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.
Homoserine dehydrogenase (HSD), which is required for the synthesis of threonine, isoleucine and methionine in fungi, is a potential target for novel antifungal drugs. In order to design effective inhibitors, the kinetic mechanism of Saccharomyces cerevisiae HSD and the stereochemistry of hydride transfer were examined. Product inhibition experiments revealed that yeast HSD follows an ordered Bi Bi kinetic mechanism, where NAD(P)H must bind the enzyme prior to aspartate semialdehyde (ASA) and homoserine is released first followed by NAD(P)+. H-(1,2,4-triazol-3-yl)-D,L-alanine was an uncompetitive inhibitor of HSD with respect to NADPH (K(ii)=3.04+/-0.18 mM) and a noncompetitive inhibitor with respect to ASA (K(is)=1.64+/-0.36 mM, K(ii)=3.84+/-0.46 mM), in agreement with the proposed substrate order. Both kinetic isotope and viscosity experiments provided evidence for a very rapid catalytic step and suggest nicotinamide release to be primarily rate limiting. Incubation of HSD with stereospecifically deuterated NADP[2H] and subsaturating amounts of aspartate semialdehyde revealed that the pro-S NADPH hydride is transferred to the aldehyde. The pH dependence of steady state kinetic parameters indicate that ionizable groups with basic pKs may be involved in substrate binding, consistent with the observation of Lys223 at the enzyme active site in the recently determined 3D structure [B. DeLaBarre, P.R. Thompson, G.D. Wright, A.M. Berghuis, Nat. Struct. Biol. 7 (2000) 238-244]. These findings provide the requisite foundation for future exploitation of fungal HSD in inhibitor design.  相似文献   

2.
An effective means of relieving the toxicity of furan aldehydes, furfural (FFA) and 5-hydroxymethylfurfural (HMF), on fermenting organisms is essential for achieving efficient fermentation of lignocellulosic biomass to ethanol and other products. Ari1p, an aldehyde reductase from Saccharomyces cerevisiae, has been shown to mitigate the toxicity of FFA and HMF by catalyzing the NADPH-dependent conversion to corresponding alcohols, furfuryl alcohol (FFOH) and 5-hydroxymethylfurfuryl alcohol (HMFOH). At pH 7.0 and 25°C, purified Ari1p catalyzes the NADPH-dependent reduction of substrates with the following values (k(cat) (s(-1)), k(cat)/K(m) (s(-1)mM(-1)), K(m) (mM)): FFA (23.3, 1.82, 12.8), HMF (4.08, 0.173, 23.6), and dl-glyceraldehyde (2.40, 0.0650, 37.0). When acting on HMF and dl-glyceraldehyde, the enzyme operates through an equilibrium ordered kinetic mechanism. In the physiological direction of the reaction, NADPH binds first and NADP(+) dissociates from the enzyme last, demonstrated by k(cat) of HMF and dl-glyceraldehyde that are independent of [NADPH] and (K(ia)(NADPH)/k(cat)) that extrapolate to zero at saturating HMF or dl-glyceraldehyde concentration. Microscopic kinetic parameters were determined for the HMF reaction (HMF+NADPH?HMFOH+NADP(+)), by applying steady-state, presteady-state, kinetic isotope effects, and dynamic modeling methods. Release of products, HMFOH and NADP(+), is 84% rate limiting to k(cat) in the forward direction. Equilibrium constants, [NADP(+)][FFOH]/[NADPH][FFA][H(+)]=5600×10(7)M(-1) and [NADP(+)][HMFOH]/[NADPH][HMF][H(+)]=4200×10(7)M(-1), favor the physiological direction mirrored by the slowness of hydride transfer in the non-physiological direction, NADP(+)-dependent oxidation of alcohols (k(cat) (s(-1)), k(cat)/K(m) (s(-1)mM(-1)), K(m) (mM)): FFOH (0.221, 0.00158, 140) and HMFOH (0.0105, 0.000104, 101).  相似文献   

3.
Fungal homoserine dehydrogenase (HSD) is required for the biosynthesis of threonine, isoleucine and methionine from aspartic acid, and is a target for antifungal agents. HSD from the yeast Saccharomyces cerevisiae was overproduced in Escherichia coli and 25 mg of soluble dimeric enzyme was purified per liter of cell culture in two steps. HSD efficiently reduces aspartate semialdehyde to homoserine (Hse) using either NADH or NADPH with kcat/Km in the order of 10(6-7) M(-1) x s(-1) at pH 7.5. The rate constant of the reverse direction (Hse oxidation) was also significant at pH 9.0 (kcat/Km approximately 10(4-5) M(-1) x s(-1)) but was minimal at pH 7.5. Chemical modification of HSD with diethyl pyrocarbonate (DEPC) resulted in a loss of activity that could be obviated by the presence of substrates. UV difference spectra revealed an increase in absorbance at 240 nm for DEPC-modified HSD consistent with the modification of two histidines (His) per subunit. Amino acid sequence alignment of HSD illustrated the conservation of two His residues among HSDs. These residues, His79 and His309, were substituted to alanine (Ala) using site directed mutagenesis. HSD H79A had similar steady state kinetics to wild type, while kcat/Km for HSD H309A decreased by almost two orders of magnitude. The recent determination of the X-ray structure of HSD revealed that His309 is located at the dimer interface [B. DeLaBarre, P.R. Thompson, G.D. Wright, A.M. Berghuis, Nat. Struct. Biol. 7 (2000) 238-244]. The His309Ala mutant enzyme was found in very high molecular weight complexes rather than the expected dimer by analytical gel filtration chromatography analysis. Thus the invariant His309 plays a structural rather than catalytic role in these enzymes.  相似文献   

4.
Transhydrogenase is a proton pump. It has three components: dI and dIII protrude from the membrane and contain the binding sites for NAD(H) and NADP(H), respectively, and dII spans the membrane. We have expressed dIII from Homo sapiens transhydrogenase (hsdIII) in Escherichia coli. The purified protein was associated with stoichiometric amounts of NADP(H) bound to the catalytic site. The NADP+ and NADPH were released only slowly from the protein, supporting the suggestion that nucleotide-binding by dIII is regulated by the membrane-spanning dII. HsdIII formed a catalytically active complex with recombinant dI from Rhodospirillum rubrum (rrdI), even in the absence of dII. The rates of forward and reverse transhydrogenation catalysed by this complex are probably limited by slow release from dIII of NADPH and NADP+, respectively. The hybrid complex also catalysed high rates of 'cyclic' transhydrogenation, indicating that hydride transfer, and exchange of nucleotides with dI, are rapid. Stopped-flow experiments revealed a rapid, monoexponential, single-turnover burst of reverse transhydrogenation in pre-steady-state. The apparent first-order rate constant of the burst increased with the concentration of rrdI. A deuterium isotope effect (kH/kD approximately 2 at 27 degrees C) was observed when [4B-1H]NADPH was replaced with [4B-2H]NADPH. The characteristics of the burst of transhydrogenation with rrdI:hsdIII differed from those previously reported for rrdI:rrdIII (J.D. Venning et al., Eur. J. Biochem. 257 (1998) 202-209), but the differences are readily explained by a greater dissociation constant of the hybrid complex. The steady-state rate of reverse transhydrogenation by the rrdI:hsdIII complex was almost independent of pH, but there was a single apparent pKa ( approximately 9.1) associated with the cyclic reaction. The reactions of the dI:dIII complex probably proceed independently of those protonation/deprotonation reactions which, in the complete enzyme, are associated with H+ translocation.  相似文献   

5.
Hydroxysteroid dehydrogenases (HSDs) are essential for the biosynthesis and mechanism of action of all steroid hormones. We report the complete kinetic mechanism of a mammalian HSD using rat 3alpha-HSD of the aldo-keto reductase superfamily (AKR1C9) with the substrate pairs androstane-3,17-dione and NADPH (reduction) and androsterone and NADP(+) (oxidation). Steady-state, transient state kinetics, and kinetic isotope effects reconciled the ordered bi-bi mechanism, which contained 9 enzyme forms and permitted the estimation of 16 kinetic constants. In both reactions, loose association of the NADP(H) was followed by two conformational changes, which increased cofactor affinity by >86-fold. For androstane-3,17-dione reduction, the release of NADP(+) controlled k(cat), whereas the chemical event also contributed to this term. k(cat) was insensitive to [(2)H]NADPH, whereas (D)k(cat)/K(m) and the (D)k(lim) (ratio of the maximum rates of single turnover) were 1.06 and 2.06, respectively. Under multiple turnover conditions partial burst kinetics were observed. For androsterone oxidation, the rate of NADPH release dominated k(cat), whereas the rates of the chemical event and the release of androstane-3,17-dione were 50-fold greater. Under multiple turnover conditions full burst kinetics were observed. Although the internal equilibrium constant favored oxidation, the overall K(eq) favored reduction. The kinetic Haldane and free energy diagram confirmed that K(eq) was governed by ligand binding terms that favored the reduction reactants. Thus, HSDs in the aldo-keto reductase superfamily thermodynamically favor ketosteroid reduction.  相似文献   

6.
Clark DD  Allen JR  Ensign SA 《Biochemistry》2000,39(6):1294-1304
The bacterial metabolism of propylene proceeds by epoxidation to epoxypropane followed by carboxylation to acetoacetate. Epoxypropane carboxylation is a minimetabolic pathway that requires four enzymes, NADPH, NAD(+), and coenzyme M (CoM; 2-mercaptoethanesulfonate) and occurs with the overall reaction stoichiometry: epoxypropane + CO(2) + NADPH + NAD(+) + CoM --> acetoacetate + H(+) + NADP(+) + NADH + CoM. The terminal enzyme of the pathway is NADPH:2-ketopropyl-CoM [2-(2-ketopropylthio)ethanesulfonate] oxidoreductase/carboxylase (2-KPCC), an FAD-containing enzyme that is a member of the NADPH:disulfide oxidoreductase family of enzymes and that catalyzes the reductive cleavage and carboxylation of 2-ketopropyl-CoM to form acetoacetate and CoM according to the reaction: 2-ketopropyl-CoM + NADPH + CO(2) --> acetoacetate + NADP(+) + CoM. In the present work, 2-KPCC has been characterized with respect to the above reaction and four newly discovered partial reactions of relevance to the catalytic mechanism, and each of which requires the formation of a stabilized enolacetone intermediate. These four reactions are (1) NADPH-dependent cleavage and protonation of 2-ketopropyl-CoM to form NADP(+), CoM, and acetone, a reaction analogous to the physiological reaction but in which H(+) is the electrophile; (2) NADP(+)-dependent synthesis of 2-ketopropyl-CoM from CoM and acetoacetate, the reverse of the physiologically important forward reaction; (3) acetoacetate decarboxylation to form acetone and CO(2); and (4) acetoacetate/(14)CO(2) exchange to form (14)C(1)-acetoacetate and CO(2). Acetoacetate decarboxylation and (14)CO(2) exchange occurred independent of NADP(H) and CoM, demonstrating that these substrates are not central to the mechanism of enolate generation and stabilization. 2-KPCC did not uncouple NADPH oxidation or NADP(+) reduction from the reactions involving cleavage or formation of 2-ketopropyl-CoM. N-Ethylmaleimide inactivated the reactions forming/using 2-ketopropyl-CoM but did not inactivate acetoacetate decarboxylation or (14)CO(2) exchange reactions. The biochemical characterization of 2-KPCC and the associated five catalytic activities has allowed the formulation of an unprecedented mechanism of substrate activation and carboxylation that involves NADPH oxidation, a redox active disulfide, thiol-mediated reductive cleavage of a C-S thioether bond, the formation of a CoM:cysteine mixed disulfide, and enolacetone stabilization.  相似文献   

7.
The lpdA (Rv3303c) gene from Mycobacterium tuberculosis encoding a new member of the flavoprotein disulfide reductases was expressed in Escherichia coli, and the recombinant LpdA protein was purified to homogeneity. LpdA is a homotetramer and co-purifies with one molecule of tightly but noncovalently bound FAD and NADP+ per monomer. Although annotated as a probable lipoamide dehydrogenase in M. tuberculosis, LpdA cannot catalyze reduction of lipoyl substrates, because it lacks one of two cysteine residues involved in dithiol-disulfide interchange with lipoyl substrates and a His-Glu pair involved in general acid catalysis. The crystal structure of LpdA was solved by multiple isomorphous replacement with anomalous scattering, which confirmed the absence of these catalytic residues from the active site. Although LpdA cannot catalyze reduction of disulfide-bonded substrates, it catalyzes the NAD(P)H-dependent reduction of alternative electron acceptors such as 2,6-dimethyl-1,4-benzoquinone and 5-hydroxy-1,4-naphthaquinone. Significant primary deuterium kinetic isotope effects were observed with [4S-2H]NADH establishing that the enzyme promotes transfer of the C4-proS hydride of NADH. The absence of an isotope effect with [4S-2H]NADPH, the low Km value of 0.5 microm for NADPH, and the potent inhibition of the NADH-dependent reduction of 2,6-dimethyl-1,4-benzoquinone by NADP+ (Ki approximately 6 nm) and 2'-phospho-ADP-ribose (Ki approximately 800 nm), demonstrate the high affinity of LpdA for 2'-phosphorylated nucleotides and that the physiological substrate/product pair is NADPH/NADP+ rather than NADH/NAD+. Modeling of NADP+ in the active site revealed that LpdA achieves the high specificity for NADP+ through interactions involving the 2'-phosphate of NADP+ and amino acid residues that are different from those in glutathione reductase.  相似文献   

8.
Rapid reaction studies presented herein show that ferredoxin:NADP+ oxidoreductase (FNR, EC 1.18.1.2) catalyzes electron transfer from spinach ferredoxin (Fd) to NADP+ via a ternary complex, Fd X FNR X NADP+. In the absence of NADP+, reduction of ferredoxin:NADP+ reductase by Fd was much slower than the catalytic rate: 37-80 s-1 versus at least 445 e-s-1; dissociation of oxidized spinach ferredoxin (Fdox) from one-electron reduced ferredoxin:NADP+ reductase (FNRsq) limited the reduction of FNR. This confirms the steady-state kinetic analysis of Masaki et al. (Masaki, R., Yoshikaya, S., and Matsubara, H. (1982) Biochim. Biophys. Acta 700, 101-109). Occupation of the NADP+ binding site of FNR by NADP+ or by 2',5'-ADP (a nonreducible NADP+ analogue) greatly increased the rate of electron transfer from Fd to FNR, releiving inhibition by Fdox. NADP+ (and 2',5'-ADP) probably facilitate the dissociation of Fdox; equilibrium studies have shown that nucleotide binding decreases the association of Fd with FNR (Batie, C. J. (1983) Ph.D. dissertation, Duke University; Batie, C. J., and Kamin, H. (1982) in Flavins and Flavoproteins VII (Massey, V., and Williams, C. H., Jr., eds) pp. 679-683, Elsevier, New York; Batie, C.J., and Kamin, H. (1982) Fed. Proc. 41, 888; and Batie, C.J., and Kamin, H. (1984) J. Biol. Chem. 259, 8832-8839). Premixing Fd with FNR was found to inhibit the reaction of the flavoprotein with NADP+ and with NADPH; thus, substrate binding may be ordered, NADP+ first, then Fd. FNRred and NADP+ very rapidly formed an FNRred X NADP+ complex with flavin to nicotinamide charge transfer bands. The Fdred X NADP+ complex then relaxed to an equilibrium species; the spectrum indicated a predominance of FNRox X NADPH charge-transfer complex. However, charge-transfer species were not observed during turnover; thus, their participation in catalysis of electron transfer from Fd to NADP+ remains uncertain. The catalytic rate of Fd to NADP+ electron transfer, as well as the rates of electron transfer from Fd to FNR, and from FNR to NADP+ were decreased when the reactants were in D2O; diaphorase activity was unaffected by solvent. On the basis of the data presented, a scheme for the catalytic mechanism of catalysis by FNR is presented.  相似文献   

9.
M Yamaguchi  Y Hatefi 《Biochemistry》1989,28(14):6050-6056
The mitochondrial nicotinamide nucleotide transhydrogenase is a dimeric enzyme of monomer Mr 110,000. It is located in the inner mitochondrial membrane and catalyzes hydride ion transfer between NAD(H) and NADP(H) in a reaction that is coupled to proton translocation across the inner membrane. The amino acid sequence and the nucleotide binding sites of the enzyme have been determined [Yamaguchi, M., Hatefi, Y., Trach, K., & Hoch, J.A. (1988) J. Biol. Chem. 263, 2761-2767; Wakabayashi, S., & Hatefi, Y. (1987) Biochem. Int. 15, 915-924]. N-Ethylmaleimide, as well as other sulfhydryl group modifiers, inhibits the transhydrogenase. The presence of NADP in the incubation mixture suppressed the inhibition rate by N-ethylmaleimide, and the presence of NADPH greatly increased it. NAD and NADH had little or no effect. The NADPH effect was concentration dependent and saturable, with a half-maximal NADPH concentration effect close to the Km of the enzyme for NADPH. Study of the effect of pH on the N-ethylmaleimide inhibition rate showed that NADPH binding by the enzyme lowers the apparent pKa of the N-ethylmaleimide-sensitive group by 0.4 of a pH unit and NADP binding raises this pKa by 0.4 of a pH unit, thus providing a rationale for the effects of NADP and NADPH on the N-ethylmaleimide inhibition rate. With the use of N-[3H]ethylmaleimide, the modified sulfhydryl group involved in the NADP(H)-modulated inhibition of the transhydrogenase was identified as that belonging to Cys-893, which is located 113 residues upstream of the tyrosyl residue modified by [p-(fluorosulfonyl)benzoyl]-5'-adenosine at the putative NADP(H) binding site of the enzyme (see above references).(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

10.
A kinetic scheme is presented for Lactobacillus casei dihydrofolate reductase that predicts steady-state kinetic parameters. This scheme was derived from measuring association and dissociation rate constants and pre-steady-state transients by using stopped-flow fluorescence and absorbance spectroscopy. Two major features of this kinetic scheme are the following: (i) product dissociation is the rate-limiting step for steady-state turnover at low pH and follows a specific, preferred pathway in which tetrahydrofolate (H4F) dissociation occurs after NADPH replaces NADP+ in the ternary complex; (ii) the rate constant for hydride transfer from NADPH to dihydrofolate (H2F) is rapid (khyd = 430 s-1), favorable (Keq = 290), and pH dependent (pKa = 6.0), reflecting ionization of a single group. Not only is this scheme identical in form with the Escherichia coli kinetic scheme [Fierke et al. (1987) Biochemistry 26, 4085] but moreover none of the rate constants vary by more than 40-fold despite there being less than 30% amino acid homology between the two enzymes. This similarity is consistent with their overall structural congruence. The role of Trp-21 of L. casei dihydrofolate reductase in binding and catalysis was probed by amino acid substitution. Trp-21, a strictly conserved residue near both the folate and coenzyme binding sites, was replaced by leucine. Two major effects of this substitution are on (i) the rate constant for hydride transfer which decreases 100-fold, becoming the rate-limiting step in steady-state turnover, and (ii) the affinities for NADPH and NADP+ which decrease by approximately 3.5 and approximately 0.5 kcal mol-1, respectively.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

11.
An analysis of the kinetic mechanism of the microsomal NADPH-linked progesterone 5 alpha-reductase obtained from female rat anterior pituitaries was performed. Initial velocity, product inhibition and dead-end inhibition studies indicate that the kinetic mechanism for the progesterone 5 alpha-reductase is equilibrium ordered sequential. Analysis of the initial velocity data resulted in intersecting double reciprocal plots suggesting a sequential mechanism [apparent Km(progesterone) = 88.2 +/- 8.2 nM; apparent Kia(NADPH) = 7.7 +/- 1.1 microM]. Furthermore, the plot of 1/v vs 1/progesterone intersected on the ordinate which is indicative of an equilibrium ordered mechanism. Additional support for ordered substrate binding was provided by the product inhibition studies with NADPH versus NADP and progesterone versus NADP. NADP is a competitive inhibitor versus NADPH (apparent Kis = 7.8 +/- 1.0 microM) and a noncompetitive inhibitor versus progesterone (apparent Kis = 9.85 +/- 2.1 microM and apparent Kii = 63.2 +/- 12.5 microM). These inhibition patterns suggest that NADPH binds prior to progesterone. In sum, these kinetic studies indicate that NADPH binds to the microsomal enzyme in rapid equilibrium and preferentially precedes the binding of progesterone.  相似文献   

12.
The latter stages of the catalytic cycle of the light-driven enzyme, protochlorophyllide oxidoreductase, have been investigated using novel laser photoexcitation methods. The formation of the ternary product complex was initiated with a 6-ns laser pulse, which allowed the product release steps to be kinetically accessed for the first time. Subsequent absorbance changes associated with the release of the NADP+ and chlorophyllide products from the enzyme could be followed on a millisecond timescale. This has facilitated a detailed kinetic and thermodynamic characterization for the interconversion of all the various bound and unbound product species. Initially, NADP+ is released from the enzyme in a biphasic process with rate constants of 1210 and 237 s(-1). The rates of both phases show a significant dependence on the viscosity of the solvent and become considerably slower at higher glycerol concentrations. The fast phase of this process exhibits no dependence on NADP+ concentration, suggesting that conformational changes are required prior to NADP+ release. Following NADP+ release, the NADPH rebinds to the enzyme with a maximum rate constant of approximately 72 s(-1). At elevated temperatures (>298 K) chlorophyllide is released from the enzyme to yield the free product with a maximum rate constant of 20 s(-1). The temperature dependencies of the rates of each of these steps were measured, and enthalpies and entropies of activation were calculated using the Eyring equation. A comprehensive kinetic and thermodynamic scheme for these final stages of the reaction mechanism is presented.  相似文献   

13.
Previous studies have shown that the interaction of P450 reductase with bound NADP(H) is essential to ensure fast electron transfer through the two flavin cofactors. In this study we investigated in detail the interaction of the house fly flavoprotein with NADP(H) and a number of nucleotide analogues. 1,4,5,6-Tetrahydro-NADP, an analogue of NADPH, was used to characterize the interaction of P450 reductase with the reduced nucleotide. This analogue is inactive as electron donor, but its binding affinity and rate constant of release are very close to those for NADPH. The 2'-phosphate contributes about 5 kcal/mol of the binding energy of NADP(H). Oxidized nicotinamide does not interact with the oxidized flavoprotein, while reduced nicotinamide contributes 1.3 kcal/mol of the binding energy. Oxidized P450 reductase binds NADPH with a K(d) of 0.3 microM, while the affinity of the reduced enzyme is considerably lower, K(d) = 1.9 microM. P450 reductase catalyzes a transhydrogenase reaction between NADPH and oxidized nucleotides, such as thionicotinamide-NADP(+), acetylpyridine-NADP(+), or [(3)H]NADP(+). The reverse reaction, reduction of [(3)H]NADP(+) by the reduced analogues, is also catalyzed by P450 reductase. We define the mechanism of the transhydrogenase reaction as follows: NADPH binding, hydride ion transfer, and release of the NADP(+) formed. An NADP(+) or its analogue binds to the two-electron-reduced flavoprotein, and the electron-transfer steps reverse to transfer hydride ion to the oxidized nucleotide, which is released. Measurements of the flavin semiquinone content, rate constant for NADPH release, and transhydrogenase turnover rates allowed us to estimate the steady-state distribution of P450 reductase species during catalysis, and to calculate equilibrium constants for the interconversion of catalytic intermediates. Our results demonstrate that equilibrium redox potentials of the flavin cofactors are not the sole factor governing rapid electron transfer during catalysis, but conformational changes must be considered to understand P450 reductase catalysis.  相似文献   

14.
Recombinant house fly (Musca domestica) cytochrome P450 reductase has been purified by anion exchange and affinity chromatography. Steady-state kinetics of cytochrome c reductase activity revealed a random Bi-Bi mechanism with formation of a ternary P450 reductase-NADPH-electron acceptor complex as catalytic intermediate. NADP(H) binding is essential for fast hydride ion transfer to FAD, as well as for electron transfer from FMN to cytochrome c. Reduced cytochrome c had no effect on the enzyme activity, while NADP+ and 2'-AMP inhibited P450 reductase competitively with respect to NADPH and noncompetitively with respect to cytochrome c. The affinity of the P450 reductase to NADPH is 10 times higher than to NADP+ (Kd of 0.31 and 3.3 microM, respectively). Such an affinity change during catalysis could account for a +30 mV shift of the redox potential of FAD. Cys560 was substituted for Tyr by site-directed mutagenesis. This mutation decreased enzyme affinity to NADPH 35-fold by decreasing the bimolecular rate constant of nucleotide binding with no detectable effect on the kinetic mechanism. The affinity of the C560Y mutant enzyme to NADP+ decreased 9-fold compared to the wild-type enzyme, while the affinity to 2'-AMP was not significantly affected, suggesting that Cys560 is located in the nicotinamide binding site of the active, full-size enzyme in solution.  相似文献   

15.
In cell extracts of Methanosarcina barkeri, the methylcoenzyme M methylreductase system with H2 as the electron donor was inhibited by NAD+ and NADP+, but NADH and NADPH had no effect on enzyme activity. NAD+ (4 and 8 mM) shifted the saturation curve for methylcoenzyme M from hyperbolic (Hill coefficient [nH] = 1.0; concentration of substrate giving half maximal velocity [Km] = 0.21 mM) to sigmoidal (nH = 1.5 and 2.0), increased Km (Km = 0.25 and 0.34 mM), and slightly decreased Vmax. Similarly NADP+ at 4m and 8 mM increased nH to 1.6 and 1.85 respectively, but the Km values (0.3 and 0.56 mM) indicated that NADP+ was a more efficient inhibitor than NAD+.  相似文献   

16.
Reaction of glycolaldehyde with the binary E-NADP complex of bovine kidney aldose reductase (ALR2) produces an enzyme-bound chromophore whose absorbance (lambd max 341 nm) and fluorescence (lambda ex max 341 nm; lambda emit max 421 nm) properties are distinct from those of NADPH or E.NADPH yet are consistent with the proposed covalent adduct structure [1,4-dihydro-4-(1-hydroxy-2-oxoethyl)nicotinamide adenine dinucleotide phosphate]. The kinetics of adduct formation, both in solution and at the enzyme active site, support a mechanism involving rate-determining enolization of glycolaldehyde at high [NADP+] or [E.NADP]. At low [NADP+] or [E.NADP] the reaction is second-order overall, but the ALR2-mediated reaction displays saturation by glycolaldehyde due to competition of the aldehyde (plus hydrate) and enol for E.NADP. Measurement of the pre-steady-state burst of E-adduct formation confirms that glycolaldehyde enol is the reactive species and gives a value of 1.3 x 10(-6) for Kenol = [enol]/[( aldehyde] + [hydrate]), similar to that determined by trapping the enol with I3-. At the ALR2 active site, the rate of adduct formation is enhanced 79,000-fold and the adduct is stabilized greater than or equal to 13,000-fold relative to the reaction with NADP+ in solution. A portion of this enhancement is ascribed to specific interaction of NADP+ with the enzyme since the 3-acetylpyridine analogue, (AP)ADP+, gives values that are 15-200-fold lower. Additional evidence for strong interaction of ALR2 with both NADP+ and NADPH is reported. Yet, because dissociation of adduct is slow, catalysis of the overall adduct formation reaction by ALR2 is less than or equal to 67-fold.  相似文献   

17.
The catalytic subunit of the cAMP-dependent protein kinase from bovine cardiac muscle phosphorylates homoserine in the synthetic peptide Leu-Arg-Arg-Ala-Hse-Leu-Gly. Phosphorylation of the primary alcohol of the homoserine residue was established via NMR spectroscopy. Two-dimensional correlated and nuclear Overhauser effect spectroscopies provided the sequence-specific chemical shift assignments of the substrate peptide and its phosphorylated counterpart. Coupled and decoupled 31P NMR experiments established the presence of phosphate on the homoserine residue. The maximal velocity (6.4 mumol/min.mg) obtained for homoserine-peptide phosphorylation at 12.5 mM Mg2+ compares favorably to the velocities observed for the corresponding serine- (21 mumol/min.mg), threonine- (3.2 mumol/min.mg), and hydroxyproline-peptides (1 mumol/min.mg). However, the Km for homoserine kinase activity is modest (1.3 mM) relative to the Km associated with the phosphorylation of the serine-containing substrate (22 microM). The effect of Mg2+ concentration on the kinetic parameters kcat, Km, and kcat/Km was investigated for both serine- and homoserine-peptides. Both substrates display similar kcat/Km versus [Mg2+] profiles, with the most notable difference that the optimal Mg2+ concentration is higher for the homoserine-containing peptide. In addition, the Km for the serine-peptide was found to be independent of [Mg2+], whereas the Km for the homoserine-peptide was observed to be dependent upon [Mg2+]. These results suggest that the long homoserine side chain may induce an unusually large off rate for the peptide and/or may misalign the hydroxyl moiety in the active site.  相似文献   

18.
A NADPH cytochrome c oxidoreductase purified from membranes of rabbit peritoneal neutrophil was shown to behave as the NADPH dehydrogenase component of the O2- generating oxidase complex. A photoactivable derivative of NADP+, azido nitrophenyl-gamma-aminobutyryl NADP+ (NAP4-NADP+), was synthesized in its labeled [3H] form and used to photolabel the NADPH cytochrome c reductase at different stages of the purification procedure. Control assays performed in dim light indicated that the reduced form of NADP4-NADP+ generated by reduction with glucose-6-phosphate and glucose-6-phosphate dehydrogenase was oxidized at virtually the same rate as NADPH. Upon photoirradiation of the purified reductase in the presence of [3H]NAP4-NADP+ and subsequent separation of the photolabeled species by sodium dodecyl sulfate polyacrylamide gel electrophoresis, radioactivity was found to be present predominantly in a protein band with a molecular mass of 77-kDa and accessorily in bands of 67-kDa and 57-kDa. Evidence is provided that the 67-kDa and 57-kDa proteins arose from the 77-kDa protein by proteolysis. Despite removal of part of the sequence, the proteolyzed proteins were still active in catalyzing electron transport from NADPH to cytochrome c and in binding the photoactivable derivative of NADP+.  相似文献   

19.
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   

20.
The reduced glutathione-linked NADP+ reduction, catalyzed by yeast glutathione reductase, follows a 'sequential' or 'ping-pong' mechanism at high or low NADP+ concentrations, respectively. The pattern of the NADPH and NADP+ cross-inhibition reflects not only the competition for the binding site, but the shift of the reaction equilibrium as well. A 'branched' scheme of the glutathione reductase reaction is presented. The enzyme standard potential (-255 mV, pH 7.0) was estimated from the ratio of the NADPH and NADP+ rate constants corresponding to the ping-pong 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