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 pH dependence of the kinetic parameters V, V/KNADH, and V/KH2O2 has been determined for the flavoenzyme NADH peroxidase. Both V/KNADH and V/KH2O2 decrease as groups exhibiting pK's of 9.2 and 9.9, respectively, are deprotonated. The V profile decreases by a factor of 5 as a group exhibiting a pK of 7.2 is deprotonated. Primary deuterium kinetic isotope effects on NADH oxidation are observed on V only, and the magnitude of DV is independent of H2O2 concentration at pH 7.5. DV/KNADH is pH independent and equal to 1.0 between pH 6 and pH 9.5, but DV is pH dependent, decreasing from a value of 7.2 at pH 5.5 to 1.9 at pH 9.5. The shape of the DV versus pH profile parallels that observed in the V profile and yields a similar pK of 6.6 for the group whose deprotonation decreases DV. Solvent kinetic isotope effects obtained with NADH or reduced nicotinamide hypoxanthine dinucleotide as the variable substrate are observed on V only, while equivalent solvent kinetic isotope effects on V and V/K are observed when H2O2 is used as the variable substrate. In all cases linear proton inventories are observed. Primary deuterium kinetic isotope effects on V for NADH oxidation decrease as the solvent isotopic composition is changed from H2O to D2O. These data are consistent with a change in the rate-limiting step from a step in the reductive half-reaction at low pH to a step in the oxidative half-reaction at high pH. Analysis of the multiple kinetic isotope effect data suggests that at high D2O concentrations the rate of a single proton transfer step in the oxidative half-reaction is slowed. These data are used to propose a chemical mechanism involving the pH-dependent protonation of a flavin hydroxide anion, following flavin peroxide bond cleavage.  相似文献   

2.
Glucose oxidase from the mold Aspergillus niger (EC 1.1.3.4) oxidizes beta-D-glucose with a wide variety of oxidizing substrates. The substrates were divided into three main groups: molecular oxygen, quinones, and one-electron acceptors. The kinetic and chemical mechanism of action for each group of substrates was examined in turn with a wide variety of kinetic methods and by means of molecular modeling of enzyme-substrate complexes. There are two proposed mechanisms for the reductive half-reaction: hydride abstraction and nucleophilic attack followed by deprotonation. The former mechanism appears plausible; here, beta-D-glucose is oxidized to glucono-delta-lactone by a concerted transfer of a proton from its C1-hydroxyl to a basic group on the enzyme (His516) and a direct hydride transfer from its C1 position to the N5 position in FAD. The oxidative half-reaction proceeds via one- or two-electron transfer mechanisms, depending on the type of the oxidizing substrate. The active site of the enzyme contains, in addition to FAD, three amino acid side chains that are intimately involved in catalysis: His516 with a pK(a)=6.9, and Glu412 with pK(a)=3.4 which is hydrogen bonded to His559, with pK(a)>8. The protonation of each of these residues has a strong influence on all rate constants in the catalytic mechanism.  相似文献   

3.
Buckman J  Miller SM 《Biochemistry》2000,39(34):10532-10541
EBP1-catalyzed reduction of alpha,beta-unsaturated ketones and aldehydes is proposed to proceed via transfer of hydride from the flavin to the beta-position of the olefinic bond, concomitant with or followed by uptake of a proton at the alpha-position. Structural analysis suggests that this proton is donated from Tyr206, and, hence, a protein was constructed in which it was replaced by phenylalanine. The mutation results in a slightly less stable protein than the wild type that nevertheless retains the fundamental flavin and phenol binding properties of EBP1 characterized previously. The pH profile for binding of phenol was characterized over the pH range 6.5-9.5 and was found to be simpler than that for the wild-type enzyme. Most importantly, a pK(a) of 8.7 that is perturbed to 9.4 upon binding of phenol to the wild-type enzyme is missing in the mutant, allowing assignment of this pK(a) to the Y206 hydroxyl group. Additionally, the pK(a) of phenol is further lowered from its value of 10.0 in solution to approximately 6.4 in the active site of the mutant, as compared to 7.1 in the wild type. Together, these perturbations lead to an increase of approximately 35-fold in the binding affinity of the mutant for phenol at high pH relative to the affinity of the wild-type enzyme. As expected, the mutation has little effect on the reductive half-reaction, in which a hydride equivalent is transferred from NADPH to the flavin. In contrast, the reduction of trans-2-hexenal by the reduced enzyme is significantly affected. The results indicate formation of a previously unobserved charge-transfer (CT) complex following formation of the Michaelis complex between substrate and reduced enzyme and preceding reduction of the substrate, which occurs at a greatly reduced rate (>/=440-fold) relative to wild type. Thus, while the oxidative half-reaction with wild-type enzyme is limited by the rate of formation of the CT complex, it is the chemical step that is rate-limiting in the reaction with EBP1:Y206F, consistent with the role of this residue as a general acid.  相似文献   

4.
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   

5.
Patel MP  Blanchard JS 《Biochemistry》2001,40(17):5119-5126
The recent identification of the enzyme in Mycobacterium tuberculosis that catalyzes the NADPH-dependent reduction of the unique low molecular weight disulfide mycothione, mycothione reductase, has led us to examine the mechanism of catalysis in greater detail. The pH dependence of the kinetic parameters V and V/K for NADPH, NADH, and an active analogue of mycothione disulfide, des-myo-inositol mycothione disulfide, has been determined. An analysis of the pH profiles has allowed the tentative assignment of catalytically significant residues crucial to the mechanism of disulfide reduction, namely, the His444-Glu449 ion pair and Cys39. Solvent kinetic isotope effects were observed on V and V/K(DIMSSM), yielding values of 1.7 +/- 0.2 and 1.4 +/- 0.2, respectively, but not on V/K(NADPH). Proton inventory studies (V versus mole fraction of D(2)O) were linear, indicative of a single proton transfer in a solvent isotopically sensitive step. Steady-state primary deuterium kinetic isotope effects on V have been determined using NADPH and NADH, yielding values of 1.27 +/- 0.03 and 1.66 +/- 0.14, respectively. The pre-steady-state primary deuterium kinetic isotope effect on enzyme reduction has values of 1.82 +/- 0.04 and 1.59 +/- 0.06 for NADPH and NADH, respectively. The steady-state primary deuterium kinetic isotope effect using NADH coincide with that obtained under single turnover conditions, suggesting the complete expression of the intrinsic primary kinetic isotope effect. Rapid reaction studies on the reductive half-reaction using NADPH and NADH yielded maximal rates of 129 +/- 2 and 20 +/- 1 s(-1), respectively, while similar studies of the oxidation of the two-electron reduced enzyme by mycothiol disulfide yielded a maximum rate of 190 +/- 10 s(-1). These data suggest a unique flavoprotein disulfide mechanism in which the rate of the oxidative half-reaction is slightly faster than the rate of the reductive half-reaction.  相似文献   

6.
Argyrou A  Blanchard JS 《Biochemistry》2001,40(38):11353-11363
The gene encoding dihydrolipoamide dehydrogenase from Mycobacterium tuberculosis, Rv0462, was expressed in Escherichia coli and the protein purified to homogeneity. The 49 kDa polypeptide forms a homodimer containing one tightly bound molecule of FAD/monomer. The results of steady-state kinetic analyses using several reduced pyridine nucleotide analogs and a variety of electron acceptors, and the ability of the enzyme to catalyze the transhydrogenation of NADH and thio-NAD(+) in the absence of D,L-lipoamide, demonstrated that the enzyme uses a ping-pong kinetic mechanism. Primary deuterium kinetic isotope effects on V and V/K at pH 7.5 using NADH deuterated at the C(4)-proS position of the nicotinamide ring are small [(D)(V/K)(NADH) = 1.12 +/- 0.15, (D)V(app) = 1.05 +/- 0.07] when D,L-lipoamide is the oxidant but large and equivalent [(D)(V/K)(NADH) = (D)V = 2.95 +/- 0.03] when 5-hydroxy-1,4-naphthoquinone is the oxidant. Solvent deuterium kinetic isotope effects at pH 5.8, using APADH as the reductant, are inverse with (D)(V/K)(APADH) = 0.73 +/- 0.03, (D)(V/K)(Lip(S))2 = 0.77 +/- 0.03, and (D)V(app) = 0.77 +/- 0.01. Solvent deuterium kinetic isotope effects with 4,4-dithiopyridine (DTP), the 4-thiopyridone product of which requires no protonation, are also inverse with (D)(V/K)(APADH) = 0.75 +/- 0.06, (D)(V/K)(DTP) = 0.71 +/- 0.02, and (D)V(app) = 0.56 +/- 0.15. All proton inventories were linear, indicating that a single proton is being transferred in the solvent isotopically sensitive step. Taken together, these results suggest that (1) the reductive half-reaction (hydride transfer from NADH to FAD) is rate limiting when a quinone is the oxidant, and (2) deprotonation of enzymic thiols, most likely Cys(46) and Cys(41), limits the reductive and oxidative half-reactions, respectively, when D,L-lipoamide is the oxidant.  相似文献   

7.
The pH dependence of the initial reaction rate catalyzed by the isolated bovine heart ubiquinol-cytochrome c reductase (bc1 complex) varying decylbenzoquinol (DBH) and decylbenzoquinone (DB) concentrations was determined. The affinity for DBH was increased threefold by the protonation of a group with pKa = 5.7 +/- 0.2, while the inhibition constant (Ki) for DB decreased 22 and 2.8 times when groups with pKa = 5.2 +/- 0.6 and 7.7 +/- 0.2, respectively, were protonated. This suggests stabilization of the protonated form of the acidic group by DBH binding. Initial rates were best fitted to a kinetic model involving three protonatable groups. The protonation of the pKa approximately 5.7 group blocked catalysis, indicating its role in proton transfer. The kinetic model assumed that the deprotonation of two groups (pKa values of 7.5 +/- 0.03 and approximately 9.2) decreases the catalytic rate by diminishing the redox potential of the iron-sulfur (Fe-S) cluster. The protonation of the pKa approximately 7.5 group also decreased the reaction rate by 80-86%, suggesting its role as acceptor of a proton from ubiquinol. The lack of effect on the Km for DBH when the pKa 7.5-7.7 group is deprotonated suggests that hydrogen bonding to this residue is not the main factor that determines substrate binding to the Qo site. The possible relationship of the pKa 5.2-5.7 and pKa 7.5-7.7 groups with Glu272 of cytochrome b and His161 of the Fe-S protein is discussed.  相似文献   

8.
2-Methyl-3-hydroxypyridine-5-carboxylic acid (MHPC) oxygenase (MHPCO) is a flavoprotein that catalyzes the oxygenation of MHPC to form alpha-(N-acetylaminomethylene)-succinic acid. Although formally similar to the oxygenation reactions catalyzed by phenol hydroxylases, MHPCO catalyzes the oxygenation of a pyridyl derivative rather than a simple phenol. Therefore, in this study, the mechanism of the reaction was investigated by replacing the natural cofactor FAD with FAD analogues having various substituents (-Cl, -CN, -NH(2), -OCH(3)) at the C8-position of the isoalloxazine. Thermodynamic and catalytic properties of the reconstituted enzyme were investigated and found to be similar to those of the native enzyme, validating that these FAD analogues are reasonable to be used as mechanistic probes. Dissociation constants for the binding of MHPC or the substrate analogue 5-hydroxynicotinate (5HN) to the reconstituted enzymes indicate that the reconstituted enzymes bind well with ligands. Redox potential values of the reconstituted enzymes were measured and found to be more positive than the values of free FAD analogues, which correlated well with the electronic effects of the 8-substituents. Studies of the reductive half-reaction of MHPCO have shown that the rates of flavin reduction by NADH could be described as a parabolic relationship with the redox potential values of the reconstituted enzymes, which is consistent with the Marcus electron transfer theory. Studies of the oxidative half-reaction of MHPCO revealed that the rate of hydroxylation depended upon the different analogues employed. The rate constants for the hydroxylation step correlated with the calculated pK(a) values of the 8-substituted C(4a)-hydroxyflavin intermediates, which are the leaving groups in the oxygen transfer step. It was observed that the rates of hydroxylation were greater when the pK(a) values of C(4a)-hydroxyflavins were lower. Although these results are not as dramatic as those from analogous studies with parahydroxybenzoate hydroxylase (Ortiz-Maldonado et al., (1999) Biochemistry 38, 8124-8137), they are consistent with the model that the oxygenation reaction of MHPCO occurs via an electrophilic aromatic substitution mechanism analogous to the mechanisms for parahydroxybenzoate and phenol hydroxylases.  相似文献   

9.
X Wu  B A Palfey  V V Mossine  V M Monnier 《Biochemistry》2001,40(43):12886-12895
Amadoriase is a flavoenzyme that catalyzes the oxidative deglycation of Amadori products (fructosyl amino acids or aliphatic amines) to yield free amine, glucosone, and hydrogen peroxide. The mechanism of action of amadoriase I from Aspergillus sp. has been investigated by stopped-flow kinetic studies using fructosyl propylamine and O(2) as substrates in 10 mM Tris HCl, pH 7.9, 4 degrees C. Using both substrate analogues and fast kinetic techniques, the active configuration of the substrate was found to be the beta-pyranose form. Stopped-flow studies showed that the reductive half-reaction is triphasic and generates intermediates that absorb at long wavelengths and is consistent either with (i) the reaction of the substrate with the flavin followed by iminium deprotonation or hydrolysis and then product release or with (ii) the formation of flavin reduction intermediates (carbanion equivalents or adducts), followed by product release. The rate of product release after flavin reduction is lower than the aerobic turnover rate, 14.4 s(-1), suggesting that it is not involved in the catalytic cycle and that reoxidation of the reduced enzyme occurs in the E(red)-product complex. In the oxidative half-reaction, the reduced flavin is oxidized by O(2) in a single phase. The observed rate constant has a linear dependence on oxygen concentration, giving a bimolecular rate constant of 4.9 x 10(4) M(-1) s(-1) in the absence of product, and 3.6 x 10(4) M(-1) s(-1) when the product is bound. The redox potentials of amadoriase have been measured at pH 7.0, 25 degrees, giving values of +48 and -52 mV for the oxidized enzyme/anionic semiquinone and anionic semiquinone/reduced enzyme couples, respectively.  相似文献   

10.
The pH variation of the kinetic parameters for the oxidative decarboxylation of L-malate and decarboxylation of oxalacetate catalyzed by malic enzyme has been used to gain information on the catalytic mechanism of this enzyme. With Mn2+ as the activator, an active-site residue with a pK of 5.4 must be protonated for oxalacetate decarboxylation and ionized for the oxidative decarboxylation of L-malate. With Mg2+ as the metal, this pK is 6, and, at high pH, V/K for L-malate decreases when groups with pKs of 7.8 and 9 are deprotonated. The group at 7.8 is a neutral acid (thought to be water coordinated to Mg2+), while the group at 9 is a cationic acid such as lysine. The V profile for reaction of malate shows these pKs displaced outward by 1.4 pH units, since the rate-limiting step is normally TPNH release, and the chemical reaction, which is pH sensitive, is 25 times faster. TPN binding is decreased by ionization of a group with pK 9.3 or protonation of a group with pK 5.3. The pH variation of the Km for Mg shows that protonation of a group with pK 8.7 (possibly SH) decreases metal binding in the presence of malate by a factor of 1400, and in the absence of malate by a factor of 20. A catalytic mechanism is proposed in which hydride transfer is accompanied by transfer of a proton to the group with pK 5.4-6, and enolpyruvate is protonated by water coordinated to the Mg2+ (pK 7.8) after decarboxylation and release of CO2.  相似文献   

11.
The structure of the flavin-dependent alkanesulfonate monooxygenase (SsuD) exists as a TIM-barrel structure with an insertion region located over the active site that contains a conserved arginine (Arg297) residue present in all SsuD homologues. Substitution of Arg297 with alanine (R297A SsuD) or lysine (R297K SsuD) was performed to determine the functional role of this conserved residue in SsuD catalysis. While the more conservative R297K SsuD possessed a lower k(cat)/K(m) value (0.04 ± 0.01 μM(-1) min(-1)) relative to wild-type (1.17 ± 0.22 μM(-1) min(-1)), there was no activity observed with the R297A SsuD variant. Each of the arginine variants had similar K(d) values for flavin binding as wild-type SsuD (0.32 ± 0.15 μM), but there was no measurable binding of octanesulfonate. The low levels of activity for the R297A and R297K SsuD variants correlated with the absence of any detectable C4a-(peroxy)flavin formation in stopped-flow kinetic studies. Single-turnover experiments were performed in the presence of SsuE to evaluate both the reductive and oxidative half-reaction. With wild-type SsuD a lag phase is observed following the reductive half-reaction by SsuE that represents flavin transfer or conformational changes associated with the binding of substrates. Evaluation of the Arg297 SsuD variants in the presence of SsuE showed no lag phase following reduction by SsuE, and the flavin was oxidized immediately following the reductive half-reaction. These results corresponded with a lack of detectable changes in the proteolytic susceptibility of R297A and R297K SsuD in the presence of reduced flavin and/or octanesulfonate, signifying the absence of a conformational change in these variants with the substitution of Arg297.  相似文献   

12.
Blazyk JL  Lippard SJ 《Biochemistry》2002,41(52):15780-15794
Soluble methane monooxygenase (sMMO) from Methylococcus capsulatus (Bath) catalyzes the selective oxidation of methane to methanol, the first step in the primary catabolic pathway of methanotrophic bacteria. A reductase (MMOR) mediates electron transfer from NADH through its FAD and [2Fe-2S] cofactors to the dinuclear non-heme iron sites housed in a hydroxylase (MMOH). The structurally distinct [2Fe-2S], FAD, and NADH binding domains of MMOR facilitated division of the protein into its functional ferredoxin (MMOR-Fd) and FAD/NADH (MMOR-FAD) component domains. The 10.9 kDa MMOR-Fd (MMOR residues 1-98) and 27.6 kDa MMOR-FAD (MMOR residues 99-348) were expressed and purified from recombinant Escherichia coli systems. The Fd and FAD domains have absorbance spectral features identical to those of the [2Fe-2S] and flavin components, respectively, of MMOR. Redox potentials, determined by reductive titrations that included indicator dyes, for the [2Fe-2S] and FAD cofactors in the domains are as follows: -205.2 +/- 1.3 mV for [2Fe-2S](ox/red), -172.4 +/- 2.0 mV for FAD(ox/sq), and -266.4 +/- 3.5 mV for FAD(sq/hq). Kinetic and spectral properties of intermediates observed in the reaction of oxidized MMOR-FAD (FAD(ox)) with NADH at 4 degrees C were established with stopped-flow UV-visible spectroscopy. Analysis of the influence of pH on MMOR-FAD optical spectra, redox potentials, and NADH reaction kinetics afforded pK(a) values for the semiquinone (FAD(sq)) and hydroquinone (FAD(hq)) MMOR-FAD species and two protonatable groups near the flavin cofactor. Electron transfer from MMOR-FAD(hq) to oxidized MMOR-Fd is extremely slow (k = 1500 M(-1) s(-1) at 25 degrees C, compared to 90 s(-1) at 4 degrees C for internal electron transfer between cofactors in MMOR), indicating that cofactor proximity is essential for efficient interdomain electron transfer.  相似文献   

13.
The maximum velocity of the malic enzyme (L-malate: NADP+ oxidoreductase (oxaloacetate-decarboxylating), EC 1.1.1.40) reductive carboxylation of pyruvate and V/KCO2 are pH-independent from pH 5.5 to pH 8.5. V/K for pyruvate exhibits pK values values of 6.50 +/- 0.25 and 7.25 +/- 0.25. These data suggest that the binding of pyruvate locks the protonation state of enzyme. In addition, the pK values are within experimental error identical for the pH dependence of V/Kmalate and V/Kpyruvate. Thus, the catalytic groups appear to have reverse protonation states in the two reaction directions. The ratio of (V/Kmalate)/(V/Kpyruvate) is 100, suggesting that the protonation state of enzyme is optimum in the malate oxidative decarboxylation direction. Thus, the group with a pK of about 6 is unprotonated and the group with a pK of 7.5 is protonated for malate decarboxylation, and the opposite is true for pyruvate reductive carboxylation.  相似文献   

14.
1. The reaction of 4-iodoacetamidosalicylate with bovine liver glutamate dehydrogenase is dependent on pH. The pH-activity curve is bell-shaped and can be described by apparent pK values of 7.8+/-0.2 and 9.1+/-0.2. 2. Enzyme in which lysine-126 has been modified by 4-iodoacetamidosalicylate has unaltered sedimentation characteristics except when measured in the presence of GTP and NADH. 3. GTP binding to the inhibited enzyme is unaltered. However, GTP can no longer promote the binding of a second molecule of NADH, since this is already bound to the inhibited enzyme without GTP. 4. The equilibrium binding of ADP, GTP, NAD-sulphite and NADH (when measured at low concentrations) was largely unchanged by modification. 5. The number of binding sites for 2-oxoglutarate to the enzyme-NADH complex were decreased by 60% in an enzyme that has been inhibited by 70%.  相似文献   

15.
Trimmer EE  Ballou DP  Matthews RG 《Biochemistry》2001,40(21):6205-6215
The flavoprotein methylenetetrahydrofolate reductase (MTHFR) from Escherichia coli catalyzes the reduction of 5,10-methylenetetrahydrofolate (CH(2)-H(4)folate) to 5-methyltetrahydrofolate (CH(3)-H(4)folate) using NADH as the source of reducing equivalents. The enzyme also catalyzes the transfer of reducing equivalents from NADH or CH(3)-H(4)folate to menadione, an artificial electron acceptor. Here, we have determined the midpoint potential of the enzyme-bound flavin to be -237 mV. We have examined the individual reductive and oxidative half-reactions constituting the enzyme's activities. In an anaerobic stopped-flow spectrophotometer, we have measured the rate constants of flavin reduction and oxidation occurring in each half-reaction and have compared these with the observed catalytic turnover numbers measured under steady-state conditions. We have shown that, in all cases, the half-reactions proceed at rates sufficiently fast to account for overall turnover, establishing that the enzyme is kinetically competent to catalyze these oxidoreductions by a ping-pong Bi-Bi mechanism. Reoxidation of the reduced flavin by CH(2)-H(4)folate is substantially rate limiting in the physiological NADH-CH(2)-H(4)folate oxidoreductase reaction. In the NADH-menadione oxidoreductase reaction, the reduction of the flavin by NADH is rate limiting as is the reduction of flavin by CH(3)-H(4)folate in the CH(3)-H(4)folate-menadione oxidoreductase reaction. We conclude that studies of individual half-reactions catalyzed by E. coli MTHFR may be used to probe mechanistic questions relevant to the overall oxidoreductase reactions.  相似文献   

16.
The crystal structure of aryl-alcohol oxidase (AAO), a flavoenzyme involved in lignin degradation, reveals two active-site histidines, whose role in the two enzyme half-reactions was investigated. The redox state of flavin during turnover of the variants obtained show a stronger histidine involvement in the reductive than in the oxidative half-reaction. This was confirmed by the k(cat)/K(m(Al)) and reduction constants that are 2-3 orders of magnitude decreased for the His546 variants and up to 5 orders for the His502 variants, while the corresponding O(2) constants only decreased up to 1 order of magnitude. These results confirm His502 as the catalytic base in the AAO reductive half-reaction. The solvent kinetic isotope effect (KIE) revealed that hydroxyl proton abstraction is partially limiting the reaction, while the α-deuterated alcohol KIE showed a stereoselective hydride transfer. Concerning the oxidative half-reaction, directed mutagenesis and computational simulations indicate that only His502 is involved. Quantum mechanical/molecular mechanical (QM/MM) reveals an initial partial electron transfer from the reduced FADH(-) to O(2), without formation of a flavin-hydroperoxide intermediate. Reaction follows with a nearly barrierless His502H(+) proton transfer that decreases the triplet/singlet gap. Spin inversion and second electron transfer, concomitant with a slower proton transfer from flavin N5, yields H(2)O(2). No solvent KIE was found for O(2) reduction confirming that the His502 proton transfer does not limit the oxidative half-reaction. However, the small KIE on k(cat)/K(m(Ox)), during steady-state oxidation of α-deuterated alcohol, suggests that the second proton transfer from N5H is partially limiting, as predicted by the QM/MM simulations.  相似文献   

17.
K K Wong  J S Blanchard 《Biochemistry》1989,28(8):3586-3590
Human erythrocyte glutathione reductase catalyzes the pyridine nucleotide dependent reduction of oxidized glutathione (GSSG). The pH dependence of the kinetic parameters V and V/K for three reduced pyridine nucleotide substrates, the Ki's for three competitive inhibitors (versus NADPH), and the temperature dependence of the V pH profile have been determined. Below pH 8, V and V/K for NADPH, 2',3'-cyclic-NADPH, and NADH are pH independent. In the basic pH region, both V and V/K for the three substrates are pH dependent. All three of the V profiles decrease with increasing pH as a group with a pKa of approximately 9.2 is titrated. The V/K profiles for NADPH, 2',3'-cyclic-NADPH, and NADH decrease at high pH as a group with a pKa of greater than 9.8, 8.9, and 8.8, respectively, is deprotonated. The Ki's for ATP-ribose and 2',5'-ADP are pH independent below pH 8 but increase in the basic region as a group with a pKa of about 8.8 and 8.5, respectively, is deprotonated. The Ki of AADP is pH independent between pH 6 and 9. These studies suggest that binding interactions between the 2'-phosphate of NADPH and the enzyme are predominately nonionic. The temperature dependence of the pK observed in all V pH profiles allows the calculation of an enthalpy of ionization of 3.2 kcal/mol for this group. The high pK and low enthalpy of ionization suggest that the protonation state of the His-467'-Glu-472' ion pair observed in the structure of human erythrocyte glutathione reductase influences proton-transfer steps occurring in the oxidative half-reaction.  相似文献   

18.
Styrene monooxygenase (SMO) from Pseudomonas putida S12 is a two-component flavoenzyme composed of the NADH-specific flavin reductase, SMOB, and FAD-specific styrene epoxidase, SMOA. Here, we report the cloning, and expression of native and histidine-tagged versions of SMOA and SMOB and studies of the flavin transfer and styrene oxygenation reactions. In the reductive half-reaction, SMOB catalyzes the two-electron reduction of FAD with a turnover number of 3200 s(-1). Single turnover studies of the reaction of reduced SMOA with substrates indicate the formation of a stable oxygen intermediate with the absorbance characteristics of a flavin hydroperoxide. Based on the results of numerical simulations of the steady-state mechanism of SMO, we find that the observed coupling of NADH and styrene oxidation can be best explained by a model, which includes both the direct transfer and passive diffusion of reduced FAD from SMOB to SMOA.  相似文献   

19.
Thermus thermophilus NADH oxidase (NOX) activity exhibits a bell-shaped pH-dependency with the maximal rate at pH 5.2 and marked inhibition at lower pH. The first pH transition, from pH 7.2 to pH 5.2, results in more than a 2-fold activity increase with protonation of a group with pKa=6.1+/-0.1. The difference in fluorescence of the free and enzyme-bound flavin strongly indicates that the increase in enzyme activity in a pH-dependent manner is related to a protein-cofactor interaction. Only one amino acid residue, His75, has an intrinsic pKa approximately 6.0 and is localized in proximity (<10 A) to N5-N10 of the isoalloxazine ring and, therefore, is able to participate in such an interaction. Solvent acidification leads to the second pH transition from pH 5.2 to 2.0 that results in complete inhibition of the enzyme with protonation of a group with an apparent pKa=4.0+/-0.1. Inactivation of NOX activity at low pH is not caused by large conformational changes in the quaternary structure as judged by intrinsic viscosity and sedimentation velocity experiments. NOX exists as a dimer even as an apoprotein at acidic conditions. There is a strong coupling between the fluorescence of the enzyme-bound flavin and the intrinsic tryptophans, as demonstrated by energy transfer between Trp47 and the isoalloxazine ring of flavin adenine dinucleotide (FAD). The pH-induced changes in intrinsic tryptophan and FAD fluorescence indicate that inhibition of the FAD-binding enzyme at low pH is related to dissociation of the flavin cofactor, due to protonation of its adenine moiety.  相似文献   

20.
Tryptophan 2-monooxygenase (TMO) from Pseudomonas savastanoi catalyzes the oxidative decarboxylation of l-tryptophan during the biosynthesis of indoleacetic acid. Structurally and mechanistically, the enzyme is a member of the family of l-amino acid oxidases. Deuterium and 15N kinetic isotope effects were used to probe the chemical mechanism of l-alanine oxidation by TMO. The primary deuterium kinetic isotope effect was pH independent over the pH range 6.5-10, with an average value of 6.0 +/- 0.5, consistent with this being the intrinsic value. The deuterium isotope effect on the rate constant for flavin reduction by alanine was 6.3 +/- 0.9; no intermediate flavin species were observed during flavin reduction. The kcat/Kala value was 1.0145 +/- 0.0007 at pH 8. NMR analyses gave an equilibrium 15N isotope effect for deprotonation of the alanine amino group of 1.0233 +/- 0.0004, allowing calculation of the 15N isotope effect on the CH bond cleavage step of 0.9917 +/- 0.0006. The results are consistent with TMO oxidation of alanine occurring through a hydride transfer mechanism.  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