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.
The glutamate dehydrogenase catalyzed reduction of delta 1-pyrroline-2-carboxylic acid (PCA; an alpha-imino acid) with reduced nicotinamide adenine dinucleotide phosphate (NADPH) to give L-proline and NADP+ is employed as a model for the redox step of the corresponding enzyme-catalyzed reductive amination of alpha-ketoglutarate. We demonstrate the reversibility of the model reaction and measure its equilibrium constant. The pH profiles for the model reactions show that the active substrates are the N-protonated imino acid in one direction and the proline anion with a neutral amino group in the other. The V/K value for the imino acid reduction is enhanced by a group Z of pK = 8.6 in the enzyme-NADPH complex, while that for the proline reaction is unaffected by any such group in the enzyme-NADP+ complex. The following conclusions emerge from a comparison of the pH dependence of the rates for the model reactions with that for the oxidative deamination of L-glutamate [Rife, J. E., & Cleland, W. W. (1980) Biochemistry 19, 2328]. The N-protonated form of alpha-iminoglutarate and the conjugate base of glutamate are the active substrates. The redox step is not sensitive to the protonation state of the groups that catalyze the hydrolysis of bound alpha-iminoglutarate. The group Z, which facilitates the PCA reaction, plays no role in the binding of alpha-ketoglutarate. We propose a chemical mechanism for the glutamate reaction where an unprotonated enzyme group of pK = 5.2 in enzyme-NADPH catalyzes the conversion of the alpha-iminoglutarate to the carbinolamine.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

2.
E A Williams  J F Morrison 《Biochemistry》1992,31(29):6801-6811
The kinetics of the NADPH-dependent reduction of 7,8-dihydrofolate, folate, and 7,8-dihydrobiopterin by human dihydrofolate reductase have been examined over the pH range from 4.0 to 9.5. The V and V/K profiles obtained with the three substrates indicate that a single ionizing residue at the active site of the enzyme must be protonated for catalysis. Both the maximum velocity of the reactions and the rate of interaction of the substrates with the enzyme-NADPH complex decrease in the order dihydrofolate greater than dihydrobiopterin much greater than folate. From the pK values of the V/K profiles, it can be concluded that, while dihydrofolate behaves as a sticky substrate and dihydrobiopterin exhibits slight stickiness, folate is not a sticky substrate. Further support for this conclusion comes from the results of deuterium isotope effects. The pK values obtained from both the V and V/Kfolate profiles are similar to the intrinsic pK value of 5.6 for both the free enzyme and the enzyme-NADPH complex. The folate analogue, 5-deazafolate, is not a substrate, but it undergoes strong interaction with the enzyme. This interaction, which is enhanced by the presence of NADPH, is due to protonation of the bound ligand that does not involve the single ionizing group at the active center of the enzyme. Difference spectra yield evidence for the protonation of bound 5-deazafolate and show that, on binding to the enzyme-NADPH complex, the pK of the N-8 atom is raised to about 10 from a value of about 4 in solution. The results are in accord with those of a recent paper on the three-dimensional structure of the enzyme-5-deazafolate complex [Davies, J.F., Delcamp, T.J., Prendergast, N.J., Ashfors, V.A., Freisheim, J.H., & Kraut, J. (1990) Biochemistry 29, 9467-9479] which indicate that there is hydrogen bond formation between N-8 of the ligand and the carbonyl group of Ile-7. However, the present findings do not support the idea that bound 5-deazafolate resembles the transition-state complex for folate reduction. Quinazolines also interact strongly with the enzyme but in a pH-independent manner. The dissociation constants for the binary complexes are an order of magnitude lower than that for the binding to the enzyme of unprotonated 5-deazafolate. This difference reflects the hydrophobic nature of the amino acid residues at the active site that are near the N-5 and N-8 nitrogens of bound pterins.  相似文献   

3.
The kinetics of glucose-fructose oxidoreductase from Zymomonas mobilis   总被引:5,自引:0,他引:5  
Glucose-fructose oxidoreductase operates by a classic ping-pong mechanism with a single site for all substrates: glucose, fructose, gluconolactone and sorbitol. The Km values for these substrates were determined. The values of kcat are 200 s-1 and 0.8 s-1 for the forward and reverse directions respectively. The overall catalytic process consists of two half-reactions with alternate reduction of NADP+ and oxidation of NADPH tightly bound to the enzyme. Reduction of enzyme-NADP+ by glucose and oxidation of enzyme-NADPH by gluconolactone involve single first-order processes. The values of the rate constants at saturating substrate are 2100 s-1 and 8 s-1 respectively; deuterium isotope effects indicate that these are for the hydrogen transfer step. Oxidation of enzyme-NADPH by fructose is first order with a limiting rate constant of at least 430 s-1. The reaction of enzyme-NADP+ with sorbitol is biphasic, with rate constants for both phases less than 1 s-1. This behaviour is explained by a mechanism in which the slow cyclisation of the acyclic form of fructose follows its dissociation from the enzyme. The rate-determining steps for the overall reaction are probably dissociation of gluconolactone in the forward direction and hydrogen transfer from sorbitol to enzyme-bound NADP+ in the reverse direction.  相似文献   

4.
Alcohol dehydrogenase (E. C. 1.1.1.1) from Thermoanaerobium brockii at 25 degrees C and at 65 degrees C is more active with secondary than primary alcohols. The enzyme utilizes NADP and NADPH as cosubstrates better than NAD and NADH. The maximum velocities (V(m)) for secondary alcohols at 65 degrees C are 10 to 100 times higher than those at 25 degrees C, whereas the K(m) values are more comparable.At both 25 degrees C and 65 degrees C the substrate analogue 1,1,1,3,3,3-hexafluoro-2-propanol inhibited the oxidation of alcohol competitively with respect to cyclopentanol, and uncompetitively with respect to NADP. Dimethylsulfoxide inhibited the reduction of cyclopentanone competitively with respect to cyclopentanone, and uncompetitively with respect to NADPH. As a product inhibitor, NADP was competitive with respect to NADPH. These results demonstrate that the enzyme binds the nucleotide and then the alcohol or ketone to form a ternary complex which is converted to a product ternary complex that releases product and nucleotide in that order.At 25 degrees C, all aldehydes and ketones examined inhibited the enzyme at concentrations above their Michaelis constants. The substrate inhibition by cyclopentanone was incomplete, and it was uncompetitive with respect to NADPH. Furthermore, cyclopentanone as a product inhibitor showed intercept-linear, slope-parabolic inhibition with respect to cyclopentanol. These results indicate that cyclopentanone binds to the enzyme-NADP complex at high concentrations. The resulting ternary complex slowly dissociates NADP and cyclopentanone.At 65 degrees C, all of the secondary alcohols, with the exception of cyclohexanol, show substrate activation at high concentration. Experiments in which NADP was the variable substrate and cyclopentanol as the constant-variable substrate over a wide range of concentrations gave double reciprocal plots in which the intercepts showed substrate activation and the slopes showed substrate inhibition. These results indicate that the secondary alcohols bind to the enzyme-NADPH complex at high concentrations and that the resulting ternary complex dissociates NADPH faster than the enzyme-NADPH complex. (c) 1993 John Wiley & Sons, Inc.  相似文献   

5.
The reduction of gluconolactone by glucose-fructose oxidoreductase containing tightly bound NADPH (enzyme-NADPH) is biphasic in nucleotide fluorescence. The initial rapid decrease, which represents quenching of the fluorescence by bound lactone, is followed by a slower decrease which corresponds to the change in absorbance. At low glucose concentrations, the oxidation of glucose by enzyme-NADP+ involves a single first-order process with similar rate constants in fluorescence and absorbance. At higher glucose concentrations, the apparent first-order rate constants for the fluorescence change are less than those for the absorbance change. This is consistent with a mechanism in which the fluorescence change occurs during the lactone dissociation step, which is slower than the hydrogen transfer step during which the absorbance change occurs. The rate constant for gluconolactone dissociation is 360 +/- 10 s-1 and this step is therefore rate-determining for the overall reaction. Reduction of fructose by enzyme-NADPH is first order with a limiting rate constant of at least 2000 s-1.  相似文献   

6.
R Srinivasan  H F Fisher 《Biochemistry》1985,24(20):5356-5360
The thermodynamic and activation parameters for the reduction of delta 1-pyrroline-2-carboxylic acid (an alpha-imino acid) by reduced nicotinamide adenine dinucleotide phosphate (NADPH) are compared with those for the reduction of the same imino acid by the glutamate dehydrogenase-NADPH complex. The enthalpies of activation and standard free energy changes for these two reactions are found to be virtually the same. The catalysis by the enzyme, expressed as the ratio of the reactivity of the enzyme--NADPH complex to that of NADPH itself in reducing the iminium ion, is entirely accounted for by a more favorable entropy of activation with enzyme--NADPH as the reductant. This entropic driving force is large enough to overcome the exergonic formation of the binary complex and still lead to considerable catalysis by glutamate dehydrogenase. Comparison of delta S not equal to and delta So values for the reduction of the iminium ion by NADPH suggests that the solvation of the transition state resembles that of the reactants, even though the substituent effects on rate have shown that the hydride transfer from the reduced coenzyme is complete at the transition state [Srinivasan, R., Medary, R. T., Fisher, H. F., Norris, D. J., & Stewart, R. (1982) J. Am. Chem. Soc. 104, 807]. The delta Go and delta S not equal to/delta So values for the reduction by the enzyme--NADPH complex indicate that this reaction has a fairly symmetric transition state, the solvation properties of which are intermediate between those of the reactants and those of the products.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

7.
Steady state kinetic analysis at pH 7.0 of the reduction of DL-glyceraldehyde by pig muscle aldose reductase showed that the enzyme follows a sequential ordered mechanism with NADPH binding first. However, the "off constant" for NADP+ in the forward direction was 1 order of magnitude less than the kcat. Analysis of this anomaly by pre-steady state kinetics using stopped-flow fluorescence spectroscopy showed that this could be accounted for by isomerization of the enzyme-NADP+ complex and that the rate of isomerization is the rate-limiting step. The rate constant for this step was of the same order of magnitude as the kcat for the forward reaction. Fluorescence emission spectra of free and NADP(H)-bound enzyme suggested a conformational change upon binding of coenzyme. In the reverse direction (oxidation of glycerol) pre-steady state and steady state kinetic analyses were consistent with the rate-limiting step occurring before isomerization of the enzyme-NADPH complex. We conclude, therefore, that during the kinetic mechanism of the reduction of aldehydes by aldose reductase, a slow (kinetically detectable) conformational change in the enzyme occurs upon coenzyme binding. Since NADPH and NADP+ bind to the enzyme very tightly, this has implications for the targeting and binding of drugs that are aldose reductase inhibitors.  相似文献   

8.
Cochliobolus lunatus 17beta-hydroxysteroid dehydrogenase (17beta-HSD) is pluripotent for several steroidal and nonsteroidal substrates. In the presence of NADPH the enzyme was found to reduce 3-keto groups of 4,5-dihydro steroids, 20-keto groups, and most efficiently, 17-keto groups of steroidal substrates. In addition, several quinones were accepted and found to be even better substrates as steroids due to their higher affinity for the enzyme-coenzyme complex and faster conversion of the enzyme-coenzyme-substrate complex into the corresponding products. As suggested by the competition studies quinones and 17-ketosteroids are converted by the same active center of the enzyme. For all tested substrates, the equilibrium ordered mechanism was established with NADPH binding first to the enzyme. According to our knowledge, the investigated 17beta-HSD is the first known fungal pluripotent enzyme of this type.  相似文献   

9.
Delta 1-Piperidine 2-carboxylic acid, an alpha-imino acid, is reduced by 1,4-dihydropyridines to pipecolic acid, an alpha-amino acid, and the corresponding pyridinium ions. This nonenzymatic reaction occurs only in the direction of pipecolic acid production. Glutamate dehydrogenase catalyzes this reaction when the reductant is NADPH and gives as products L-pipecolic acid and NADP+. The reaction velocity for the enzyme-catalyzed reaction is measurable in either direction. The pH-independent equilibrium constant, Keq, for the reduction of the imino acid by NADPH to give pipecolic acid anion and NADP+ was determined from the equilibrium conditions and the pKa values of pipecolic acid (10.72) and of the cyclic imino acid (8.10). The value of Keq was found to be 175 +/- 30; the values of delta G0, delta H0 and delta S0 are -3.1 +/- 0.1 kcal/mol, 5 +/- 1 kcal/mol and 27 +/- 4 e.u., respectively. The data indicate that the reactants are far more solvated than the products and that there must be a large degree of solvent reorganization during the course of the reaction. If these thermodynamic parameters apply to the redox step of the enzyme-catalyzed glutamate reaction, then the burst phase which results upon mixing the enzyme, L-glutamate and NADP+ in stoichiometric amounts must contain a hidden nonredox step of large delta H0 value to account for the curved Arrhenius plot observed for this phase (A.H. Colen, R.T. Medary and H.F. Fisher, Biopolymers 20 (1981) 879).  相似文献   

10.
The reaction of pentaerythritol tetranitrate reductase with reducing and oxidizing substrates has been studied by stopped-flow spectrophotometry, redox potentiometry, and X-ray crystallography. We show in the reductive half-reaction of pentaerythritol tetranitrate (PETN) reductase that NADPH binds to form an enzyme-NADPH charge transfer intermediate prior to hydride transfer from the nicotinamide coenzyme to FMN. In the oxidative half-reaction, the two-electron-reduced enzyme reacts with several substrates including nitroester explosives (glycerol trinitrate and PETN), nitroaromatic explosives (trinitrotoluene (TNT) and picric acid), and alpha,beta-unsaturated carbonyl compounds (2-cyclohexenone). Oxidation of the flavin by the nitroaromatic substrate TNT is kinetically indistinguishable from formation of its hydride-Meisenheimer complex, consistent with a mechanism involving direct nucleophilic attack by hydride from the flavin N5 atom at the electron-deficient aromatic nucleus of the substrate. The crystal structures of complexes of the oxidized enzyme bound to picric acid and TNT are consistent with direct hydride transfer from the reduced flavin to nitroaromatic substrates. The mode of binding the inhibitor 2,4-dinitrophenol (2,4-DNP) is similar to that observed with picric acid and TNT. In this position, however, the aromatic nucleus is not activated for hydride transfer from the flavin N5 atom, thus accounting for the lack of reactivity with 2,4-DNP. Our work with PETN reductase establishes further a close relationship to the Old Yellow Enzyme family of proteins but at the same time highlights important differences compared with the reactivity of Old Yellow Enzyme. Our studies provide a structural and mechanistic rationale for the ability of PETN reductase to react with the nitroaromatic explosive compounds TNT and picric acid and for the inhibition of enzyme activity with 2,4-DNP.  相似文献   

11.
The Glu alpha-carboxylate of glutathione contributes to the catalytic function of the glutathione transferases. The catalytic efficiency of human glutathione transferase A1-1 (GST A1-1) in the conjugation reaction with 1-chloro-2,4-dinitrobenzene is reduced 15 000-fold if the decarboxylated analogue of glutathione, dGSH (GABA-Cys-Gly), is used as an alternative thiol substrate. The decrease is partially due to an inability of the enzyme to promote ionization of dGSH. The pK(a) value of the thiol group of the natural substrate glutathione decreases from 9.2 to 6.7 upon binding to GST A1-1. However, the lack of the Glu alpha-carboxylate in dGSH raised the pK(a) value of the thiol in the enzymatic reaction to that of the nonenzymatic reaction. Furthermore, K(M)(dGSH) was 100-fold higher than K(M)(GSH). The active-site residue Thr68 forms a hydrogen bond to the Glu alpha-carboxylate of glutathione. Introduction of a carboxylate into GST A1-1 by a T68E mutation increased the catalytic efficiency with dGSH 10-fold and reduced the pK(a) value of the active site bound dGSH by approximately 1 pH unit. The altered pK(a) value is consistent with a catalytic mechanism where the carboxylate contributes to ionization of the glutathione thiol group. With Delta(5)-androstene-3,17-dione as substrate the efficiency of the enzyme is decreased 24 000-fold while with 4-nitrocinnamaldehyde (NCA) the decrease is less than 150-fold. In the latter reaction NCA accepts a proton and, unlike the other reactions studied, may not be dependent on the Glu alpha-carboxylate for deprotonation of the thiol group. An additional function of the Glu alpha-carboxylate may be productive orientation of glutathione within the active site.  相似文献   

12.
Zhao G  Jorns MS 《Biochemistry》2005,44(51):16866-16874
Monomeric sarcosine oxidase (MSOX) binds the L-proline zwitterion (pKa = 10.6). The reactive substrate anion is generated by ionization of the ES complex (pKa = 8.0). Tyr317 was mutated to Phe to determine whether this step might involve proton transfer to an active site base. The mutation does not eliminate the ionizable group in the ES complex (pKa = 8.9) but does cause a 20-fold decrease in the maximum rate of the reductive half-reaction. Kinetically determined Kd values for the ES complex formed with L-proline agree with results obtained in spectral titrations with the wild-type or mutant enzyme. Unlike the wild-type enzyme, Kd values with the mutant enzyme are pH-dependent, suggesting that the mutation has perturbed the pKa of a group that affects the Kd. As compared with the wild-type enzyme, an increase in charge transfer band energy is observed for mutant enzyme complexes with substrate analogues while a 10-fold decrease in the charge transfer band extinction coefficient is found for the complex with the L-proline anion. The results eliminate Tyr317 as a possible acceptor of the proton released upon substrate ionization. Since previous studies rule out the only other nearby base, we conclude that L-proline is the ionizable group in the ES complex and that amino acids are activated for oxidation upon binding to MSOX by stabilization of the reactive substrate anion. Tyr317 may play a role in substrate activation and optimizing binding, as judged by the effects of its mutation on the observed pKa, reaction rates, and charge transfer bands.  相似文献   

13.
Steady-state kinetic parameters of the human kidney aldehyde reductase-catalyzed reduction of para-substituted benzaldehydes by 3-acetyl pyridine dinucleotide phosphate (3-APADPH) were determined. The kcat of aldehyde reduction by 3-APADPH was 2- to 4-fold lower than by NADPH. The dissociation constant of 3-APADPH from the enzyme-coenzyme complex was higher (77 microM) than that of NADPH (5.3 microM). Primary deuterium kinetic isotope effects on both kcat and kcat/Km for para-substituted benzaldehyde reduction by 3-APADPH (with the exception of para-carboxybenzaldehyde) were equal and on average 2.82 +/- 0.21, suggesting that these reactions follow a rapid equilibrium-ordered reaction scheme in which the hydride transfer step is rate-limiting. Multiple regression analysis of the data suggests that benzaldehyde reduction depends upon electronic substituent effects, characterized by a rho value of 0.5. These data are consistent with a transition state in which the charge on the aldehyde carbonyl increases relative to the charge on this group in the ground state. A positive deviation of para-carboxybenzaldehyde from the linear correlation between other benzaldehydes and the substituent constant sigma + suggests a specific interaction of the carboxyl substituent of the substrate with the enzyme.  相似文献   

14.
Ketol acid reductoisomerase is an enzyme of the branched-chain amino acid biosynthetic pathway. It catalyzes two separate reactions: an acetoin rearrangement and a reduction. This paper reports on the purification of the enzyme from a recombinant Escherichia coli and on the steady-state kinetics of the enzyme. The kinetics of the reaction were determined for the forward and reverse reaction by using the appropriate chiral substrates. At saturating metal ion concentrations the mechanism follows an ordered pathway where NADPH binds before acetolactate. The product of the rearrangement of acetolactate, 3-hydroxy-3-methyl-2-oxobutyrate, is shown to be kinetically competent as an intermediate in the enzyme-catalyzed reaction. Starting with acetolactate, Mg2+ is the only divalent metal ion that will support enzyme catalysis. For the reduction of 3-hydroxy-3-methyl-2-oxobutyrate, Mn2+ is catalytically active. Product and dead-end inhibition studies indicate that the binding of metal ion and NADPH occurs randomly. In the forward reaction direction, the deuterium kinetic isotope effect on V/K is 1.07 when acetolactate is the substrate and 1.39 when 3-hydroxy-3-methyl-2-oxobutyrate is the substrate.  相似文献   

15.
The structure of pentaerythritol tetranitrate (PETN) reductase in complex with the nitroaromatic substrate picric acid determined previously at 1.55 A resolution indicated additional electron density between the indole ring of residue Trp-102 and the nitro group at C-6 of picrate. The data suggested the presence of an unusual bond between substrate and the tryptophan side chain. Herein, we have extended the resolution of the PETN reductase-picric acid complex to 0.9 A. This high-resolution analysis indicates that the active site is partially occupied with picric acid and that the anomalous density seen in the original study is attributed to the population of multiple conformational states of Trp-102 and not a formal covalent bond between the indole ring of Trp-102 and picric acid. The significance of any interaction between Trp-102 and nitroaromatic substrates was probed further in solution and crystal complexes with wild-type and mutant (W102Y and W102F) enzymes. Unlike with wild-type enzyme, in the crystalline form picric acid was bound at full occupancy in the mutant enzymes, and there was no evidence for multiple conformations of active site residues. Solution studies indicate tighter binding of picric acid in the active sites of the W102Y and W102F enzymes. Mutation of Trp-102 does not impair significantly enzyme reduction by NADPH, but the kinetics of decay of the hydride-Meisenheimer complex are accelerated in the mutant enzymes. The data reveal that decay of the hydride-Meisenheimer complex is enzyme catalyzed and that the final distribution of reaction products for the mutant enzymes is substantially different from wild-type enzyme. Implications for the mechanism of high explosive degradation by PETN reductase are discussed.  相似文献   

16.
Zhao G  Jorns MS 《Biochemistry》2002,41(31):9747-9750
Monomeric sarcosine oxidase (MSOX) contains covalently bound FAD and catalyzes the oxidation of sarcosine (N-methylglycine) and other secondary amino acids, including L-proline. The reductive half-reaction with L-proline proceeds via a rapidly attained equilibrium (K(d)) between free E(ox) and the E(ox).S complex, followed by a practically irreversible reduction step (E(ox).S --> E(red).P) associated with a rate constant, k(lim). The effect of pH on the reductive half-reaction shows that the K(d) for L-proline binding is pH-independent (pH 6.46-9.0). This indicates that MSOX binds the zwitterionic form of L-proline, the predominant species in solution at neutral pH (pK(a) = 10.6). Values for the limiting rate of reduction (k(lim)) are, however, strongly pH-dependent and indicate that an ionizable group in the E(ox).L-proline complex (pK(a) = 8.02) must be unprotonated for conversion to E(red).P. Charge-transfer interaction with L-proline as the donor and FAD as acceptor is possible only with the anionic form of L-proline. The ionizable group in the E(ox).L-proline complex is required for conversion of enzyme-bound L-proline from the zwitterionic to the reactive anionic form, as judged by the independently determined pK(a) for charge-transfer complex formation with the MSOX flavin (pK(a) = 7.94). The observation that the anionic form of L-proline with a neutral amino group is the reactive species in the reduction of MSOX is similar to that observed for other flavoenzymes that oxidize amines, including monoamine oxidase and trimethylamine dehydrogenase.  相似文献   

17.
1. The oxidation of NADPH2 by leucocyte granules, as measured at acid pH in the presence of Mn-2+, was found to be inhibited by superoxide dismutase. 2. Omission of Mn-2+ markedly lowered the oxidase activity at acid pH, which was still inhibited by superoxide dismutase. 3. At alkaline pH the oxidase activity was lower than at acid pH. 4. During oxidation of NADPH2 by leucocyte granules, reduction of cytochrome c occurred which was partially inhibited by superoxide dismutase. 5. It was concluded that NADPH2 oxidation occurs through an enzymatic reaction and a nonenzymatic chain reaction. Superoxide anion (O-minus-2 and NADPH- free radical would be involved in the chain reaction. The differential sensitivity of NADPH2 oxidation to superoxide dismutase in different experimental conditions (see above 1, 2 and 3) was explained on the basis of changes in the properties of the chain reaction.  相似文献   

18.
Mammalian thioredoxin reductase catalyzes NADPH dependent reduction of a wide variety of substrates and plays a central role in redox regulation and antioxidant defence. Recently the enzyme was discovered to be a selenoprotein with a catalytically active penultimate selenocysteine residue. Dinitrohalobenzenes irreversibly inhibit the enzyme with a concomitant induction of an NADPH oxidase activity, producing superoxide. A model explaining the reactivity of dinitrohalobenzenes with thioredoxin reductase is presented, involving dinitrophenyl-derivatization of both the selenocysteine residue and its neighboring cysteine residue, reduction by NADPH of the enzyme-bound flavin in dinitrophenyl-alkylated enzyme (dnp-TrxR), followed by two consecutive one-electron transfers from the flavin to nitro groups of the dnp-moieties in dnp-TrxR, forming nitro anion radicals. The nitro radicals react with oxygen to form superoxide, again generating dnp-TrxR with an oxidized flavin, which may then follow another cycle of NADPH-dependent superoxide production. Dinitrohalobenzene compounds are well known for their immunostimulatory properties. Here it is proposed that the inflammatory components of this immunostimulation can be mediated by interaction with the thioredoxin system, via effects on cell function by superoxide production, oxidative stress and increased extracellular levels of thioredoxin.  相似文献   

19.
The nucleotides 8-amino-, 8-methylamino-, and 8-dimethylaminoadenylic acid have been synthesized and their preferred conformations about the glycosyl bond in qaueous solution have been determined by 1H nuclear magnetic resonance spectroscopy. Paramagnetic relaxation studies, nuclear Overhauser enhancement measurements, chemical shifts, and coupling constant comparisons indicate that their is rotation about the glycosyl bond and that preference for either the anti or syn conformation depends on the extent of alkyl substitution on the 8-amino group. The primary and secondary amines 8-amino- and 8-methylaminoadenylic acid adopt a perferential anti conformation about the glycosyl bond, while the tertiary amine 8-dimethylaminoadenylic acid exists predominantly in the syn form. These three analogs provide a system to study interactions of a dehydrogenase with coenzyme inhibitors having different glycosyl conformer populations. All three analogs are competitive inhibitors of NADH in reaction with chicken muscle lactate dehydrogenase, and the Ki values show little dependence on the nature of the amino substitution. This demonstrates that the distribution of conformations about the nucleotide glycosyl bond does not effect the competition of the nucleotide for lactate dehydrogenase apoenzyme. Several models for enzyme-coenzyme binding are discussed. The available data cannot distinguish whether the enzyme binds nucleotide in both the anti and syn conformations or in purely the anti conformation. However, at some stage of the enzyme-coenzyme interaction, there appears to be a strong stabilization of the nucleotide in the anti conformation about the glycosyl bond.  相似文献   

20.
An enzymatic asymmetric synthesis was carried out for the preparation of enantiomerically pure L-diphenylalanine using the rationally engineered aromatic L-amino acid transaminase (eAroATEs) obtained from Enterobacter sp. BK2K-1. To rationally redesign the enzyme, structural model was constructed by the homology modeling. The structural model was experimentally validated by the site-directed mutagenesis of the predicted pyridoxal-5'-phosphate (PLP) binding site and the substrate-recognition region, and the cell-free protein synthesis of mutated enzymes. It was suggested that Arg281 and Arg375 were the key residues to recognize the distal carboxylate and alpha-carboxylate group of the substrates, respectively. The model also predicted that Tyr66 forms hydrogen bond with the phosphate moiety of PLP and interacts with the side chain attached to beta-carbon of the amino acid substrate. Among the various site-directed mutants, Y66L variant was able to synthesize L-diphenylalanine with 23% conversion yield for 10 h, whereas the wild-type AroATEs was inactive for the transamination between diphenylpyruvate and L-phenylalanine as amino acceptor and amino donor, respectively.  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