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 steady state kinetic mechanism, molecular isotope exchange and the positional isotope exchange (PIX) reactions of D-alanyl-D-alanine ligase from Salmonella typhimurium have been studied. The kinetic mechanism has been determined to be ordered Ter-Ter from initial velocity and product inhibition experiments. The first substrate to bind is ATP followed by the addition of 2 mol of D-alanine. Pi is released, and then D-alanyl-D-alanine and ADP dissociate from the enzyme surface. In the reverse direction D-alanyl-D-alanine exhibits complete substrate inhibition (Ki = 1.15 +/- 0.05 mM) by binding to the enzyme-ATP complex. In the presence of D-alanine, D-alanyl-D-alanine ligase catalyzed the positional exchange of the beta,gamma-bridge oxygen in [gamma-18O4]ATP to a beta-nonbridge position. Two possible alternate dead-end substrate analogs, D-2-chloropropionic acid and isobutyric acid, did not induce a positional isotope exchange in [gamma-18O4]ATP. The positional isotope exchange rate is diminished relative to the net substrate turnover as the concentration of D-alanine is increased. This is consistent with the ordered Ter-Ter mechanism as determined by the steady state kinetic experiments. The ratio of the positional isotope exchange rate relative to the net chemical turnover of substrate (Vex/Vchem) approaches a value of 1.4 as the concentration of D-alanine becomes very small. This ratio is 100 times larger than the ratio of the maximal reverse and forward chemical reaction velocities (V2/V1). This situation is only possible when the reaction mechanism proceeds in two distinct steps and the first step is much faster than the second step. The enzyme was also found to catalyze the molecular isotope exchange of radiolabeled D-alanine with D-alanyl-D-alanine in the presence of phosphate. These results are consistent with the formation of D-alanyl phosphate as a kinetically competent intermediate.  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.
P Marcotte  C Walsh 《Biochemistry》1978,17(14):2864-2868
Upon oxidation of D-propargylglycine by D-amino acid oxidase, the enzyme is converted by covalent alkylation to catalytic species with different properties from those of native enzyme. At least five distinct modified enzyme species are present in the preparation, as determined by gel electro-focusing. Individual characterization of the components has not yet been attempted. The combined kinetic and spectral properties of the preparation have been studied. The modified enzymes have a marked preference for hydrophobic amino acids: the rates of oxidation decrease in the series D-phenylalanine, D-methionine, D-norleucine, D-norvaline, D-alpha-aminobutyrate, D-alanine. In addition, the observed Kms of the amino acids are increased, especially those of the smaller substrates (D-alanine and D-alpha-aminobutyrate). A primary kinetic isotope effect is observed upon oxidation of amino acids by the modified enzymes, evidence that this catalysis exhibits a different rate-determining step from catalysis by native enzyme. The modified apoenzyme exhibits intense absorbance at 318--320 nm, not present in native enzyme. This chromophore can be partially (75%) removed by treatment of the modified enzyme with hydrazine. However, the activity of native enzyme is not substantially restored by this process, suggesting the existence of superficial alkylations in addition to the modification responsible for the observed changes in kinetic parameters.  相似文献   

4.
Nandigama RK  Edmondson DE 《Biochemistry》2000,39(49):15258-15265
The interaction of recombinant human liver monoamine oxidase A (MAO A) with a series of phenethylamine substrate analogues has been investigated by steady-state and stopped-flow kinetic techniques. Substrate analogues with para substituents exhibit large deuterium kinetic isotope effect on k(cat), on k(cat)/K(m), and on the limiting rate of enzyme reduction in reductive half-reaction experiments. These kinetic isotope effect values range from 5 to 10 with the exception of tyramine, which exhibited smaller steady-state isotope effects (2.3-3.5) than that observed on the rate of flavin reduction (6.9). The stopped-flow data show that imine release from the reduced enzyme is slower than the rate of catalytic turnover. Phenethylamine oxidation by MAO A can be described as the C-H bond cleavage step being rate limiting in catalysis and with oxygen reacting with the reduced enzyme-imine complex. In the case of tyramine, the product release from the oxidized enzyme-imine complex contributes to the rate limitation in catalysis. The binding affinities of a series of para-substituted phenethylamine analogues to MAO A show an increase in affinity of the deprotonated amine with increasing van der Waals volume of the substituent. The limiting rate of enzyme reduction decreases with increasing van der Waals volume of the substituent in a linear manner with no observable electronic contribution as observed previously with benzylamine reduction of MAO A [Miller, J. R., and Edmondson, D. E. (1999) Biochemistry 38, 13670-13683]. Examination of side chain analogues of phenethylamine show 3-phenylpropylamine to be oxidized 2.5-fold more slowly and bound 75-fold more tightly than phenethylamine. 4-Phenylbutylamine is not a substrate for MAO A but is a good competitive inhibitor with a K(i) value of 31 +/- 5 microM. Analysis of the effect of alkyl side chain alterations on binding affinities of a series of arylalkylamine analogues taken from this study and from the literature show a linear correlation with the Taft steric value (E(s)) of the side chain. These results suggest that the binding site for the aryl ring is identical for phenethylamine and for benzylamine analogues and that steric interactions of the alkyl side chain with the enzyme strongly contribute to the binding affinities of a series of reversible inhibitors of MAO A.  相似文献   

5.
J R Miller  D E Edmondson 《Biochemistry》1999,38(41):13670-13683
Monoamine oxidase A (MAO A) plays a central role in the oxidation of amine neurotransmitters. To investigate the structure and mechanism of this enzyme, recombinant human liver MAO A was expressed and purified from Saccharomyces cerevisiae. Anaerobic titrations of the enzyme require only 1 mol of substrate per mole of enzyme-bound flavin for complete reduction. This demonstrates that only one redox-active group (i.e., the covalent FAD cofactor) is involved in catalysis. The reaction rates and binding affinities of 17 para-substituted benzylamine analogues with purified MAO A were determined by steady state and stopped flow kinetic experiments. For each substrate analogue that was tested, the rates of steady state turnover (k(cat)) and anaerobic flavin reduction (k(red)) are similar in value. Deuterium kinetic isotope effects on k(cat), k(red), k(cat)/K(m), and k(red)/K(s) with alpha, alpha-[(2)H]benzylamines are similar for each substrate analogue that was tested and range in value from 6 to 13, indicating that alpha-C-H bond cleavage is rate-limiting in catalysis. Substrate analogue dissociation constants determined from reductive half-reaction experiments as well as from steady state kinetic isotope effect data [Klinman, J. P., and Matthews, R. G. (1985) J. Am. Chem. Soc. 107, 1058-1060] are in excellent agreement. Quantitative structure-activity relationship (QSAR) analysis of dissociation constants shows that the binding of para-substituted benzylamine analogues to MAO A is best correlated with the van der Waals volume of the substituent, with larger substituents binding most tightly. The rate of para-substituted benzylamine analogue oxidation and/or substrate analogue-dependent flavin reduction is best correlated with substituent electronic effects (sigma). Separation of the electronic substituent parameter (sigma) into field-inductive and resonance effects provides a more comprehensive treatment of the electronic correlations. The positive correlation of rate with sigma (rho approximately 2.0) suggests negative charge development at the benzyl carbon position occurs and supports proton abstraction as the mode of alpha-C-H bond cleavage. These results are discussed in terms of several mechanisms proposed for MAO catalysis and with previous structure-activity studies published with bovine liver MAO B [Walker, M. C., and Edmondson, D. E. (1994) Biochemistry 33, 7088-7098].  相似文献   

6.
Rapid reaction kinetics of the flavoprotein anthranilate hydroxylase from Trichosporon cutaneum were examined for reactions involving anthranilate, the native substrate. As was reported earlier for the nonhydroxylated substrate analogue, salicylate, some reactions in the first turnover with anthranilate occur slower than those in subsequent turnovers (Powlowski, J., Massey, V., and Ballou, D. P. (1989) J. Biol. Chem. 264, 5606-5612). Evidence is presented for slow conformational changes that occur both on binding of the aromatic ligand and on reduction of the enzyme. These changes are apparently important for rapid anthranilate binding to occur in turnovers subsequent to the first. Moreover, bound anthranilate is required for rapid reduction of enzyme-bound FAD by NADPH. Studies to probe the accessibility of reagents to modified flavins that had been incorporated into the apoenzyme indicate that anthranilate binding causes a conformational change in the protein, allowing increased access to the benzene ring moiety of the flavin. An unusual isotope effect with (R)-NADPD (4(R)-2H] NADPH) is observed on Kd rather than on kred, which is consistent with a model involving slow interconversion of enzyme-substrate complexes before productive binding of NADPH and reduction of the enzyme flavin.  相似文献   

7.
J M Denu  P F Fitzpatrick 《Biochemistry》1992,31(35):8207-8215
Primary deuterium kinetic isotope and pH effects on the reduction of D-amino acid oxidase by amino acid substrates were determined using steady-state and rapid reaction methods. With D-serine as substrate, reduction of the enzyme-bound FAD requires that a group with a pKa value of 8.7 be unprotonated and that a group with a pKa value of 10.7 be protonated. The DV/Kser value of 4.5 is pH-independent, establishing that these pKa values are intrinsic. The limiting rate of reduction of the enzyme shows a kinetic isotope effect of 4.75, consistent with this as the intrinsic value. At high enzyme concentration (approximately 15 microM) at pH 9,D-serine is slightly sticky (k3/k2 = 0.8), consistent with a decrease in the rate of substrate dissociation. With D-alanine as substrate, the pKa values are perturbed to 8.1 and 11.5. The DV/Kala value increases from 1.3 at pH 9.5 to 5.1 at pH 4, establishing that D-alanine is sticky with a forward commitment of approximately 10. The effect of pH on the DV/Kala value is consistent with a model in which exchange with solvent of the proton from the group with pKa 8.7 is hindered and is catalyzed by H2O and OH- above pH 7 and by H3O+ and H2O below pH 7. With glycine, the pH optimum is shifted to a more basic value, 10.3. The DV/Kgly value increases from 1.26 at pH 6.5 to 3.1 at pH 10.7, consistent with fully reversible CH bond cleavage followed by a pH-dependent step. At pH 10.5, the kinetic isotope effect on the limiting rate of reduction is 3.4.  相似文献   

8.
S-Adenosylhomocysteine hydrolase (AdoHcy hydrolase) crystallizes from solutions containing the intermediate analogue neplanocin A with the analogue bound in its 3'-keto form at the active sites of all of its four subunits and the four tightly bound cofactors in their reduced (NADH) state. The enzyme is in the closed conformation, which corresponds to the structure in which the catalytic chemistry occurs. Examination of the structure in the light of available, very detailed kinetic studies [Porter, D. J., Boyd, F. L. (1991) J. Biol. Chem. 266, 21616-21625. Porter, D. J., Boyd, F. L. (1992) J. Biol. Chem. 267, 3205-3213. Porter, D. J. (1998) J. Biol. Chem. 268, 66-73] suggests elements of the catalytic strategy of AdoHcy hydrolase for acceleration of the reversible conversion of AdoHcy to adenosine (Ado) and homocysteine (Hcy). The enzyme, each subunit of which possesses a substrate-binding domain that in the absence of substrate is in rapid motion relative to the tetrameric core of the enzyme, first binds substrate and ceases motion. Probably concurrently with oxidation of the substrate to its 3'-keto form, the closed active site is "sealed off" from the environment, as indicated by a large (10(8)(-)(9)-fold) reduction in the rate of departure of ligands, a feature that prevents exposure of the labile 3'-keto intermediates to the aqueous environment. Elimination of the 5'-substituent (Hcy in the hydrolytic direction, water in the synthetic direction) generates the central intermediate 4',5'-didehydro-5'-deoxy-3'-ketoadenosine. Abortive 3'-reduction of the central intermediate is prevented by a temporary suspension of all or part of the redox catalytic power of the enzyme during the existence of the central intermediate. The abortive reduction is 10(4)-fold slower than the productive reductions at the ends of the catalytic cycle and has a rate constant similar to those of nonenzymic intramolecular model reactions. The mechanism for suspending the redox catalytic power appears to be a conformationally induced increase in the distance across which hydride transfer must occur between cofactor and substrate, the responsible conformational change again being that which "seals" the active site. The crystal structure reveals a well-defined chain of three water molecules leading from the active site to the subunit surface, which may serve as a relay for proton exchange between solvent and active site in the closed form of the enzyme, permitting maintenance of active-site functional groups in catalytically suitable protonation states.  相似文献   

9.
M A Vanoni  R G Matthews 《Biochemistry》1984,23(22):5272-5279
Previous work from this laboratory has established that the NADPH-menadione oxidoreductase reaction catalyzed by methylenetetrahydrofolate reductase from pig liver proceeds by Ping Pong Bi Bi kinetics and that the reductive half-reaction is rate limiting in steady-state turnover. We have now shown that methylenetetrahydrofolate reductase stereo-specifically removes the pro-S hydrogen from the 4-position of NADPH. During the oxidation of [4(S)-3H]NADPH, we observed a kinetic isotope on V/KNADPH of 10.8 +/- 0.4. When comparing the rates of oxidation of [4(S)-2H]NADPH and [4(S)-1H]NADPH, we measure kinetic isotope effects on V of 4.78 +/- 0.15 and on V/KNADPH of 4.54 +/- 0.59. When oxidation of [4(R)-2H]NADPH and [4(R)-1H]NADPH is compared, the secondary kinetic isotope effect on V is 1.04 +/- 0.01. When the NADPH-menadione oxidoreductase reaction is catalyzed in tritiated water, no incorporation of solvent tritium into residual NADPH is observed. We conclude from these observations that the oxidation of NADPH is largely or entirely rate limiting in the reductive half-reaction and, hence, in NADPH-menadione oxidoreductase turnover at saturating menadione concentration. In the presence of saturating NADPH, the flavin reduction proceeds with a rate constant of 160 S-1, which is at least 29-fold slower than estimates of the lower limit for the diffusion-limited rate constant characterizing NADPH binding to the enzyme under physiological conditions. Albery & Knowles have defined criteria for perfection in enzyme catalysis [Albery, W. J., & Knowles, J.R. (1976) Biochemistry 15, 5631-5640].(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

10.
The effect of isotopic substitution of the 8-H of xanthine (with 2H and 3H) on the rate of oxidation by bovine xanthine oxidase and by chicken xanthine dehydrogenase has been measured. V/K isotope effects were determined from competition experiments. No difference in H/T(V/K) values was observed between xanthine oxidase (3.59 +/- 0.1) and xanthine dehydrogenase (3.60 +/- 0.09). Xanthine dehydrogenase exhibited a larger T/D(V/K) value (0.616 +/- 0.028) than that observed for xanthine oxidase (0.551 +/- 0.016). Observed H/T(V/K) values for either enzyme are less than those H/T(V/K) values calculated with D/T(V/K) data. These discrepancies are suggested to arise from the presence of a rate-limiting step(s) prior to the irreversible C-H bond cleavage step in the mechanistic pathways of both enzymes. These kinetic complexities preclude examination of whether tunneling contributes to the reaction coordinate for the H-transfer step in each enzyme. No observable exchange of tritium with solvent is observed during the anaerobic incubation of [8-3H]xanthine with either enzyme, which suggests the reverse commitment to catalysis (Cr) is essentially zero. With the assumption of adherence to reduced mass relationships, the intrinsic deuterium isotope effect (Dk) for xanthine oxidation is calculated to be 7.4 +/- 0.7 for xanthine oxidase and 4.2 +/- 0.2 for xanthine dehydrogenase. By use of these values and steady-state kinetic data, the minimal rate for the hydrogen-transfer step is calculated to be approximately 75-fold faster than kcat for xanthine oxidase and approximately 10-fold faster than kcat for xanthine dehydrogenase. This calculated rate is consistent with data obtained by rapid-quench experiments with XO. A stoichiometry of 1.0 +/- 0.3 mol of uric acid/mol of functional enzyme is formed within the mixing time of the instrument (5-10 ms). The kinetic isotope effect data also permitted the calculation of the Kd values [Klinman, J. P., & Mathews, R. G. (1985) J. Am. Chem. Soc. 107, 1058-1060] for substrate dissociation, including all reversible steps prior to C-H bond cleavage. Values calculated for each enzyme (Kd = 120 microM) were found to be identical within experimental uncertainty.  相似文献   

11.
C A Gates  D B Northrop 《Biochemistry》1988,27(10):3820-3825
Aminoglycoside nucleotidyltransferase 2'-I (formerly gentamicin adenylyltransferase) conveys antibiotic resistance to Gram-negative bacteria by transfer of AMP to the 2'-hydroxyl group of 4,6-substituted deoxystreptamine-containing aminoglycosides. The kinetics constants of thirteen aminoglycoside antibiotics and the magnesium chelates of eight nucleotide triphosphates were determined with purified enzyme. Eleven of the antibiotics exhibit substrate inhibition attributed to secondary binding of the aminoglycoside to an enzyme-AMP-aminoglycoside complex. Maximal velocities vary by only 4-fold, versus variation of values of Vmax/Km for the aminoglycosides of nearly 4000-fold, consistent with a Theorell-Chance kinetic mechanism as proposed for this enzyme [Gates, C. A., & Northrop, D. B. (1988) Biochemistry (second of three papers in this issue)] with the added specification that the binding of aminoglycosides is in rapid equilibrium. Under these conditions, Vmax/Km becomes kcat/Kd, where kcat is the net rate constant for catalysis (but not turnover) and Kd is the dissociation constant of aminoglycosides from a complex with enzyme and nucleotide. Values of kcat fall closely together into three distinct sets, with the 3',4'-dideoxygentamicins greater than gentamicins greater than kanamycins. These sets reflect unusual structure-activity correlations which are specific for catalysis but have nothing to do with the maximal velocity of this enzyme.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

12.
Serine 335 at the active site of D-amino acid oxidase from the yeast Rhodotorula gracilis (RgDAAO) is not conserved in other DAAO sequences. To assess its role in catalysis, it was mutated to Gly, the residue present in mammalian DAAO, an enzyme with a 35-fold lower turnover number with D-alanine. The spectral and ligand binding properties of the S335G mutant are similar to those of wild-type enzyme, suggesting an active site with minimally altered electrostatic properties. The S335G mutant is catalytically active, excluding an essential role of S335 in catalysis. However, S335-OH contributes to the high efficiency of the mutant enzyme since the catalytic activity of the latter is lower due to a decreased rate of flavin reduction relative to wild-type RgDAAO. Catalytic rates are pH-dependent and appear to converge to very low, but finite and similar values at low pH for both wild-type and S335G RgDAAO. While this dependence exhibits two apparent pKs with wild-type RgDAAO, with the S335G mutant a single, apparent pK approximately 8 is observed, which is attributed to the ionization of the alphaNH2 group of the bound substrate. Removal of S335-OH thus suppresses an apparent pK approximately 6. Both wild-type RgDAAO and the S335G mutant exhibit a substantial deuterium solvent kinetic isotope effect (> or =4) at pH<7 that disappears with increasing pH and reflects a pKapp=6.9 +/- 0.4. Interestingly, the substitution suppresses the activity towards d-lactate, suggesting a role of the serine 335 in removal of the substrate alpha-OH hydrogen.  相似文献   

13.
Liver microsomal flavin-containing monooxygenase (MFMO) has been shown to exhibit a stable 4a-flavin hydroperoxide intermediate in the absence of oxygenatable substrate (Poulsen, L. L., and Ziegler, D. M. (1979) J. Biol. Chem. 254, 6449-6455; Beaty, N. B., and Ballou, D. P. (1981) J. Biol. Chem. 256, 4619-4625). The reaction of this intermediate with an assortment of substrates was studied by stopped flow techniques. The first observed spectral change is a small blue shift in the absorbance peak of the 4a-flavin intermediate. The rate of this spectral change is dependent on the concentration of the substrate. This small spectral change is succeeded by a large increase in the absorbance at 450 nm. The rate of appearance of oxidized flavin is independent of substrate concentration but does increase at higher pH. Steady state turnover rates also greater at higher pH, consistent with earlier observations that the formation of oxidized flavin is rate determining in catalysis. Upon oxygenation by MFMO, thiobenzamide and iodide each undergo a spectral change which is dependent on substrate concentration. The spectral changes corresponding to oxygenation of these substrates occur at the same rates as do the initial small spectral changes contributed by the flavin chromophore as observed with all substrates. However, no substrate tested to date shows any effect on the rate of formation of oxidized flavin. Previous work has shown MFMO to catalyze the oxygenation of a variety of nitrogen- and sulfur-containing hydrophobic compounds. Two new classes of compounds are shown here to be substrates for this enzyme. The nucleophilic anions, iodide and thiocyanate, catalyze the decomposition of the 4a-flavin hydroperoxide. Organic boronic acids (e.g. phenylboronic acid and butylboronic acid) also appear to be oxygenated with no striking differences in kinetic characteristics from those of nucleophilic substrates. These organic boronic acids are classic electrophiles and suggest that like peracids, the 4a-flavin hydroperoxide is capable of oxygenating both nucleophiles and electrophiles (Lee, J. B., and Uff, B. C. (1967) Quart. Rev. 21, 429-457).  相似文献   

14.
The kinetic mechanism of octopine dehydrogenase has been investigated by stopped-flow and isotope replacement techniques. When the enzyme is saturated by substrate and coenzyme, both for NADH oxidation and NAD+ reduction, the stationary phase is preceded by a rapid burst. Under these saturation conditions, furthermore, the stationary phase shows a secondary isotope effect when 4S-[4(2)H]NADH is substituted for NADH and when (on the other reaction end) D-[2H] octopine is substituted for D-octopine. The data are taken to indicate that the rate-limiting step for enzyme turnover is a step following a very fast chemical transformation of the reagents. However, when the substrate concentration is lowered below the corresponding Km value keeping the coenzyme concentration at saturating levels, the time course of the reaction shows no burst and the stationary phase has a larger isotope effect. This indicated that under those non-saturating conditions, the enzyme turnover has a larger contribution than the hydrogen-transfer step. Changing the coenzyme concentration alone has very little or no effect on the amplitude of the burst or on the isotope effect. These features are discussed in terms of the other known kinetic properties of the enzyme, and in terms of analogous studies reported in the literature for other dehydrogenases.  相似文献   

15.
Human methionine synthase reductase (MSR) is a key enzyme in folate and methionine metabolism as it reactivates the catalytically inert cob(II)alamin form of methionine synthase (MS). Electron transfer from MSR to the cob(II)alamin cofactor coupled with methyl transfer from S-adenosyl methionine returns MS to the active methylcob(III)alamin state. MSR contains stoichiometric amounts of FAD and FMN, which shuttle NADPH-derived electrons to the MS cob(II)alamin cofactor. Herein, we have investigated the pre-steady state kinetic behavior of the reductive half-reaction of MSR by anaerobic stopped-flow absorbance and fluorescence spectroscopy. Photodiode array and single-wavelength spectroscopy performed on both full-length MSR and the isolated FAD domain enabled assignment of observed kinetic phases to mechanistic steps in reduction of the flavins. Under single turnover conditions, reduction of the isolated FAD domain by NADPH occurs in two kinetically resolved steps: a rapid (120 s(-1)) phase, characterized by the formation of a charge-transfer complex between oxidized FAD and NADPH, is followed by a slower (20 s(-1)) phase involving flavin reduction. These two kinetic phases are also observed for reduction of full-length MSR by NADPH, and are followed by two slower and additional kinetic phases (0.2 and 0.016 s(-1)) involving electron transfer between FAD and FMN (thus yielding the disemiquinoid form of MSR) and further reduction of MSR by a second molecule of NADPH. The observed rate constants associated with flavin reduction are dependent hyperbolically on NADPH and [4(R)-2H]NADPH concentration, and the observed primary kinetic isotope effect on this step is 2.2 and 1.7 for the isolated FAD domain and full-length MSR, respectively. Both full-length MSR and the separated FAD domain that have been reduced with dithionite catalyze the reduction of NADP+. The observed rate constant of reverse hydride transfer increases hyperbolically with NADP+ concentration with the FAD domain. The stopped-flow kinetic data, in conjunction with the reported redox potentials of the flavin cofactors for MSR [Wolthers, K. R., Basran, J., Munro, A. W., and Scrutton, N. S. (2003) Biochemistry, 42, 3911-3920], are used to define the mechanism of electron transfer for the reductive half-reaction of MSR. Comparisons are made with similar stopped-flow kinetic studies of the structurally related enzymes cytochrome P450 reductase and nitric oxide synthase.  相似文献   

16.
17.
R M Gould  B V Plapp 《Biochemistry》1990,29(23):5463-5468
Molecular modeling of alcohol dehydrogenase suggests that His-47 in the yeast enzyme (His-44 in the protein sequence, corresponding to Arg-47 in the horse liver enzyme) binds the pyrophosphate of the NAD coenzyme. His-47 in the Saccharomyces cerevisiae isoenzyme I was substituted with an arginine by a directed mutation. Steady-state kinetic results at pH 7.3 and 30 degrees C of the mutant and wild-type enzymes were consistent with an ordered Bi-Bi mechanism. The substitution decreased dissociation constants by 4-fold for NAD+ and 2-fold for NADH while turnover numbers were decreased by 4-fold for ethanol oxidation and 6-fold for acetaldehyde reduction. The magnitudes of these effects are smaller than those found for the same mutation in the human liver beta enzyme, suggesting that other amino acid residues in the active site modulate the effects of the substitution. The pH dependencies of dissociation constants and other kinetic constants were similar in the two yeast enzymes. Thus, it appears that His-47 is not solely responsible for a pK value near 7 that controls activity and coenzyme binding rates in the wild-type enzyme. The small substrate deuterium isotope effect above pH 7 and the single exponential phase of NADH production during the transient oxidation of ethanol by the Arg-47 enzyme suggest that the mutation makes an isomerization of the enzyme-NAD+ complex limiting for turnover with ethanol.  相似文献   

18.
The reaction of PQQ-dependent methanol dehydrogenase (MDH) from Methylophilus methylotrophus has been studied by steady-state and stopped-flow kinetic methods, with particular reference to multiple ligand binding and the kinetic isotope effect (KIE) for PQQ reduction. Phenazine ethosulfate (PES; an artificial electron acceptor) and cyanide (a suppressant of endogenous activity), but not ammonium (an activator of MDH), compete for binding at the catalytic methanol-binding site. Cyanide does not activate turnover in M. methylotrophus MDH, as reported previously for the Paracoccus denitrificans enzyme. Activity is dependent on activation by ammonium but is inhibited at high ammonium concentrations. PES and methanol also influence the stimulatory and inhibitory effects of ammonium through competitive binding. Reaction profiles as a function of ammonium and PES concentration differ between methanol and deuterated methanol, owing to force constant effects on the binding of methanol to the stimulatory and inhibitory ammonium binding sites. Differential binding gives rise to unusual KIEs for PQQ reduction as a function of ammonium and PES concentration. The observed KIEs at different ligand concentrations are independent of temperature, consistent with their origin in differential binding affinities of protiated and deuterated substrate at the ammonium binding sites. Stopped-flow studies indicate that enzyme oxidation is not rate-limiting at low ammonium concentrations (<4 mM) during steady-state turnover. At higher ammonium concentrations (>20 mM), the low effective concentration of PES in the active site owing to the competitive binding of ammonium lowers the second-order rate constant for enzyme oxidation, and the oxidative half-reaction becomes more rate limiting. A sequential stopped-flow method is reported that has enabled, for the first time, a detailed study of the reductive half-reaction of MDH and comparison with steady-state data. The limiting rate of PQQ reduction (0.48 s(-1)) is less than the steady-state turnover number, and the observed KIE in stopped-flow studies is unity. Although catalytically active, we propose reduction of the oxidized enzyme generated in stopped-flow analyses is gated by conformational change or ligand exchange. Slow recovery from this trapped state on mixing with methanol accounts for the slow reduction of PQQ and a KIE of 1. This study emphasizes the need for caution in using inflated KIEs, and the temperature dependence of KIEs, as a probe for hydrogen tunneling.  相似文献   

19.
The reaction of Old Yellow Enzyme (OYE) with pyridine nucleotides has been examined using steady state kinetics, rapid reaction kinetics, and equilibrium binding. alpha-NADPH, beta-NADPH, and the acid breakdown products of NADPH all bind to oxidized OYE with dissociation constants below 1 microM. These complexes produce characteristic red shifts in the absorption spectrum of OYE. A similar red shift which occurs after multiple turnovers of OYE with NADPH has been found to be due to an impurity in the NADPH preparation, possibly an acid breakdown product. Anions such as chloride, acetate, azide, and phenolates compete with the pyridine nucleotides for binding to a common site in oxidized OYE. Anaerobic reduction of OYE by NADPH proceeds via two intermediates to establish a readily reversible equilibrium. In contrast to most other NADPH-dependent enzymes, both alpha- and beta-NADPH are capable of reducing OYE, and alpha-NADPH is more effective. Using beta-[4(R)-2H]NADPH, a primary deuterium isotope effect was observed in the reduction reaction. Results from rapid reaction and steady state studies showed that reduction of OYE was rate limiting in turnover. Consistent with this, the turnover number with alpha-NADPH was significantly higher than that with beta-NADPH.  相似文献   

20.
D-amino acid transaminase, which contains pyridoxal 5'-phosphate (vitamin B6) as coenzyme, catalyzes the formation of D-alanine and D-glutamate from their corresponding alpha-keto acids; these D-amino acids are required for bacterial cell wall biosynthesis. Under conditions usually used for kinetic assay of enzyme activity, i.e., short incubation times with dilute enzyme concentrations, D-alanine behaves as one of the best substrates. However, the enzyme slowly loses activity over a period of hours when exposed to substrates, intermediates, and products at equilibrium. The rate of inactivation is dependent on enzyme concentration but independent of substrate concentration greater than Km values. Continuous removal of the product pyruvate by enzymic reduction precludes the establishment of equilibrium and prevents inactivation. The formation of small but detectable amounts of a quinonoid intermediate absorbing at 493 nm is proportional to inactivation. Studies with [14C]-D-alanine labeled on different carbon atoms indicate that the alpha-carboxyl group of the substrate is absent in the inactive enzyme; such decarboxylation is not a usual function of this enzyme. The inactive transaminase contains 1.1 mol of [14C]-D-alanine-derived adduct per mole of dimeric enzyme; this finding is consistent with the 50% reduction in the fluorescence intensity at 390 nm (due to the PMP form of the coenzyme) for the inactive enzyme. Thus, inactivation of one subunit of the dimeric enzyme renders the entire molecule inactive. Inactivation may occur when a coenzyme intermediate, perhaps the ketimine, is slowly decarboxylated and then undergoes a conformational change from its catalytically competent location.(ABSTRACT TRUNCATED AT 250 WORDS)  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