首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 15 毫秒
1.
The catalytic mechanism of the phosphoglycerate dehydrogenase reaction in both directions was investigated by studying: (a) pre-steady state transients in reduced coenzyme appearance or disappearance or disappearance and in protein fluorescence; (b) deuterium isotope effects on the transients and on the steady state reactions; and (c) the partial reaction between the enzyme-NADH complex and hydroxypyruvate-P. These studies led to the scheme below for the ternary complex interconversion. E1-NADH-hydroxypyruvate-P(1)equilibriumE2-NADH-hydroxypyruvate-P(2)equilibriumE3-NADH-hydroxypyruvate-P + H+(3)equilibriumE3-NAD+-3-phosphoglycerate(4)equilibriumE4-NAD+-3-phosphoglycerate Steps 1,2, and 4 are ternary complex isomerizations. Step 3 is the hydride transfer. Under steady state conditions isomerization 2 is the rate-determining step in the direction of hydroxypyruvate-P reduction at higher pH values. At lower pH values, the hydride transfer step is also partially rate-determining. The rate-determining step in the direction of 3-phosphoglycerate oxidation occurs subsequent to the hydride transfer step at higher pH values. At lower pH values the rate is determined by both isomerization 4 and the hydride transfer step. Isomerizations 1, 2, and 4 were inhibited by serine, an allosteric inhibitor, indicating that the inactive conformation of the enzyme is incapable of performing any of the steps of the ternary complex interconversion. Phosphoglycerate dehydrogenase corresponds to a V-type allosteric enzyme. When the enzyme-NADH complex was mixed with hydroxypyruvate-P at pH 8.5, a rapid quenching of enzymebound NADH fluorescence occurred. This process was studied under pseudo-first order conditions and shown to be the result of hydroxypyruvate-P binding.  相似文献   

2.
The reduction of the melilotate hydroxylase . 2-OH-phenyl propionate complex by NADH and reduced 3-acetyl pyridine adenine dinucleotide (AcPyNADH) has been investigated using steady state kinetic and rapid reaction techniques. Reduction by NADH appeared to involve only one charge-transfer-type intermediate (between reduced enzyme and NAD) as previously described (Strickland, S., and Massey, V. (1973) J. Biol. Chem. 248, 2953-2962). Reduction by AcPyNADH was shown to involve two charge-transfer-type intermediates. The first was between oxidized enzyme and AcPyNADH and the second was between reduced enzyme and AcPyNAD. Reaction of AcPyNADH with oxidized enzyme . 2-OH-phenyl propionate complex to form the first charge-transfer complex reached equilibrium within the mixing time of the stopped flow apparatus (5 ms). Subsequent steps in the reaction appeared to be first order and were independent of the AcPyNADH concentration. An 8-fold deuterium isotope effect on the step involving flavin reduction was found when reduced 3-acetyl[4A-2H]pyridine adenine dinucleotide (AcPyNADD) was used as the reductant. Analysis of the rapid reaction results for the reaction of oxidized pyridine nucleotide with reduced enzyme . 2-OH-phenyl propionate complex indicated the presence of two forms of reduced enzyme (in equilibrium) of which only one form was capable of reacting with the oxidized pyridine nucleotide. Based on the rapid reaction data, a mechanism for the reduction half-reaction is proposed. The turnover number calculated from this mechanism is in good agreement with that determined from the steady state data.  相似文献   

3.
A kinetic scheme is presented for Escherichia coli dihydrofolate reductase that predicts steady-state kinetic parameters and full time course kinetics under a variety of substrate concentrations and pHs. This scheme was derived from measuring association and dissociation rate constants and pre-steady-state transients by using stopped-flow fluorescence and absorbance spectroscopy. The binding kinetics suggest that during steady-state turnover product dissociation follows a specific, preferred pathway in which tetrahydrofolate (H4F) dissociation occurs after NADPH replaces NADP+ in the ternary complex. This step, H4F dissociation from the E X NADPH X H4F ternary complex, is proposed to be the rate-limiting step for steady-state turnover at low pH because koff = VM. The rate constant for hydride transfer from NADPH to dihydrofolate (H2F), measured by pre-steady-state transients, has a deuterium isotope effect of 3 and is rapid, khyd = 950 s-1, essentially irreversible, Keq = 1700, and pH dependent, pKa = 6.5, reflecting ionization of a single group in the active site. This scheme accounts for the apparent pKa = 8.4 observed in the steady state as due to a change in the rate-determining step from product release at low pH to hydride transfer above pH 8.4. This kinetic scheme is a necessary background to analyze the effects of single amino acid substitutions on individual rate constants.  相似文献   

4.
C T Grubmeyer  K W Chu  S Insinga 《Biochemistry》1987,26(12):3369-3373
Salmonella typhimurium histidinol dehydrogenase produces histidine from the amino alcohol histidinol by two sequential NAD-linked oxidations which form and oxidize a stable enzyme-bound histidinaldehyde intermediate. The enzyme was found to catalyze the exchange of 3H between histidinol and [4(R)-3H]NADH and between NAD and [4(S)-3H]NADH. The latter reaction proceeded at rates greater than kcat for the net reaction and was about 3-fold faster than the former. Histidine did not support an NAD/NADH exchange, demonstrating kinetic irreversibility in the second half-reaction. Specific activity measurements on [3H]histidinol produced during the histidinol/NADH exchange reaction showed that only a single hydrogen was exchanged between the two reactants, demonstrating that under the conditions employed this exchange reaction arises only from the reversal of the alcohol dehydrogenase step and not the aldehyde dehydrogenase reaction. The kinetics of the NAD/NADH exchange reaction demonstrated a hyperbolic dependence on the concentration of NAD and NADH when the two were present in a 1:2 molar ratio. The histidinol/NADH exchange showed severe inhibition by high NAD and NADH under the same conditions, indicating that histidinol cannot dissociate directly from the ternary enzyme-NAD-histidinol complex; in other words, the binding of substrate is ordered with histidinol leading. Binding studies indicated that [3H]histidinol bound to 1.7 sites on the dimeric enzyme (0.85 site/monomer) with a KD of 10 microM. No binding of [3H]NAD or [3H]NADH was detected. The nucleotides could, however, displace histidinol dehydrogenase from Cibacron Blue-agarose.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

5.
l-Lactate oxidase (LOX) belongs to a family of flavin mononucleotide (FMN)-dependent α-hydroxy acid-oxidizing enzymes. Previously, the crystal structure of LOX (pH 8.0) from Aerococcus viridans was solved, revealing that the active site residues are located around the FMN. Here, we solved the crystal structures of the same enzyme at pH 4.5 and its complex with d-lactate at pH 4.5, in an attempt to analyze the intermediate steps. In the complex structure, the d-lactate resides in the substrate-binding site, but interestingly, an active site base, His265, flips far away from the d-lactate, as compared with its conformation in the unbound state at pH 8.0. This movement probably results from the protonation of His265 during the crystallization at pH 4.5, because the same flip is observed in the structure of the unbound state at pH 4.5. Thus, the present structure appears to mimic an intermediate after His265 abstracts a proton from the substrate. The flip of His265 triggers a large structural rearrangement, creating a new hydrogen bonding network between His265-Asp174-Lys221 and, furthermore, brings molecular oxygen in between d-lactate and His265. This mimic of the ternary complex intermediate enzyme-substrate-O2 could explain the reductive half-reaction mechanism to release pyruvate through hydride transfer. In the mechanism of the subsequent oxidative half-reaction, His265 flips back, pushing molecular oxygen into the substrate-binding site as the second substrate, and the reverse reaction takes place to produce hydrogen peroxide. During the reaction, the flip-flop action of His265 has a dual role as an active base/acid to define the major chemical steps. Our proposed reaction mechanism appears to be a common mechanistic strategy for this family of enzymes.  相似文献   

6.
UDP-glucose dehydrogenase catalyzes the NAD+-dependent twofold oxidation of UDP-glucose to give UDP-glucuronic acid. A sequestered aldehyde intermediate is produced in the first oxidation step and a covalently bound thioester is produced in the second oxidation step. This work demonstrates that the Streptococcus pyogenes enzyme incorporates a single solvent-derived oxygen atom during catalysis and probably does not generate an imine intermediate. The reaction of UDP-[6",6"-di-2H]-d-glucose is not accompanied by a primary kinetic isotope effect, indicating that hydride transfer is not rate determining in this reaction. Studies with a mutant of the key active site nucleophile, Cys260Ala, show that it is capable of both reducing the aldehyde intermediate, and oxidizing the hydrated form of the aldehyde intermediate but is incapable of oxidizing UDP-glucose to UDP-glucuronic acid. In the latter case, a ternary Cys260Ala/aldehyde intermediate/NADH complex is presumably formed, but it does not proceed to product as both release and hydration of the bound aldehyde occur slowly. A washout experiment demonstrates that the NADH in this ternary complex is not exchangeable with external NADH, indicating that dissociation only occurs after the addition of a nucleophile to the aldehyde carbonyl. Studies on Thr118Ala show that the value of kcat is reduced 160-fold by this mutation, and that the reaction of UDP-D-[6",6"-di-2H]-glucose is now accompanied by a primary kinetic isotope effect. This indicates that the barriers for the hydride transfer steps have been selectively increased and supports a mechanism in which an ordered water molecule (H-bonded to Thr118) serves as the catalytic base in these steps.  相似文献   

7.
Using the quenched flow technique the mechanism of seryl tRNA synthetase action has been investigated with respect to the presteady state kinetics of individual steps. Under conditions where the strong binding sites of the enzyme are nearly saturated and the steady state turnover number is about 1 s-1, rate constants of four different processes have been determined: steps connected with substrate associations are relatively slow (12 s-1 for the entire process); activation of serine is the rate determining step (about 1.2 s-1 in presence of tRNASer); whereas the transfer of serine onto tRNASer (35 s-1) and the dissociation of seryl tRNASer (70 s-1) are fast. Similar kinetic parameter seem to hold also for the steady state reactions. This conclusion is based on a detailed study of the substrate, product, and Mg2+ concentration dependence of the transfer reaction. The results also indicate that a second serine binding site is operative. Since the transfer of serine from a preformed adenylate complex onto tRNASer is fast, seryl adenylate seems to be a kinetically competent intermediate of the aminoacylation reaction although, of course, alternative mechanisms cannot be excluded.  相似文献   

8.
Enhanced inhibition of thymidylate synthase by methotrexate polyglutamates   总被引:15,自引:0,他引:15  
We have studied the effects of methotrexate (MTX-Glu1) and the polyglutamate derivatives of methotrexate (MTXPGs) with 2, 3, 4, and 5 glutamyl residues on the catalytic activity of thymidylate synthase purified from MCF-7 human breast cancer cells and on the kinetics of the ternary complex formation by 5-fluoro-2'-deoxyuridine 5'-monophosphate, folate cofactor, and thymidylate synthase. MTX-Glu1 exhibited uncompetitive inhibition of thymidylate synthase when reaction kinetics were analyzed by either double reciprocal plots or a computerized mathematical model based on nonlinear least-squares curve fitting. The Ki for MTX-Glu1 inhibition was 13 microM and the I50 was 22 microM, irrespective of the degree of polyglutamation of the folate. In contrast, the polyglutamated derivatives of MTX all acted as noncompetitive inhibitors. The MTXPGs had 75-300-fold greater potency than MTX-Glu1 as inhibitors of thymidylate synthase catalytic activity, with Ki values from 0.17 to 0.047 microM for MTX-Glu2 to MTX-Glu5, respectively. Neither MTX-Glu1 nor MTXPGs promoted the formation of a charcoal-stable ternary complex with thymidylate synthase and 5-fluoro-2'-deoxyuridine 5'-monophosphate. CH2-H4PteGlu5 (where PteGlu represents pteroylglutamic acid) was found to be 40-fold more potent than CH2-H4PteGlu1 in participating in the formation of a ternary complex, and 10 microM MTX-Glu5 significantly inhibited the formation of a ternary complex containing this folate as cofactor. The inhibition was determined to be due to a reduction in the kon. The potency of this inhibition was markedly greater in the presence of CH2-H4PteGlu1 as compared to CH2-H4PteGlu5. This finding suggests that the degree of interference with complex formation in intact cells would depend on the state of polyglutamation of available folate cofactor. Ternary complex formation with H2PteGlu5 as the folate cofactor was also investigated, and a 50% reduction in complex formation was found in the presence of a 2 microM concentration of MTX-Glu5. These findings have significant implications regarding the mechanism of action of MTX-Glu1 and contribute to an understanding of the complex interactions of MTX-Glu1 and 5-fluorouracil.  相似文献   

9.
Kinetic methods for studying the reactions of the “general” fatty acyl CoA dehydrogenase under three sets of substrate and enzyme concentration conditions have been developed. The reaction of butyryl-CoA and electron transfer flavoprotein (ETF) can be studied either under steady-state conditions with enzyme at catalytic concentration or under single-turnover conditions with enzyme in excess. Under the latter conditions, acyl-CoA dehydrogenase acts both as a catalyst and an ultimate electron-transfer acceptor. The reductive half-reaction of butyryl-CoA and enzyme can also be studied in a separate kinetic experiment. Comparison of the pH dependences of the rate constants and isotope effects of the steady-state reaction of butyryl-CoA and ETF with the same parameters for the reductive half-reaction is consistent with a mechanism involving transfer of electrons from butyryl-CoA to ETF within a ternary complex. An alternative mechanism in which the reductive half-reaction takes place prior to the binding and reaction of ETF seems unlikely because the pH 8.5 isotope effect on the reductive half-reaction is much larger than that on the complete reaction in spite of the fact that the rates of the reactions are comparable. The pH dependence of the Km for substrate and KI for inhibitor is consistent with a mechanism for transfer of electrons within the ternary complex which involves protonation of the C group of substrates. The protonation labilizes the C-2 proton and base catalysis of the removal of the C-2 proton results in the production of the active enzyme-substrate species, namely the C-2 anion of substrate.  相似文献   

10.
A fluorescence titration assay was used to detect the effects of various modifications of E.coli elongation factor Tu on the formation of the ternary complex with aminoacyl-tRNAs. The treatment of EF-Tu.GDP with TPCK, an analogue of the 3'terminus of aminoacyl-tRNA, was found to have no influence on the conversion of EF-Tu.GDP to 'active' EF-Tu.GTP, but does decrease the affinity of the activated protein for yeast aminoacyl-tRNA by more than three orders of magnitude. Modification of the elongation factor by limited cleavage with trypsin, leading to the excision of amino acid residues 45-58, has only a minor influence on ternary complex formation. The equilibrium dissociation constant of the ternary complex with this trypsin-treated EF-Tu.GTP and E.coli Phe-tRNA(Phe) is only one order of magnitude higher than that of the ternary complex with native EF-Tu. Mutations in the amino acid residues 222 and 375 of EF-Tu also have little effect on ternary complex formation. Compared with TPCK-treated EF-Tu, the affinities of the two mutant species, designated EF-tuAR and EF-TuBO respectively, for [AEDANS-s2C]Tyr-tRNA(Tyr) are only slightly reduced and in the same range as trypsin-cleaved EF-Tu.  相似文献   

11.
The kinetic mechanism of hog kidney D-amino acid oxidase with D-alpha-aminobutyrate as substrate has been examined in detail using a combination of steady state and rapid reaction methods. At concentrations of D-alpha-aminobutyrate below 0.5 mM, the rapid reaction and steady state results are consistent with the mechanism previously proposed for D-alanine (Massey, V., and Gibson, Q. H. (1964) Fed. Proc. 23, 18-29; Porter, D. J. T., Voet, J. G., and Bright, H. J. (1977) J. Biol. Chem. 252, 4464-4473). Both flavin reduction by D-alpha-aminobutyrate and reoxidation are quite rapid. Release of product from the oxidized enzyme has been measured directly and matches the turnover number at infinite concentrations of both substrates. Substitution of deuterium for the alpha-hydrogen decreases the rate of reduction 1.4-fold, without any effect on the apparent Kd. Computer simulations show that the kinetic isotope effects on the reductive half-reaction with D-alanine reported by Porter et al. (see above reference) can be explained using a two-step model with a kinetic isotope effect of 1.75 on the limiting rate of reduction. The effect of enzyme concentration on the kinetics has been examined in some detail. With D-alanine as substrate, increasing the enzyme concentration over the range 29 nM to 17 microM resulted in less than a 2-fold decrease in the turnover number. The Kd for benzoate binding also decreased marginally with increasing enzyme concentration. The effect of enzyme concentration is consistent with a decrease in the rate of release of ligands from the oxidized enzyme as the enzyme concentration is increased.  相似文献   

12.
Ab initio molecular orbital calculations have been performed on the reaction profile for the addition/elimination reaction between ammonia and formic acid, proceeding via a tetrahedral intermediate: NH3 + HCO2H----H2NCH(OH)2----NH2CHO + H2O. Calculated transition state energies for the first addition step of the reaction revealed that a lone pair on the oxygen of the OH group, which is antiperiplanar to the attacking nitrogen, stabilized the transition state by 3.9 kcal/mol, thus supporting the hypothesis of stereoelectronic control for this reaction. In addition, a secondary, counterbalancing stereoelectronic effect stabilizes the second step, water elimination, transition state by 3.1 kcal/mol if the lone pair on the leaving water oxygen is not antiperiplanar to the C-N bond. The best conformation for the transition states was thus one with a lone pair antiperiplanar to the adjacent scissile bond and also one without a lone-pair orbital on the scissile bond oxygen or nitrogen antiperiplanar to the adjacent polar bond. The significance of these stereoelectronic effects for the mechanism of action of serine proteases is discussed.  相似文献   

13.
The kinetic properties of the [3H]ADP-ATP exchange reaction catalyzed by Na+, K+-dependent ATPase [EC 3.6.1,3] were investigated, using NaI-treated microsomes from bovine brain, and the following results were obtained. 1. The rates of the Na+-dependent exchange reaction in the steady state were measured in a solution containing 45 micronM free Mg2+, 100 mMNaCl, 80 micronM ATP, and 160 micronM ADP at pH 6.5 and 4-5 degrees. The rate and amount of decrease in phosphorylated intermediate on adding ADP, i.e., the amount of ADP-sensitive EP, were measured while varying one of the reaction parameters and fixing the others mentioned above. Plots of the exchange rate and the amount of ADP-sensitive EP against the logarithm of free Mg2+ concentration gave bell-shaped curves with maximum values at 50-60 micronM free Mg2+. Plots of the exchange rate and the amount of ADP-sensitive EP against pH also gave bell-shaped curves with maximum values at pH 6.9-7. They both increased with increase in the concentration of NaCl to maximum values at 150-200 mM NaCl, and then decreased rapidly with increase in the NaCl concentration above 200 mM. The dependences of the exchange rate and the amount of ADP-sensitive EP on the concentration of ADP followed the Michaelis-Menten equation, and the Michaelis constants Km, for both were 43 micronM. The dependence of the exchange rate on the ATP concentration also followed the Michaelis-Menten equation, and the Km value was 30 micronM. The amount of ADP-sensitive EP increased with increase in the ATP concentration, and reached a maximum value at about 5 micronM ATP. 2. The N+-dependent [3H]ADP-ATP exchange reaction was started by adding [3H]ADP to EP at low Mg2+-concentration. The reaction consisted of a rapid initial phase and a slow steady phase. The amount of [3H]ATP formed during the rapid initial phase, i.e. the size of the ATP burst, was equal to that of ADP-sensitive EP, and was proportional to the rate in the steady state. At high Mg2+ concentration, the rate of Na+-dependent exchange in the steady state was almost zero, and EP did not show any ADP sensitivity. However, rapid formation of [3H]ATP was observed in the pre-steady state, and the size of the ATP burst increased with increase in the KCl concentration. From these findings, we concluded that an enzyme-ATP complex (E2ATP) formed at low Mg2+ concentration is in equilibrium with EP + ADP, that the rate-limiting step for the exchange reaction is the release of ATP from the enzyme-ATP complex, that the ADP-insensitive EP (formula: see text) produced at high Mg2+ concentration is in equilibrium with the enzyme-ATP complex, and that the equilibrium shifts towards the enzyme-ATP complex on adding KCl. Actually, the ratio of the size of the ATP burst to the amount of EP was equal to the reciprocal of the equilibrium constant of step (formula: see text), determined by a method previously reported by us.  相似文献   

14.
Plasmodium falciparum dihydroorotate dehydrogenase (pfDHODH) is a flavin-dependent mitochondrial enzyme that provides the only route to pyrimidine biosynthesis in the parasite. Clinically significant inhibitors of human DHODH (e.g., A77 1726) bind to a pocket on the opposite face of the flavin cofactor from dihydroorotate (DHO). This pocket demonstrates considerable sequence variability, which has allowed species-specific inhibitors of the malarial enzyme to be identified. Ubiquinone (CoQ), the physiological oxidant in the reaction, has been postulated to bind this site despite a lack of structural evidence. To more clearly define the residues involved in CoQ binding and catalysis, we undertook site-directed mutagenesis of seven residues in the structurally defined A77 1726 binding site, which we term the species-selective inhibitor site. Mutation of several of these residues (H185, F188, and F227) to Ala substantially decreased the affinity of pfDHODH-specific inhibitors (40-240-fold). In contrast, only a modest increase in the Kmapp for CoQ was observed, although mutation of Y528 in particular caused a substantial reduction in kcat (40-100-fold decrease). Pre-steady-state kinetic analysis by single wavelength stopped-flow spectroscopy showed that the mutations had no effect on the rate of the DHO-dependent reductive half-reaction, but most reduced the rate of the CoQ-dependent flavin oxidation step (3-20-fold decrease), while not significantly altering the Kdox for CoQ. As with the mutants, inhibitors that bind this site block the CoQ-dependent oxidative half-reaction without affecting the DHO-dependent step. These results identify residues involved in inhibitor binding and electron transfer to CoQ. Importantly, the data provide compelling evidence that the binding sites for CoQ and species-selective site inhibitors do not overlap, and they suggest instead that inhibitors act either by blocking the electron path between flavin and CoQ or by stabilizing a conformation that excludes CoQ binding.  相似文献   

15.
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   

16.
D-amino acid oxidase (DAAO) from pig has been reported to catalyze the β-elimination of Cl(-) from βCl-D-alanine via abstraction of the substrate α-H as H(+) ("carbanion mechanism") (Walsh, C. T., Schonbrunn, A., and Abeles, R. H. (1971) J. Biol. Chem. 246, 6855-6866). In view of the fundamental mechanistic importance of this reaction and of the recent reinterpretation of the DAAO dehydrogenation step as occurring via a hydride mechanism, we reinvestigated the elimination reaction using yeast DAAO. That enzyme catalyzes the same reactions as the pig enzyme but with a much higher efficiency and a substantially different kinetic behavior. The reaction is initiated by a very rapid and fully reversible dehydrogenation step. This leads to an equilibrium (k(on) ≈ k(reverse)) between the complexes of oxidized enzyme-βCl-D-alanine and reduced enzyme-βCl-iminopyruvate. In the presence of O(2) the latter complex can partition between an oxidative half-reaction and elimination of Cl(-), which proceeds at a rate of ≈50 s(-1). This step forms a complex between oxidized enzyme and enamine that is characterized by a charge transfer absorption (which describes its rates of formation and decay). A minimal scheme that lists relevant steps of the reductive and oxidative half-reactions and elimination pathways along with the estimate of the corresponding rate constants is presented. β-Elimination of Cl(-) is proposed to originate at the locus of the enzyme-βCl-iminopyruvate complex. A chemical mechanism that can account for elimination is discussed in detail.  相似文献   

17.
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   

18.
The reaction between ovarian-cyst glycoproteins and H2O2 was investigated in the presence of a number of inhibitors and catalysts. Azide and 2H2O were separately found to have little effect, implying that singlet oxygen was not involved. Superoxide dismutase was destroyed by H2O2, but mannitol had no effect: thus generalized attack by OH., whether originating from HO2.- or more directly, is not indicated. The glycoproteins contained trace quantities of Cu and Fe, amounting to about 2 atoms of metal per glycoprotein molecule. Treatment of the glycoproteins with the strong chelator DETAPAC (diethylenetriaminepenta-acetic acid) or Chelex resin eliminated the reaction with H2O2; activity could be restored by addition of Cu2+ or Fe2+ in millimolar quantities. It was concluded that metal-ion catalysis is an essential step in the attack of H2O2 on glycoproteins. Spectroscopic and other evidence showed that Cu2+ (and probably Fe2+) complexes strongly with poly-L-histidine, and implies that the Cu2+ or Fe2+ in the glycoproteins is complexed with some of the histidine residues in the glycosylated backbone. Neither polyhistidine nor polyproline reacted with H2O2 in the absence of metal ions, but small quantities of Cu2+ or Fe3+ caused degradation. This was rapid with polyhistidine, which was converted largely into aspartic acid, but slower with polyproline, where limited conversion into glutamic acid occurs. These findings confirm the original hypothesis that peroxide attack on glycoproteins occurs largely at the histidine residues, with simultaneous peptidolysis. The mechanism most probably involves the liberation of OH. by an oxidation-reduction cycle involving, e.g. Cu+/Cu2+: specificity of attack at histidine is due to the location of the metal at these residues only.  相似文献   

19.
2-Hydroxybiphenyl 3-monooxygenase (EC 1.14.13.44) from Pseudomonas azelaica HBP1 is an FAD-dependent aromatic hydroxylase that catalyzes the conversion of 2-hydroxybiphenyl to 2, 3-dihydroxybiphenyl in the presence of NADH and oxygen. The catalytic mechanism of this three-substrate reaction was investigated at 7 degrees C by stopped-flow absorption spectroscopy. Various individual steps associated with catalysis were readily observed at pH 7.5, the optimum pH for enzyme turnover. Anaerobic reduction of the free enzyme by NADH is a biphasic process, most likely reflecting the presence of two distinct enzyme forms. Binding of 2-hydroxybiphenyl stimulated the rate of enzyme reduction by NADH by 2 orders of magnitude. The anaerobic reduction of the enzyme-substrate complex involved the formation of a transient charge-transfer complex between the reduced flavin and NAD(+). A similar transient intermediate was formed when the enzyme was complexed with the substrate analog 2-sec-butylphenol or with the non-substrate effector 2,3-dihydroxybiphenyl. Excess NAD(+) strongly stabilized the charge-transfer complexes but did not give rise to the appearance of any intermediate during the reduction of uncomplexed enzyme. Free reduced 2-hydroxybiphenyl 3-monooxygenase reacted rapidly with oxygen to form oxidized enzyme with no appearance of intermediates during this reaction. In the presence of 2-hydroxybiphenyl, two consecutive spectral intermediates were observed which were assigned to the flavin C(4a)-hydroperoxide and the flavin C(4a)-hydroxide, respectively. No oxygenated flavin intermediates were observed when the enzyme was in complex with 2, 3-dihydroxybiphenyl. Monovalent anions retarded the dehydration of the flavin C(4a)-hydroxide without stabilization of additional intermediates. The kinetic data for 2-hydroxybiphenyl 3-monooxygenase are consistent with a ternary complex mechanism in which the aromatic substrate has strict control in both the reductive and oxidative half-reaction in a way that reactions leading to substrate hydroxylation are favored over those leading to the futile formation of hydrogen peroxide. NAD(+) release from the reduced enzyme-substrate complex is the slowest step in catalysis.  相似文献   

20.
Horse-liver alcohol dehydrogenase was carboxymethylated with iodoacetate, which is known to selectively alkylate cysteine-46 in the polypeptide sequence. Carboxymethyl and native enzyme had the same electrophoretic mobility on starch or polyacrylamide gel, but some separation was achieved when isobutyramide and a low concentration of NADH were present (under these conditions NADH was bound by native enzyme but not by Carboxymethyl enzyme).The Carboxymethyl enzyme formed ternary complexes with NAD+ and pyrazole or decanoate. The fluorescence emission of NADH was enhanced 7- to 8-fold (at 410 nm), and a dissociation-constant of 1.7 μM was calculated at pH 7.4; but, in contrast to native enzyme, neither the affinity nor fluorescence were increased by amides (acetamide or isobutyramide).Carboxymethyl alcohol dehydrogenase possesses catalytic activity. Higher alcohols gave maximum velocities up to 7-fold higher than ethanol (reaching nearly 20% of the activity of native enzyme) while [2H]ethanol showed an isotope-rate effect of 3.3. Although the affinity for aldehydes was considerably increased, the maximum velocity of aldehyde-reduction was always at least 20% of that shown by native enzyme, and at pH 9.9 it was almost 2-fold greater than with native enzyme. The rate-limiting step in alcohol-oxidation is likely to be the interconversion of ternary complexes (possibly the hydride-transfer step), while in aldehyde-reduction it could still be the dissociation of the enzyme/NAD+ complex. This is also indicated by inhibition experiments with decanoate, pyrazole, and isobutyramide.These results suggest that a major effect of carboxymethylation is upon ternary complexes of enzyme and NADH, which become much more reluctant to form, either by combination of NADH and ligand with the modified enzyme, or by catalytic conversion of the enzyme/NAD +/alcohol complex.  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