroscopy was used to measure the rates of nonenzymatic racemization of (S)-trifluorolactate at elevated temperatures. The values of ΔH(?) and ΔS(?) for the nonenzymatic racemization reaction were determined to be 28.0 (±0.7) kcal/mol and -15.7 (±1.7) cal K(-1) mol(-1), respectively, corresponding to a free energy of activation equal to 33 (±4) kcal/mol at 25 °C. Hence, mandelate racemase stabilizes the altered trifluorolactate in the transition state (ΔG(tx)) by at least 20 kcal/mol.  相似文献   

2.
Mandelate racemase (MR, EC 5.1.2.2) from Pseudomonas putida catalyzes the Mg(2+)-dependent 1,1-proton transfer that interconverts the enantiomers of mandelate. Crystal structures of MR reveal that the phenyl group of all ground-state ligands is located within a hydrophobic cavity, remote from the site of proton abstraction. MR forms numerous electrostatic and H-bonding interactions with the alpha-OH and carboxyl groups of the substrate, suggesting that these polar groups may remain relatively fixed in position during catalysis while the phenyl group is free to move between two binding sites [i.e., the R-pocket and the S-pocket for binding the phenyl group of (R)-mandelate and (S)-mandelate, respectively]. We show that MR binds benzilate (K(i) = 0.67 +/- 0.12 mM) and (S)-cyclohexylphenylglycolate (K(i) = 0.50 +/- 0.03 mM) as competitive inhibitors with affinities similar to that which the enzyme exhibits for the substrate. Therefore, the active site can simultaneously accommodate two phenyl groups, consistent with the existence of an R-pocket and an S-pocket. Wild-type MR exhibits a slightly higher affinity for (S)-mandelate [i.e., K(m)(S)(-)(man) < K(m)(R)(-)(man)] but catalyzes the turnover of (R)-mandelate slightly more rapidly (i.e., k(cat)(R)(-->)(S) > k(cat)(S)(-->)(R)). Upon introduction of steric bulk into the S-pocket using site-directed mutagenesis (i.e., the F52W, Y54W, and F52W/Y54W mutants), this catalytic preference is reversed. Although the catalytic efficiency (k(cat)/K(m)) of all the mutants was reduced (11-280-fold), all mutants exhibited a higher affinity for (R)-mandelate than for (S)-mandelate, and higher turnover numbers with (S)-mandelate as the substrate, relative to those with (R)-mandelate. (R)- and (S)-2-hydroxybutyrate are expected to be less sensitive to the additional steric bulk in the S-pocket. Unlike those for mandelate, the relative binding affinities for these substrate analogues are not reversed. These results are consistent with steric obstruction in the S-pocket and support the hypothesis that the phenyl group of the substrate may move between an R-pocket and an S-pocket during racemization. These conclusions were also supported by modeling of the binary complexes of the wild-type and F52W/Y54W enzymes with the substrate analogues (R)- and (S)-atrolactate, and of wild-type MR with bound benzilate using molecular dynamics simulations.  相似文献   

3.
St Maurice M  Bearne SL 《Biochemistry》2000,39(44):13324-13335
Mandelate racemase (MR) catalyzes the interconversion of the enantiomers of mandelic acid, stabilizing the altered substrate in the transition state by 26 kcal/mol relative to the substrate in the ground state. To understand the origins of this binding discrimination, carboxylate-, phosphonate-, and hydroxamate-containing substrate and intermediate analogues were examined for their ability to inhibit MR. Comparison of the competitive inhibition constants revealed that an alpha-hydroxyl function is required for recognition of the ligand as an intermediate analogue. Two intermediate analogues, alpha-hydroxybenzylphosphonate (alpha-HBP) and benzohydroxamate, were bound with affinities approximately 100-fold greater than that observed for the substrate. Furthermore, MR bound alpha-HBP enantioselectively, displaying a 35-fold higher affinity for the (S)-enantiomer relative to the (R)-enantiomer. In the X-ray structure of mandelate racemase [Landro, J. A., Gerlt, J. A., Kozarich, J. W., Koo, C. W., Shah, V. J., Kenyon, G. L., Neidhart, D. J., Fujita, J., and Petsko, G. A. (1994) Biochemistry 33, 635-643], the alpha-hydroxyl function of the competitive inhibitor (S)-atrolactate is within hydrogen bonding distance of Asn 197. To demonstrate the importance of the alpha-hydroxyl function in intermediate binding, the N197A mutant was constructed. The values of k(cat) for N197A were reduced 30-fold for (R)-mandelate and 179-fold for (S)-mandelate relative to wild-type MR; the values of k(cat)/K(m) were reduced 208-fold for (R)-mandelate and 556-fold for (S)-mandelate. N197A shows only a 3.5-fold reduction in its affinity for the substrate analogue (R)-atrolactate but a 51- and 18-fold reduction in affinity for alpha-HBP and benzohydroxamate, respectively. Thus, interaction between Asn 197 and the substrate's alpha-hydroxyl function provides approximately 3.5 kcal/mol of transition-state stabilization free energy to differentially stabilize the transition state relative to the ground state.  相似文献   

4.
Bourque JR  Bearne SL 《Biochemistry》2008,47(2):566-578
Mandelate racemase from Pseudomonas putida catalyzes the Mg2+-dependent 1,1-proton transfer that interconverts the enantiomers of mandelate. Residues of the 20s and 50s loops determine, in part, the topology and polarity of the active site and hence the substrate specificity. Previously, we proposed that, during racemization, the phenyl ring of mandelate moves between an S-pocket comprised of residues from the 50s loop and an R-pocket comprised of residues from the 20s loop [Siddiqi, F., Bourque, J. R., Jiang, H., Gardner, M., St. Maurice, M., Blouin, C., and Bearne, S. L. (2005) Biochemistry 44, 9013-9021]. The 20s loop constitutes a mobile beta-meander flap that covers the active site cavity shielding it from solvent and controlling entry and egress of ligands. To understand the role of the 20s loop in catalysis and substrate specificity, we constructed a series of mutants (V22A, V22I, V22F, T24S, A25V, V26A, V26L, V26F, V29A, V29L, V29F, V26A/V29L, and V22I/V29L) in which the sizes of hydrophobic side chains of the loop residues were varied. Catalytic efficiencies (kcat/Km) for all mutants were reduced between 6- and 40-fold with the exception of those of V22I, V26A, V29L, and V22I/V29L which had near wild-type efficiencies with mandelate. Thr 24 and Ala 25, located at the tip of the 20s loop, were particularly sensitive to minor alterations in the size of their hydrophobic side chains; however, most mutations were tolerated quite well, suggesting that flap mobility could compensate for increases in the steric bulk of hydrophobic side chains. With the exception of V29L, with mandelate as the substrate, and V22F and V26A/V29L, with 2-naphthylglycolate (2-NG) as the substrate, the values of kcat and Km were not altered in a manner consistent with steric obstruction of the R-pocket, perhaps due to flap mobility compensating for the increased size of the hydrophobic side chains. Surprisingly, V22I and V29L catalyzed the racemization of the bulkier substrate 2-NG with kcat/Km values approximately 2-fold greater than those observed for wild-type mandelate racemase. Although minor changes in substrate specificity were achieved through alterations of the active site flap of mandelate racemase, our results suggest that hydrophobic residues that reside on a flexible flap and define the topology of an active site through their van der Waals contacts with the substrate are quite tolerant of a variety of steric substitutions.  相似文献   

5.
Mandelate racemase (MR, EC 5.1.2.2) from Pseudomonas putida catalyzes the Mg(2+)-dependent interconversion of the enantiomers of mandelate, stabilizing the altered substrate in the transition state by 26 kcal/mol relative to the substrate in the ground state. To understand the origins of this binding discrimination, we determined the X-ray crystal structures of wild-type MR complexed with two analogues of the putative aci-carboxylate intermediate, benzohydroxamate and Cupferron, to 2.2-? resolution. Benzohydroxamate is shown to be a reasonable mimic of the transition state and/or intermediate because its binding affinity for 21 MR variants correlates well with changes in the free energy of transition state stabilization afforded by these variants. Both benzohydroxamate and Cupferron chelate the active site divalent metal ion and are bound in a conformation with the phenyl ring coplanar with the hydroxamate and diazeniumdiolate moieties, respectively. Structural overlays of MR complexed with benzohydroxamate, Cupferron, and the ground state analogue (S)-atrolactate reveal that the para carbon of the substrate phenyl ring moves by 0.8-1.2 ? between the ground state and intermediate state, consistent with the proposal that the phenyl ring moves during MR catalysis while the polar groups remain relatively fixed. Although the overall protein structure of MR with bound intermediate analogues is very similar to that of MR with bound (S)-atrolactate, the intermediate-Mg(2+) distance becomes shorter, suggesting a tighter complex with the catalytic Mg(2+). In addition, Tyr 54 moves closer to the phenyl ring of the bound intermediate analogues, contributing to an overall constriction of the active site cavity. However, site-directed mutagenesis experiments revealed that the role of Tyr 54 in MR catalysis is relatively minor, suggesting that alterations in enzyme structure that contribute to discrimination between the altered substrate in the transition state and the ground state by this proficient enzyme are extremely subtle.  相似文献   

6.
Mandelate racemase (MR) catalyzes the 1,1-proton transfer that interconverts the enantiomers of mandelate. The transition state/intermediate analogues N-hydroxyformanilide (K(i)=2.79+/-0.19 microM) and cupferron (K(i)=2.67+/-0.09 microM) are identified as potent competitive inhibitors of MR. The pH-pK(i) profile indicates that MR can bind either the protonated or deprotonated forms of N-hydroxyformanilide, with a 10-fold greater affinity for the latter form.  相似文献   

7.
St Maurice M  Bearne SL 《Biochemistry》2002,41(12):4048-4058
Mandelate racemase (EC 5.1.2.2) from Pseudomonas putida catalyzes the interconversion of the two enantiomers of mandelic acid with remarkable proficiency, producing a rate enhancement exceeding 15 orders of magnitude. The rates of the forward and reverse reactions catalyzed by the wild-type enzyme and by a sluggish mutant (N197A) have been studied in the absence and presence of several viscosogenic agents. A partial dependence on relative solvent viscosity was observed for values of kcat and kcat/Km for the wild-type enzyme in sucrose-containing solutions. The value of kcat for the sluggish mutant was unaffected by varying solvent viscosity. However, sucrose did have a slight activating effect on mutant enzyme efficiency. In the presence of the polymeric viscosogens poly(ethylene glycol) and Ficoll, no effect on kcat or kcat/Km for the wild-type enzyme was observed. These results are consistent with both substrate binding and product dissociation being partially rate-determining in both directions. The viscosity variation method was used to estimate the rate constants comprising the steady-state expressions for kcat and kcat/Km. The rate constant for the conversion of bound (R)-mandelate to bound (S)-mandelate (k2) was found to be 889 +/- 40 s(-1) compared with a value of 654 +/- 58 s(-1) for kcat in the same direction. From the temperature dependence of Km (shown to equal K(S)), k2, and the rate constant for the uncatalyzed reaction [Bearne, S. L., and Wolfenden, R. (1997) Biochemistry 36, 1646-1656], we estimated the enthalpic and entropic changes associated with substrate binding (DeltaH = -8.9 +/- 0.8 kcal/mol, TDeltaS = -4.8 +/- 0.8 kcal/mol), the activation barrier for conversion of bound substrate to bound product (DeltaH# = +15.4 +/- 0.4 kcal/mol, TDeltaS# = +2.0 +/- 0.1 kcal/mol), and transition state stabilization (DeltaH(tx) = -22.9 +/- 0.8 kcal/mol, TDeltaS(tx) = +1.8 +/- 0.8 kcal/mol) during mandelate racemase-catalyzed racemization of (R)-mandelate at 25 degrees C. Although the high proficiency of mandelate racemase is achieved principally by enthalpic reduction, there is also a favorable and significant entropic contribution.  相似文献   

8.
Catalytic reaction pathway for the mitogen-activated protein kinase ERK2   总被引:2,自引:0,他引:2  
Prowse CN  Hagopian JC  Cobb MH  Ahn NG  Lew J 《Biochemistry》2000,39(20):6258-6266
The structural, functional, and regulatory properties of the mitogen-activated protein kinases (MAP kinases) have long attracted considerable attention owing to the critical role that these enzymes play in signal transduction. While several MAP kinase X-ray crystal structures currently exist, there is by comparison little mechanistic information available to correlate the structural data with the known biochemical properties of these molecules. We have employed steady-state kinetic and solvent viscosometric techniques to characterize the catalytic reaction pathway of the MAP kinase ERK2 with respect to the phosphorylation of a protein substrate, myelin basic protein (MBP), and a synthetic peptide substrate, ERKtide. A minor viscosity effect on k(cat) with respect to the phosphorylation of MBP was observed (k(cat) = 10 +/- 2 s(-1), k(cat)(eta) = 0.18 +/- 0.05), indicating that substrate processing occurs via slow phosphoryl group transfer (12 +/- 4 s(-1)) followed by the faster release of products (56 +/- 4 s(-1)). At an MBP concentration extrapolated to infinity, no significant viscosity effect on k(cat)/K(m(ATP)) was observed (k(cat)/K(m(ATP)) = 0.2 +/- 0.1 microM(-1) s(-1), k(cat)/K(m(ATP))(eta) = -0.08 +/- 0.04), consistent with rapid-equilibrium binding of the nucleotide. In contrast, at saturating ATP, a full viscosity effect on k(cat)/K(m) for MBP was apparent (k(cat)/K(m(MBP)) = 2.4 +/- 1 microM(-1) s(-1), k(cat)/K(m(MBP))(eta) = 1.0 +/- 0.1), while no viscosity effect was observed on k(cat)/K(m) for the phosphorylation of ERKtide (k(cat)/K(m(ERKtide)) = (4 +/- 2) x 10(-3) microM(-1) s(-1), k(cat)/K(m(ERKtide))(eta) = -0.02 +/- 0.02). This is consistent with the diffusion-limited binding of MBP, in contrast to the rapid-equilibrium binding of ERKtide, to form the ternary Michaelis complex. Calculated values for binding constants show that the estimated value for K(d(MBP)) (/= 1.5 mM). The dramatically higher catalytic efficiency of MBP in comparison to that of ERKtide ( approximately 600-fold difference) is largely attributable to the slow dissociation rate of MBP (/=56 s(-1)), from the ERK2 active site.  相似文献   

9.
恶臭假单胞菌扁桃酸消旋酶的Val22位于20 s环状结构上, 是与底物结合相关的氨基酸之一。其中Val被替换为Arg后酶活性下降了75.9%。除了酶与底物疏水作用减弱以外, 静电排斥作用增强也可能引起活性的下降。利用分子动力学模拟对酶与底物的米氏复合物进行分析, 结果表明: 突变后第22位氨基酸侧链与底物的静电势从0.036 kJ/mol升高至0.124 kJ/mol。这说明氨基酸侧链极性的改变增加了侧链与底物分子之间的静电排斥作用, 因而静电排斥作用也是导致突变体活性下降的原因之一。同时, 突变后系统势能增加了283 kJ/mol, 进一步证实了第22位氨基酸侧链极性和带电性质的改变导致酶与底物结合状态的势能增大, 从而引起活性大幅下降。因此, 将来对酶的结合口袋区域进行理性设计时, 除了考虑空间位阻效应外, 还需考虑疏水作用和静电作用。  相似文献   

10.
Mandelate racemase has been studied as a paradigm for enzyme-catalyzed abstraction of a proton from carbon acids with relatively high pKa values. 1,1-Diphenyl-1-hydroxymethylphosphonate is a substrate-intermediate-product analogue and is a modest competitive inhibitor of the enzyme (Ki=1.41+/-0.09 mM), suggesting that simultaneous binding of the two phenyl groups obviates mimicry of the aci-carboxylate function of the intermediate by the phosphonate group.  相似文献   

11.
Chen HP  Lin CF  Lee YJ  Tsay SS  Wu SH 《Journal of bacteriology》2000,182(7):2052-2054
Ornithine racemase has been purified to homogeneity from Clostridium sticklandii, as shown by sodium dodecyl sulfate-polyacrylamide gel electrophoresis. This is the first racemase known to be highly specific to ornithine. This PLP-dependent enzyme has an M(r) of 92, 000, with a K(m) for L-ornithine of 0.77 +/- 0.05 mM and a k(cat) of 980 +/- 20 s(-1).  相似文献   

12.
Studies of ligand binding to acetylcholinesterase (AChE) have demonstrated two sites of interaction. An acyl-enzyme intermediate is formed at the acylation site, and catalytic activity can be inhibited by ligand binding to a peripheral site. The three-dimensional structures of AChE-ligand complexes reveal a narrow and deep active site gorge and indicate that ligands specific for the acylation site at the base of the gorge must first traverse the peripheral site near the gorge entrance. In recent studies attempting to clarify the role of the peripheral site in the catalytic pathway for AChE, we showed that ligands which bind specifically to the peripheral site can slow the rates at which other ligands enter and exit the acylation site, a feature we called steric blockade [Szegletes, T., Mallender, W. D., and Rosenberry, T. L. (1998) Biochemistry 37, 4206-4216]. We also demonstrated that cationic substrates can form a low-affinity complex at the peripheral site that accelerates catalytic hydrolysis at low substrate concentrations but results in substrate inhibition at high concentrations because of steric blockade of product release [Szegletes, T., Mallender, W. D., Thomas, P. J., and Rosenberry, T. L. (1999) Biochemistry 38, 122-133]. In this report, we demonstrate that a key residue in the human AChE peripheral site with which the substrate acetylthiocholine interacts is D74. We extend our kinetic model to evaluate the substrate affinity for the peripheral site, indicated by the equilibrium dissociation constant K(S), from the dependence of the substrate hydrolysis rate on substrate concentration. For human AChE, a K(S) of 1.9+/-0.7 mM obtained by fitting this substrate inhibition curve agreed with a K(S) of 1.3+/-1.0 mM measured directly from acetylthiocholine inhibition of the binding of the neurotoxin fasciculin to the peripheral site. For Torpedo AChE, a K(S) of 0.5+/- 0.2 mM obtained from substrate inhibition agreed with a K(S) of 0.4+/- 0.2 mM measured with fasciculin. Introduction of the D72G mutation (corresponding to D74G in human AChE) increased the K(S) to 4-10 mM in the Torpedo enzyme and to about 33 mM in the human enzyme. While the turnover number k(cat) was unchanged in the human D74G mutant, the roughly 20-fold decrease in acetylthiocholine affinity for the peripheral site in D74G resulted in a corresponding decrease in k(cat)/K(app), the second-order hydrolysis rate constant, in the mutant. In addition, we show that D74 is important in conveying to the acylation site an inhibitory conformational effect induced by the binding of fasciculin to the peripheral site. This inhibitory effect, measured by the relative decrease in the first-order phosphorylation rate constant k(OP) for the neutral organophosphate 7-[(methylethoxyphosphonyl)oxy]-4-methylcoumarin (EMPC) that resulted from fasciculin binding, decreased from 0.002 in wild-type human AChE to 0.24 in the D74G mutant.  相似文献   

13.
We describe here a new enzyme-coupled assay for the quantitation of d-xylose using readily available enzymes that allows kinetic evaluation of hemicellulolytic enzymes using natural xylooligosaccharide substrates. Hydrogen peroxide is generated as an intermediary analyte, which allows flexibility in the choice of the chromophore or fluorophore used as the final reporter. Thus, we present d-xylose quantitation results for solution-phase assays performed with both the fluorescent reporter resorufin, generated from N-acetyl-3,7-dihydroxyphenoxazine (Amplex Red), and 2,2'-azino-bis(3-ethylbenzothiazoline-6-sulfonate) (ABTS), whose corresponding radical cation has an absorbance maximum at approximately 400 nm. We also describe a useful solid-phase variation of the assay performed with the peroxidase substrate 3,3'-diaminobenzidine tetrahydrochloride, which produces an insoluble brown precipitate. In addition, kinetic parameters for hydrolysis of the natural substrates xylobiose and xylotriose were obtained using this assay for a glycosyl hydrolase family 39 beta-xylosidase from Thermoanaerobacterium sp. strain JW/SL YS485 (Swiss-Prot accession no. O30360). At higher xylobiose substrate concentrations the enzyme showed an increase in the rate indicative of transglycosylation, while for xylotriose marked substrate inhibition was observed. At lower xylobiose concentrations k(cat) was 2.7 +/- 0.4 s(-1), K(m) was 3.3 +/- 0.7 mM, and k(cat)/K(m) was 0.82 +/- 0.21 mM(-1) . s(-1). Nonlinear curve fitting to a substrate inhibition model showed that for xylotriose K(i) was 1.7 +/- 0.1 mM, k(cat) was 2.0 +/- 0.1 s(-1), K(m) was 0.144 +/- 0.011 mM, and k(cat)/K(m) was 14 +/- 1.3 mM(-1) . s(-1).  相似文献   

14.
Mandelate racemase [EC 5.1.2.2] from Pseudomonas putida ATCC 12336 was efficiently immobilized through ionic binding onto DEAE- and TEAE 23-cellulose. The activity of the immobilized enzyme was significantly enhanced as compared to the native protein, i.e., 2.7- and 2.5-fold, respectively. DEAE-cellulose-immobilized mandelate racemase could be efficiently used in repeated batch reactions for the racemization of (R)-mandelic acid under mild conditions.  相似文献   

15.
Eubanks LM  Poulter CD 《Biochemistry》2003,42(4):1140-1149
1-Deoxy-d-xylulose 5-phosphate synthase (DXP synthase) catalyzes the thiamine diphosphate (TPP)-dependent condensation of pyruvate and d-glyceraldehyde 3-phosphate (GAP) to yield DXP in the first step of the methylerythritol phosphate pathway for isoprenoid biosynthesis. Steady-state kinetic constants for DXP synthase calculated from the initial velocities measured at varying concentrations of substrates were as follows: k(cat) = 1.9 +/- 0.1 s(-1), K(m)(GAP) = 0.068 +/- 0.001 mM, and K(m)(pyruvate) = 0.44 +/- 0.05 mM for pyruvate and GAP; k(cat) = 1.7 +/- 0.1 s(-1), K(m)(d-glyceraldehyde) = 33 +/- 3 mM, and K(m)(pyruvate) = 1.9 +/- 0.5 mM for d-glyceraldehyde and pyruvate. beta-Fluoropyruvate was investigated as a dead-end inhibitor for pyruvate. Double-reciprocal plots showed a competitive inhibition pattern with respect to pyruvate and noncompetitive inhibition with respect to GAP/d-glyceraldehyde. (14)CO(2) trapping experiments demonstrated that the binding of both substrates (pyruvate and GAP/d-glyceraldehyde) is required for the formation of a catalytically competent enzyme-substrate complex. These results are consistent with an ordered mechanism for DXP synthase where pyruvate binds before GAP/d-glyceraldehyde.  相似文献   

16.
Mandelate racemase (EC 5.1.2.2) catalyzes the interconversion of the two stereoisomers of mandelic acid. A fixed-time assay for the quantification of mandelate racemase activity has been developed. The assay involves enzymatic conversion of R-mandelate to S-mandelate (or the reverse reaction) followed by separation and detection of the substrate and product using isocratic reversed-phase high-performance liquid chromatography on a Sumichiral OA-6100 column and absorbance detection. This method offers an economical and efficient alternative to the existing circular dichroism-based and coupled assays.  相似文献   

17.
Mandelate racemase from Pseudomonas putida is a promising candidate for the dynamic kinetic resolution of α-hydroxy carboxylic acids. In the present study, the thermal stability of mandelate racemase was investigated through molecular dynamics simulations in the temperature range of 303–363 K, which can guide the design of mandelate racemase with higher stability. The basic features such as radius of gyration, surface accessibility, and secondary structure content suggested the instability of mandelate racemase at high temperatures. With increase in temperature, α-helix content reduced significantly, especially the α-helices exposed to the environment. At the simulation time scale considered, intra-protein hydrogen bonds, hydrogen bonds between protein and water decreased at 363 K, while the number of salt-bridges increased. The long-distance networks remarkably changed at 363 K. A considerable number of long-lived (percentage existence time higher than 90%) hydrogen bonds and Cα contacts were lost. Root mean square fluctuation analysis revealed regions with high fluctuation, which should be helpful in the reengineering of mandelate racemase for enhanced thermal stability.  相似文献   

18.
Guinea pig liver transglutaminase (TGase) reacts with 0.1 mM N-Cbz-L-Glu(gamma-p-nitrophenyl ester)Gly (5, prepared herein, K(M) = 0.02 mM) to undergo rapid acylation that can be followed spectrophotometrically at 400 nm (pH 7.0, 25 degrees C). Deacylation of the transiently formed thiolester acyl enzyme intermediate via catalytic aminolysis was studied in the presence of six primary amines of widely varying basicity (pK(NH+) = 5.6-10.5). Steady-state kinetic studies were performed to measure k(cat) and K(M) values for each amine substrate. A Br?nsted plot constructed through the correlation of log(k(cat)/K(M)) and pK(NH+) for each amine substrate displays a linear free-energy relationship with a slope beta(nuc) = -0.37 +/- 0.08. The shallow negative slope is consistent with a general-base-catalyzed deacylation mechanism in which a proton is removed from the amine substrate during its rate-limiting nucleophilic attack on the thiolester carbonyl. Kinetic isotope effects were measured for four acceptor substrates (water, kie = 1.1 +/- 0.1; aminoacetonitrile, kie = 5.9 +/- 1.2; glycine methyl ester, kie = 3.4 +/- 0.7; N-Ac-L-lysine methyl ester, kie = 1.1 +/- 0.1) and are consistent with a proton in flight at the rate-limiting transition state. The active site general-base implicated by these kinetic results is believed to be His-334, of the highly conserved TGase Cys-His-Asp catalytic triad.  相似文献   

19.
C Carles  B Ribadeau-Dumas 《Biochemistry》1984,23(26):6839-6843
The first steps of proteolysis of bovine beta-casein by chymosin were studied quantitatively by using reverse-phase high-performance liquid chromatography (RP-HPLC). Although chymosin has a broad specificity, it has been possible to selectively study the hydrolysis of two bonds (Ala-189-Phe-190 and Leu-192-Tyr-193) by choosing appropriate conditions. The disappearance of the substrate and the appearance of the reaction products as a function of time were followed at 220 nm by RP-HPLC. For concentrations where beta-casein was in a micellar form, the Michaelian parameters corresponding to the cleavage of bond 192-193 were determined by measuring initial rates of reaction at different substrate concentrations in a time period for which splitting of bond 189-190 was negligible. The following results were obtained; k1cat = 1.54 s-1, K1m = 0.075 mM, and k1cat/K1m = 20.6 mM-1 s-1. Under conditions where the protein was in a monomeric state, the following parameters were determined for the splitting of bond 192-193 by integrating the Michaelis equation: k2cat = 0.056 s-1, K2m = 0.007 mM, and k2cat/K2m = 79.7 mM-1 s-1. Under the latter conditions the four enzymic reactions involved in the cleavage of bonds 189-190 and 192-193 were first-order reactions. The four corresponding apparent rate constants were calculated by using a computer program. Excellent agreement was obtained between concentrations of four molecular species measured during the reaction period and those calculated by using the four apparent rate constants.  相似文献   

20.
Modeling of transition state by molecular dynamic method often requires modification of the force field parameters to describe energy profile accurately. In this work, we avoided the modification by modeling a series of mutants at binding-related site. In predicting the catalytic efficiency (k(cat)/K(m)) of the mutants of mandelate racemase (MR), the prediction performance of three energy subsets was investigated. It was indicated that the interaction-energy subset exhibited better prediction performance than whole-system subset and binding-site subset in both quantity and trend. When prediction error (PE) criterion was equal to 5%, 10 out of 12 samples were predicted correctly within interaction-energy subset, which demonstrated a great application potential of this method in prediction of enzyme catalytic efficiency and enzyme rational design.  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