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 roles of His181, His184 and Tyr186 in PETN reductase have been examined by mutagenesis, spectroscopic and stopped-flow kinetics, and by determination of crystallographic structures for the Y186F PETN reductase and reduced wild-type enzyme-progesterone complex. Residues His181 and His184 are important in the binding of coenzyme, steroids, nitroaromatic ligands and the substrate 2-cyclohexen-1-one. The H181A and H184A enzymes retain activity in reductive and oxidative half-reactions, and thus do not play an essential role in catalysis. Ligand binding and catalysis is not substantially impaired in Y186F PETN reductase, which contrasts with data for the equivalent mutation (Y196F) in Old Yellow Enzyme. The structure of Y186F PETN reductase is identical to wild-type enzyme, with the obvious exception of the mutation. We show in PETN reductase that Tyr186 is not a key proton donor in the reduction of alpha/beta unsaturated carbonyl compounds. The structure of two electron-reduced PETN reductase bound to the inhibitor progesterone mimics the catalytic enzyme-steroid substrate complex and is similar to the structure of the oxidized enzyme-inhibitor complex. The reactive C1-C2 unsaturated bond of the steroid is inappropriately orientated with the flavin N5 atom for hydride transfer. With steroid substrates, the productive conformation is achieved by orientating the steroid through flipping by 180 degrees , consistent with known geometries for hydride transfer in flavoenzymes. Our data highlight mechanistic differences between Old Yellow Enzyme and PETN reductase and indicate that catalysis requires a metastable enzyme-steroid complex and not the most stable complex observed in crystallographic studies.  相似文献   

2.
Nidetzky B  Klimacek M  Mayr P 《Biochemistry》2001,40(34):10371-10381
Microbial xylose reductase, a representative aldo-keto reductase of primary sugar metabolism, catalyzes the NAD(P)H-dependent reduction of D-xylose with a turnover number approximately 100 times that of human aldose reductase for the same reaction. To determine the mechanistic basis for that physiologically relevant difference and pinpoint features that are unique to the microbial enzyme among other aldo/keto reductases, we carried out stopped-flow studies with wild-type xylose reductase from the yeast Candida tenuis. Analysis of transient kinetic data for binding of NAD(+) and NADH, and reduction of D-xylose and oxidation of xylitol at pH 7.0 and 25 degrees C provided estimates of rate constants for the following mechanism: E + NADH right arrow over left arrow E.NADH right arrow over left arrow E.NADH + D-xylose right arrow over left arrow E.NADH.D-xylose right arrow over left arrow E.NAD(+).xylitol right arrow over left arrow E.NAD(+) right arrow over left arrow E.NAD(+) right arrow over left arrow E + NAD(+). The net rate constant of dissociation of NAD(+) is approximately 90% rate limiting for k(cat) of D-xylose reduction. It is controlled by the conformational change which precedes nucleotide release and whose rate constant of 40 s(-)(1) is 200 times that of completely rate-limiting E.NADP(+) --> E.NADP(+) step in aldehyde reduction catalyzed by human aldose reductase [Grimshaw, C. E., et al. (1995) Biochemistry 34, 14356-14365]. Hydride transfer from NADH occurs with a rate constant of approximately 170 s(-1). In reverse reaction, the E.NADH --> E.NADH step takes place with a rate constant of 15 s(-1), and the rate constant of ternary-complex interconversion (3.8 s(-1)) largely determines xylitol turnover (0.9 s(-1)). The bound-state equilibrium constant for C. tenuis xylose reductase is estimated to be approximately 45 (=170/3.8), thus greatly favoring aldehyde reduction. Formation of productive complexes, E.NAD(+) and E.NADH, leads to a 7- and 9-fold decrease of dissociation constants of initial binary complexes, respectively, demonstrating that 12-fold differential binding of NADH (K(i) = 16 microM) vs NAD(+) (K(i) = 195 microM) chiefly reflects difference in stabilities of E.NADH and E.NAD(+). Primary deuterium isotope effects on k(cat) and k(cat)/K(xylose) were, respectively, 1.55 +/- 0.09 and 2.09 +/- 0.31 in H(2)O, and 1.26 +/- 0.06 and 1.58 +/- 0.17 in D(2)O. No deuterium solvent isotope effect on k(cat)/K(xylose) was observed. When deuteration of coenzyme selectively slowed the hydride transfer step, (D)()2(O)(k(cat)/K(xylose)) was inverse (0.89 +/- 0.14). The isotope effect data suggest a chemical mechanism of carbonyl reduction by xylose reductase in which transfer of hydride ion is a partially rate-limiting step and precedes the proton-transfer step.  相似文献   

3.
Roitel O  Scrutton NS  Munro AW 《Biochemistry》2003,42(36):10809-10821
Cys-999 is one component of a triad (Cys-999, Ser-830, and Asp-1044) located in the FAD domain of flavocytochrome P450 BM3 that is almost entirely conserved throughout the diflavin reductase family of enzymes. The role of Cys-999 has been studied by steady-state kinetics, stopped-flow spectroscopy, and potentiometry. The C999A mutants of BM3 reductase (containing both FAD and FMN cofactors) and the isolated FAD domain are substantially compromised in their capacity to reduce artificial electron acceptors in steady-state turnover with either NADPH or NADH as electron donors. Stopped-flow studies indicate that this is due primarily to a substantially slower rate of hydride transfer from nicotinamide coenzyme to FAD cofactor in the C999A enzymes. The compromised rates of hydride transfer are not attributable to altered thermodynamic properties of the flavins. A reduced enzyme-NADP(+) charge-transfer species is populated following hydride transfer in the wild-type FAD domain, consistent with the slow release of NADP(+) from the 2-electron-reduced enzyme. This intermediate does not accumulate in the C999A FAD domain or wild-type and C999A BM3 reductases, suggesting more rapid release of NADP(+) from these enzyme forms. Rapid internal electron transfer from FAD to FMN in wild-type BM3 reductase releases NADP(+) from the nicotinamide-binding site, thus preventing the inhibition of enzyme activity through the accumulation of a stable FADH(2)-NADP(+) charge-transfer complex. Hydride transfer is reversible, and the observed rate of oxidation of the 2-electron-reduced C999A BM3 reductase and FAD domain is hyperbolically dependent on NADP(+) concentration. With the wild-type BM3 reductase and FAD domain, the rate of flavin oxidation displays an unusual dependence on NADP(+) concentration, consistent with a two-site binding model in which two coenzyme molecules bind to catalytic and regulatory regions (or sites) within a bipartite coenzyme binding site. A kinetic model is proposed in which binding of coenzyme to the regulatory site hinders sterically the release of NADPH from the catalytic site. The results are discussed in the light of kinetic and structural studies on mammalian cytochrome P450 reductase.  相似文献   

4.
The NADH-dependent persulfide reductase (Npsr), a recently discovered member of the PNDOR family of flavoproteins that contains both the canonical flavoprotein reductase domain and a rhodanese domain, is proposed to be involved in the dissimilatory reduction of S(0) for Shewanella loihica PV-4. We have previously shown that polysulfide is a substrate for this enzyme, and a recently determined structure of a closely related enzyme (CoADR-Rhod from Bacillus anthracis) suggested the importance of a bound coenzyme A in the mechanism. The work described here shows that the in vivo oxidizing substrates of Npsr are the persulfides of small thiols such as CoA and glutathione. C43S, C531S, and C43,531S mutants were created to determine the role of the flavoprotein domain cysteine (C43) and the rhodanese domain cysteine (C531) in the mechanism. The absolute requirement for C43 in persulfide or DTNB reductase activity shows that this residue is involved in S-S bond breakage. C531 contributes to, but is not required for, catalysis of DTNB reduction, while it is absolutely required for reduction of any persulfide substrates. Titrations of the enzyme with NADH, dithionite, titanium(III), or TCEP demonstrate the presence of a mixed-disulfide between C43 and a tightly bound CoA, and structures of the C43 and C43,531S mutants confirm that this coenzyme A remains tightly bound to the enzyme in the absence of a C43-CoA S-S bond. The structure of Npsr suggests a likely site for binding and reaction with the persulfide substrate on the rhodanese domain. On the basis of kinetic, titration, and structural data, a mechanism for the reduction of persulfides by Npsr is proposed.  相似文献   

5.
Crystallographic analysis revealed that the nicotinamide ring of NAD can bind with multiconformations to aldehyde dehydrogenase (ALDH) (Ni, L., Zhou, J., Hurley, T. D., and Weiner, H. (1999) Protein Sci. 8, 2784-2790). Electron densities can be defined for two conformations, neither of which appears to be compatible with the catalytic reaction. In one conformation, it would prevent glutamate 268 from functioning as a general base needed to activate the catalytic nucleophile, cysteine 302. In the other conformation, the nicotinamide is too far from the enzyme-substrate adduct for efficient hydride transfer. In this study, NMR and fluorescence spectroscopies were used to demonstrate that NAD and NADH bind to human liver cytosol and mitochondrial ALDH such that the nicotinamide samples a population of conformations while the adenosine region remains relatively immobile. Although the nicotinamide possesses extensive conformational heterogeneity, the catalyzed reaction leads to the stereospecific transfer of hydride to the coenzyme. Mobility allows the nicotinamide to move into position to be reduced by the enzyme-substrate adduct. Although the reduced nicotinamide ring retains mobility after NADH formation, the extent of the motion is less than that of NAD. It appears that after reduction the population of favored nicotinamide conformations shifts toward those that do not interfere with the ability of the enzyme to release the reaction product. In the case of the mitochondrial, but not the cytosolic, enzyme this change in conformational preference is promoted by the presence of Mg2+ ions. Coenzyme conformational mobility appears to be beneficial to catalysis by ALDH throughout the catalytic cycle.  相似文献   

6.
Rubach JK  Plapp BV 《Biochemistry》2003,42(10):2907-2915
Amino acid residues Thr-178, Val-203, and Val-292, which interact with the nicotinamide ring of the coenzyme bound to alcohol dehydrogenase (ADH), may facilitate hydride transfer and hydrogen tunneling by orientation and dynamic effects. The T178S, T178V, V203A, V292A, V292S, and V292T substitutions significantly alter the steady state and transient kinetics of the enzyme. The V292A, V292S, and V292T enzymes have decreased affinity for coenzyme (NAD+ by 30-50-fold and NADH by 35-75-fold) as compared to the wild-type enzyme. The substitutions in the nicotinamide binding site decrease the rate constant of hydride transfer for benzyl alcohol oxidation by 3-fold (for V292T ADH) to 16-fold (for V203A ADH). The modest effects suggest that catalysis does not depend critically on individual residues and that several residues in the nicotinamide binding site contribute to catalysis. The structures of the V292T ADH-NAD+-pyrazole and wild-type ADH-NAD+-4-iodopyrazole ternary complexes are very similar. Only subtle changes in the V292T enzyme cause the large changes in coenzyme binding and the small change in hydride transfer. In these complexes, one pyrazole nitrogen binds to the catalytic zinc, and the other nitrogen forms a partial covalent bond with C4 of the nicotinamide ring, which adopts a boat conformation that is postulated to be relevant for hydride transfer. The results provide an experimental basis for evaluating the contributions of dynamics to hydride transfer.  相似文献   

7.
The three-component toluene dioxygenase system consists of an FAD-containing reductase, a Rieske-type [2Fe-2S] ferredoxin, and a Rieske-type dioxygenase. The task of the FAD-containing reductase is to shuttle electrons from NADH to the ferredoxin, a reaction the enzyme has to catalyze in the presence of dioxygen. We investigated the kinetics of the reductase in the reductive and oxidative half-reaction and detected a stable charge transfer complex between the reduced reductase and NAD+ at the end of the reductive half-reaction, which is substantially less reactive toward dioxygen than the reduced reductase in the absence of NAD+. A plausible reason for the low reactivity toward dioxygen is revealed by the crystal structure of the complex between NAD+ and reduced reductase, which shows that the nicotinamide ring and the protein matrix shield the reactive C4a position of the isoalloxazine ring and force the tricycle into an atypical planar conformation, both factors disfavoring the reaction of the reduced flavin with dioxygen. A rapid electron transfer from the charge transfer complex to electron acceptors further reduces the risk of unwanted side reactions, and the crystal structure of a complex between the reductase and its cognate ferredoxin shows a short distance between the electron-donating and -accepting cofactors. Attraction between the two proteins is likely mediated by opposite charges at one large patch of the complex interface. The stability, specificity, and reactivity of the observed charge transfer and electron transfer complexes are thought to prevent the reaction of reductaseTOL with dioxygen and thus present a solution toward conflicting requirements.  相似文献   

8.
The mechanism of flavin reduction in morphinone reductase (MR) and pentaerythritol tetranitrate (PETN) reductase, and flavin oxidation in MR, has been studied by stopped-flow and steady-state kinetic methods. The temperature dependence of the primary kinetic isotope effect for flavin reduction in MR and PETN reductase by nicotinamide coenzyme indicates that quantum mechanical tunneling plays a major role in hydride transfer. In PETN reductase, the kinetic isotope effect (KIE) is essentially independent of temperature in the experimentally accessible range, contrasting with strongly temperature-dependent reaction rates, consistent with a tunneling mechanism from the vibrational ground state of the reactive C-H/D bond. In MR, both the reaction rates and the KIE are dependent on temperature, and analysis using the Eyring equation suggests that hydride transfer has a major tunneling component, which, unlike PETN reductase, is gated by thermally induced vibrations in the protein. The oxidative half-reaction of MR is fully rate-limiting in steady-state turnover with the substrate 2-cyclohexenone and NADH at saturating concentrations. The KIE for hydride transfer from reduced flavin to the alpha/beta unsaturated bond of 2-cyclohexenone is independent of temperature, contrasting with strongly temperature-dependent reaction rates, again consistent with ground-state tunneling. A large solvent isotope effect (SIE) accompanies the oxidative half-reaction, which is also independent of temperature in the experimentally accessible range. Double isotope effects indicate that hydride transfer from the flavin N5 atom to 2-cyclohexenone, and the protonation of 2-cyclohexenone, are concerted and both the temperature-independent KIE and SIE suggest that this reaction also proceeds by ground-state quantum tunneling. Our results demonstrate the importance of quantum tunneling in the reduction of flavins by nicotinamide coenzymes. This is the first observation of (i) three H-nuclei in an enzymic reaction being transferred by tunneling and (ii) the utilization of both passive and active dynamics within the same native enzyme.  相似文献   

9.
Recombinant house fly (Musca domestica) cytochrome P450 reductase has been purified by anion exchange and affinity chromatography. Steady-state kinetics of cytochrome c reductase activity revealed a random Bi-Bi mechanism with formation of a ternary P450 reductase-NADPH-electron acceptor complex as catalytic intermediate. NADP(H) binding is essential for fast hydride ion transfer to FAD, as well as for electron transfer from FMN to cytochrome c. Reduced cytochrome c had no effect on the enzyme activity, while NADP+ and 2'-AMP inhibited P450 reductase competitively with respect to NADPH and noncompetitively with respect to cytochrome c. The affinity of the P450 reductase to NADPH is 10 times higher than to NADP+ (Kd of 0.31 and 3.3 microM, respectively). Such an affinity change during catalysis could account for a +30 mV shift of the redox potential of FAD. Cys560 was substituted for Tyr by site-directed mutagenesis. This mutation decreased enzyme affinity to NADPH 35-fold by decreasing the bimolecular rate constant of nucleotide binding with no detectable effect on the kinetic mechanism. The affinity of the C560Y mutant enzyme to NADP+ decreased 9-fold compared to the wild-type enzyme, while the affinity to 2'-AMP was not significantly affected, suggesting that Cys560 is located in the nicotinamide binding site of the active, full-size enzyme in solution.  相似文献   

10.
Rubach JK  Ramaswamy S  Plapp BV 《Biochemistry》2001,40(42):12686-12694
The participation of Val-292 in catalysis by alcohol dehydrogenase and the involvement of dynamics were investigated. Val-292 interacts with the nicotinamide ring of the bound coenzyme and may facilitate hydride transfer. The substitution of Val-292 with Ser (V292S) increases the dissociation constants for the coenzymes (NAD(+) by 50-fold, NADH by 75-fold) and the turnover numbers by 3-7-fold. The V292S enzyme crystallized in the presence of NAD(+) and 2,3,4,5,6-pentafluorobenzyl alcohol has an open conformation similar to the structure of the wild-type apo-enzyme, rather than the closed conformation observed for ternary complexes with wild-type enzyme. The V292S substitution perturbs the conformational equilibrium of the enzyme and decreases the kinetic complexity, which permits study of the hydride transfer step with steady-state kinetics. Eyring plots show that the DeltaH for the oxidation (V(1)) of the protio and deuterio benzyl alcohols is 13 kcal/mol and that the kinetic isotope effect of 4.1 is essentially temperature-independent. Eyring plots for the catalytic efficiency for reduction of benzaldehyde (V(2)/K(p)) with NADH or NADD are distinctly convex, being temperature-dependent from 5 to 25 degrees C and temperature-independent from 25 to 50 degrees C; the kinetic isotope effect of 3.2 for V(2)/K(p) is essentially independent of the temperature. The temperature dependencies and isotope effects for V(1) and V(2)/K(p) are not adequately explained by semiclassical transition state theory and are better explained by hydride transfer occurring through vibrationally assisted tunneling.  相似文献   

11.
Biliverdin reductase (BVR) catalyzes the last step in heme degradation by reducing the gamma-methene bridge of the open tetrapyrrole, biliverdin IXalpha, to bilirubin with the concomitant oxidation of a beta-nicotinamide adenine dinucleotide (NADH) or beta-nicotinamide adenine dinucleotide phosphate (NADPH) cofactor. Bilirubin is the major bile pigment in mammals and has antioxidant and anticompliment activity. We have determined X-ray crystal structures of apo rat BVR and its complex with NADH at 1.2 A and 1.5 A resolution, respectively. In agreement with an independent structure determination of the apo-enzyme, BVR consists of an N-terminal dinucleotide-binding domain (Rossmann-fold) and a C-terminal domain that contains a six-stranded beta-sheet that is flanked on one face by several alpha-helices. The C-terminal and N-terminal domains interact extensively, forming the active site cleft at their interface. The cofactor complex structure reported here reveals that the cofactor nicotinamide ring extends into the active site cleft, where it is adjacent to conserved amino acid residues and, consistent with the known stereochemistry of the reaction catalyzed by BVR, the si face of the ring is accessible for hydride transfer. The only titratable side-chain that appears to be suitably positioned to function as a general acid in catalysis is Tyr97. This residue, however, is not essential for catalysis, since the Tyr97Phe mutant protein retains 50% activity. This finding suggests that the dominant role in catalysis may be performed by hydride transfer from the cofactor, a process that may be promoted by proximity of the invariant residues Glu96, Glu123, and Glu126, to the nicotinamide ring.  相似文献   

12.
Previous studies have shown that the interaction of P450 reductase with bound NADP(H) is essential to ensure fast electron transfer through the two flavin cofactors. In this study we investigated in detail the interaction of the house fly flavoprotein with NADP(H) and a number of nucleotide analogues. 1,4,5,6-Tetrahydro-NADP, an analogue of NADPH, was used to characterize the interaction of P450 reductase with the reduced nucleotide. This analogue is inactive as electron donor, but its binding affinity and rate constant of release are very close to those for NADPH. The 2'-phosphate contributes about 5 kcal/mol of the binding energy of NADP(H). Oxidized nicotinamide does not interact with the oxidized flavoprotein, while reduced nicotinamide contributes 1.3 kcal/mol of the binding energy. Oxidized P450 reductase binds NADPH with a K(d) of 0.3 microM, while the affinity of the reduced enzyme is considerably lower, K(d) = 1.9 microM. P450 reductase catalyzes a transhydrogenase reaction between NADPH and oxidized nucleotides, such as thionicotinamide-NADP(+), acetylpyridine-NADP(+), or [(3)H]NADP(+). The reverse reaction, reduction of [(3)H]NADP(+) by the reduced analogues, is also catalyzed by P450 reductase. We define the mechanism of the transhydrogenase reaction as follows: NADPH binding, hydride ion transfer, and release of the NADP(+) formed. An NADP(+) or its analogue binds to the two-electron-reduced flavoprotein, and the electron-transfer steps reverse to transfer hydride ion to the oxidized nucleotide, which is released. Measurements of the flavin semiquinone content, rate constant for NADPH release, and transhydrogenase turnover rates allowed us to estimate the steady-state distribution of P450 reductase species during catalysis, and to calculate equilibrium constants for the interconversion of catalytic intermediates. Our results demonstrate that equilibrium redox potentials of the flavin cofactors are not the sole factor governing rapid electron transfer during catalysis, but conformational changes must be considered to understand P450 reductase catalysis.  相似文献   

13.
Jackson JB 《FEBS letters》2003,545(1):18-24
Transhydrogenase, in animal mitochondria and bacteria, couples hydride transfer between NADH and NADP(+) to proton translocation across a membrane. Within the protein, the redox reaction occurs at some distance from the proton translocation pathway and coupling is achieved through conformational changes. In an 'open' conformation of transhydrogenase, in which substrate nucleotides bind and product nucleotides dissociate, the dihydronicotinamide and nicotinamide rings are held apart to block hydride transfer; in an 'occluded' conformation, they are moved into apposition to permit the redox chemistry. In the two monomers of transhydrogenase, there is a reciprocating, out-of-phase alternation of these conformations during turnover.  相似文献   

14.
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   

15.
As shown by X-ray crystallography, horse liver alcohol dehydrogenase undergoes a global conformational change upon binding of NAD+ or NADH, involving a rotation of the catalytic domain relative to the coenzyme binding domain and the closing up of the active site to produce a catalytically efficient enzyme. The conformational change requires a complete coenzyme and is affected by various chemical or mutational substitutions that can increase the catalytic turnover by altering the kinetics of the isomerization and rate of dissociation of coenzymes. The binding of NAD+ is kinetically limited by a unimolecular isomerization (corresponding to the conformational change) that is controlled by deprotonation of the catalytic zinc-water to produce a negatively-charged zinc-hydroxide, which can attract the positively-charged nicotinamide ring. The deprotonation is facilitated by His-51 acting through a hydrogen-bonded network to relay the proton to solvent. Binding of NADH also involves a conformational change, but the rate is very fast. After the enzyme binds NAD+ and closes up, the substrate displaces the hydroxide bound to the catalytic zinc; this exchange may involve a double displacement reaction where the carboxylate group of a glutamate residue first displaces the hydroxide (inverting the tetrahedral coordination of the zinc), and then the exogenous ligand displaces the glutamate. The resulting enzyme-NAD+-alcoholate complex is poised for hydrogen transfer, and small conformational fluctuations may bring the reactants together so that the hydride ion is transferred by quantum mechanical tunneling. In the process, the nicotinamide ring may become puckered, as seen in structures of complexes of the enzyme with NADH. The conformational changes of alcohol dehydrogenase demonstrate the importance of protein dynamics in catalysis.  相似文献   

16.
17.
Proton-translocating transhydrogenase is found in the inner membranes of animal mitochondria, and in the cytoplasmic membranes of many bacteria. It catalyses hydride transfer from NADH to NADP(+) coupled to inward proton translocation. Evidence is reviewed suggesting the enzyme operates by a "binding-change" mechanism. Experiments with Escherichia coli transhydrogenase indicate the enzyme is driven between "open" and "occluded" states by protonation and deprotonation reactions associated with proton translocation. In the open states NADP(+)/NADPH can rapidly associate with, or dissociate from, the enzyme, and hydride transfer is prevented. In the occluded states bound NADP(+)/NADPH cannot dissociate, and hydride transfer is allowed. Crystal structures of a complex of the nucleotide-binding components of Rhodospirillum rubrum transhydrogenase show how hydride transfer is enabled and disabled at appropriate steps in catalysis, and how release of NADP(+)/NADPH is restricted in the occluded state. Thermodynamic and kinetic studies indicate that the equilibrium constant for hydride transfer on the enzyme is elevated as a consequence of the tight binding of NADPH relative to NADP(+). The protonation site in the translocation pathway must face the outside if NADP(+) is bound, the inside if NADPH is bound. Chemical shift changes detected by NMR may show where alterations in protein conformation resulting from NADP(+) reduction are initiated. This article is part of a Special Issue entitled: 17th European Bioenergetics Conference (EBEC 2012).  相似文献   

18.
In rat neuronal nitric oxide synthase, Phe1395 is positioned over the FAD isoalloxazine ring. This is replaced by Trp676 in human cytochrome P450 reductase, a tryptophan in related diflavin reductases (e.g. methionine synthase reductase and novel reductase 1), and tyrosine in plant ferredoxin-NADP(+) reductase. Trp676 in human cytochrome P450 reductase is conformationally mobile, and plays a key role in enzyme reduction. Mutagenesis of Trp676 to alanine results in a functional NADH-dependent reductase. Herein, we describe studies of rat neuronal nitric oxide synthase FAD domains, in which the aromatic shielding residue Phe1395 is replaced by tryptophan, alanine and serine. In steady-state assays the F1395A and F1395S domains have a greater preference for NADH compared with F1395W and wild-type. Stopped-flow studies indicate flavin reduction by NADH is significantly faster with F1395S and F1395A domains, suggesting that this contributes to altered preference in coenzyme specificity. Unlike cytochrome P450 reductase, the switch in coenzyme specificity is not attributed to differential binding of NADPH and NADH, but probably results from improved geometry for hydride transfer in the F1395S- and F1395A-NADH complexes. Potentiometry indicates that the substitutions do not significantly perturb thermodynamic properties of the FAD, although considerable changes in electronic absorption properties are observed in oxidized F1395A and F1395S, consistent with changes in hydrophobicity of the flavin environment. In wild-type and F1395W FAD domains, prolonged incubation with NADPH results in development of the neutral blue semiquinone FAD species. This reaction is suppressed in the mutant FAD domains lacking the shielding aromatic residue.  相似文献   

19.
Perez-Miller SJ  Hurley TD 《Biochemistry》2003,42(23):7100-7109
Crystal structures of many enzymes in the aldehyde dehydrogenase superfamily determined in the presence of bound NAD(P)(+) have exhibited conformational flexibility for the nicotinamide half of the cofactor. This has been hypothesized to be important in catalysis because one conformation would block the second half of the reaction, but no firm evidence has been put forth which shows whether the oxidized and reduced cofactors preferentially occupy the two observed conformations. We present here two structures of the wild type and two structures of a Cys302Ser mutant of human mitochondrial aldehyde dehydrogenase in binary complexes with NAD(+) and NADH. These structures, including the Cys302Ser mutant in complex with NAD(+) at 1.4 A resolution and the wild-type enzyme in complex with NADH at 1.9 A resolution, provide strong evidence that bound NAD(+) prefers an extended conformation ideal for hydride transfer and bound NADH prefers a contracted conformation ideal for acyl-enzyme hydrolysis. Unique interactions between the cofactor and the Rossmann fold make isomerization possible while allowing the remainder of the active site complex to remain intact. In addition, these structures clarify the role of magnesium in activating the human class 2 enzyme. Our data suggest that the presence of magnesium may lead to selection of particular conformations and speed isomerization of the reduced cofactor following hydride transfer.  相似文献   

20.
Argyrou A  Blanchard JS 《Biochemistry》2004,43(14):4375-4384
1-Deoxy-D-xylulose-5-phosphate (DXP) isomeroreductase catalyzes the isomerization and reduced nicotinamide adenine dinucleotide phosphate- (NADPH-) dependent reduction of DXP to generate 2-C-methylerythritol 4-phosphate (MEP) in the first committed step of the MEP pathway of isoprenoid biosynthesis. We have cloned the gene encoding the Mycobacterium tuberculosis DXP isomeroreductase, expressed the protein in Escherichia coli, and purified the enzyme to homogeneity using conventional column chromatography methods. DXP isomeroreductase is a metal ion-activated enzyme displaying superior specificity for Co(2+), good specificity for Mn(2+), and poor specificity for Mg(2+). Although NADPH is preferred over reduced nicotinamide adenine dinucleotide (NADH) about 100-fold as evaluated by the relative k(cat)/K(m) values, the maximum turnover numbers are similar, suggesting that the 2'-phosphate of NADPH contributes predominantly to binding and not to catalysis. While k(cat) was independent of pH in the region 6.0 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