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.
P T Tuazon  S L Johnson 《Biochemistry》1977,16(6):1183-1188
1,4-Dihydronicotinamide adenine dinucleotide (NADH) and its analogues undergo two reactions in sulfite buffers in the pH range 5.5-7.1: (1) an oxygen-mediated free-radical chain reaction which results in the oxidation of the dihydropyridine to the pyridinium salt, and (2) an ionic reaction which results in the hydration of the 5,6 double bond of the dihydropyridine. The free-radical reaction is inhibited by superoxide dismutase (indicating the involvement of superoxide radicals) and by free-radical inhibitors. The ionic reaction is not affected by free-radical inhibitors and follows the rate law: rate = [substrate] [HSO3-] (K + SIGMAK' [HA]), where HA is a general acid of hydronium ion. The occurrence of third-order terms of the type [substrate] X [HSO3-] [HA] is consistent with the formation of a reactive bisulfite-substrate complex, which undergoes general acid catalyzed hydration.  相似文献   

2.
1,4,5,6-Tetrahydronicotinamide adenine dinucleotide (H2NADH) has been investigated as a reduced coenzyme analog in the reaction between trans-4-N,N-dimethylaminocinnamaldehyde (I) (lambdamax 398 nm, epsilonmax 3.15 X 10-4 M-minus 1 cm-minus 1) and the horse liver alcohol dehydrogenase-NADH complex. These equilibrium binding and temperature-jump kinetic studies establish the following. (i) Substitution of H2NADH for NADH limits reaction to the reversible formation of a new chromophoric species, lambdamax 468 nm, epsilonmax 5.8 x 10-4 M-minus 1 cm-minus 1. This chromophore is demonstrated to be structurally analogous to the transient intermediate formed during the reaction of I with the enzyme-NADH complex [Dunn, M. F., and Hutchison, J. S. (1973), Biochemistry 12, 4882]. (ii) The process of intermediate formation with the enzyme-NADH complex is independent of pH over the range 6.13-10.54. Although studies were limited to the pH range 5.98-8.72, a similar pH independence appears to hold for the H2NADH system. (iii) Within the ternary complex, I is bound within van der Waal's contact distance of the coenzyme nicotinamide ring. (iv) Formation of the transient intermediate does not involve covalent modification of coenzyme. Based on these findings, we conclude that zinc ion has a Lewis acid function in facilitating the chemical activation of the aldehyde carbonyl for reduction, and that reduced coenzyme plays a noncovalent effector role in this substrate activating step.  相似文献   

3.
K H Dahl  M F Dunn 《Biochemistry》1984,23(26):6829-6839
Liver alcohol dehydrogenase (LADH) carboxymethylated at Cys-46 (CMLADH) forms two different ternary complexes with 4-trans-(N,N-dimethylamino)cinnamaldehyde (DACA). The complex with reduced nicotinamide adenine dinucleotide (NADH) is characterized by a 38-nm red shift of the long-wavelength pi, pi* transition to 436 nm, while the complex with oxidized nicotinamide adenine dinucleotide (NAD+) is characterized by a 60-nm red shift to 458 nm. CMLADH also forms a ternary complex with NAD+ and the Z isomer of 4-trans-(N,N-dimethylamino)cinnamaldoxime in which the absorption of the oxime (lambda max = 354 nm) is red shifted 80 nm to 434 nm. Pyrazole and 4-methylpyrazole are weak competitive inhibitors of ligand binding to the substrate site of native LADH. These inhibitors were found to form ternary complexes with CMLADH and NADH which are more stable than the corresponding complexes with the native enzyme. The transient reductions of the aldehydes DACA and p-nitrobenzaldehyde (NBZA) were studied under single-turnover conditions. Carboxymethylation decreases the DACA reduction rate 80-fold and renders the process essentially independent of pH over the region 5-9, whereas this process depends on a pKa of 6.0 in the native enzyme. At pH 7.0, the rate constant for NBZA reduction also is decreased at least 80-fold to a value of 7.7 +/- 0.3 s-1. Since primary kinetic isotope effects are observed when NADH is substituted with (4R)-4-deuterio-NADH (kH/kD = 3.0 for DACA and kH/kD = 2.3 for NBZA), the rate-limiting step for both aldehydes involves hydride transfer. The altered pH dependence is concluded to be due to an increase in the pK value of the zinc-coordinated DACA-alcohol in the ternary complex with NAD+ by more than 3 units. This perturbation is brought about by the close proximity of the negatively charged carboxymethyl carboxylate.  相似文献   

4.
Acholeplasma laidlawii A possesses a nicotinamide adenine dinucleotide (NAD)-dependent l(+)-lactate dehydrogenase (LDH) which is activated specifically by low concentrations of fructose-1, 6-diphosphate (FDP). Studies with partially purified enzyme show that the kinetic response to FDP is hyperbolic. The enzyme is inhibited by inorganic phosphate, adenosine triphosphate, and high concentrations of reduced NAD (NADH). Low activity is demonstrable in the absence of FDP at pH 6.0 to 7.2, but FDP is absolutely required in the region of pH 8. FDP causes an upward shift in the optimum pH of the enzyme, which is near 7.2 in tris (hydroxymethyl)aminomethane buffer. Activation of the enzyme by FDP is markedly affected by substrate concentration; FDP lowers the apparent K(m) for pyruvate and NADH. The affinity of the enzyme for pyruvate is also influenced by H(+) concentration. The pyruvate analogue alpha-ketobutyrate serves as an effective substrate for the enzyme; when it is utilized, the enzyme is still activated by FDP. Reversal of the pyruvate reduction reaction catalyzed by the enzyme can be demonstrated with the 3-acetylpyridine analogue of NAD. The catalytic properties of the A. laidlawii enzyme and the known FDP-activated LDHs which occur among lactic acid bacteria are discussed.  相似文献   

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

6.
The effect of pH on steady state kinetic parameters for the yeast alcohol dehydrogenase-catalyzed reduction of aldehydes and oxidation of alcohols has been studied. The oxidation of p-CH3 benzyl alcohol-1,1-h2 and -1,1-d2 by NAD+ was found to be characterized by large deuterium isotope effects (kH/kD = 4.1 plus or minus 0.1) between pH 7.5 and 9.5, indicating a rate-limiting hydride trahsfer step in this pH range; a plot of kCAT versus pH could be fit to a theoretical titration curve, pK = 8.25, where kCAT increases with increasing pH. The Michaelis constnat for p-CH3 benzyl alcohol was independent of pH. The reduction of p-CH3 benzaldehyde by NADH and reduced nicotinamide adenine dinucleotide with deuterium in the 4-A position (NADD) cound not be studied below pH 8.5 due to substrate inhibition; however, between pH 8.5 and 9.5, kCAT was found to decrease with increasing pH and to be characterized by significant isotope effects (kH/kD = 3.3 plus or minus 0.3). In the case of acetaldehyde reduction by NADH and NADD, isotope effects were found to be small and exxentially invariant (kH/kD = 2.O plus or minus 0.4) between pH 7.2 and 9.5, suggesting a partially rate-limiting hydride transger step for this substrate; a plot of kCAT/K'b (where K'b is the Michaelis constant for acetaldehyde) versus pH could be fit to a titration curve, pK = 8.25. The titration curve for acetaldehyde reduction has the same pK but is opposite in direction to that observed for p-CH3 benzyl alcohol oxidation. The data presented in this paper indicate a dependence on different enzyme forms for aldehyde reduction and alcohol oxidation and are consistent with a single active site side chain, pK = 8.25, which functions in acid-base catalysis of the hydride transfer step.  相似文献   

7.
The pH dependence of the 13C chemical shifts for nicotinamide adenine dinucleotide (NAD+), thionicotinamide adenine dinucleotide (TNAD+), pyridine adenine dinucleotide (PyrAD+), N-methyl-nicotinamide adenine dinucleotide (N-Me-NAD+), acetylpyridine adenine dinucleotide (AcPyAD+), nicotinamide hypoxanthine dinucleotide (NHD+), and nicotinamide adenine dinucleotide phosphate (NADP+) are reported. In these analogs the 13C chemical shifts of the pyridinium moiety reflect the pKa of the opposing purine base, while the 13C chemical shift dependence on pD for the pyridinium carbons of nicotinamide mononucleotide (NMN+) and adenosine monophosphate (AMP), 1,4-dihydronicotinamide adenine dinucleotide (NADH), 1,4-dihydronicotinamide adenine dinucleotide phosphate (NADPH), and nicotinic acid adenine dinucleotide (N(a)AD+) are not influenced by the adenine ring in the pD range tested. Through the use of 13C-labeled NAD+, the source of the pH dependence of the 13C chemical shifts was shown to be intramolecular in origin. However, serious doubt is cast on the utility of employing the pD dependence of chemical shift data to determine the nature of solution conformers or their relative populations.  相似文献   

8.
Lineweaver-Burk plots of reduced nicotinamide adenine dinucleotide (NADH) oxidation by membrane preparations from Bacillus subtilis are biphasic, with two K(m) values for NADH. The higher K(m) corresponds to the only K(m) observed for NADH oxidation by whole cells, whereas the lower K(m) corresponds to that observed with open cell envelopes. Membrane preparations apparently contain a small fraction of open or inverted vesicles which is responsible for the low K(m) reaction, whereas entry of NADH into the larger portion of closed, normally oriented vesicles is rate limiting and responsible for the high K(m) reaction. In contrast, the oxidation of l-alpha-glycerol-phosphate (glycerol-P) by membrane preparations shows only one K(m) that corresponds to that of glycerol-P oxidation by whole cells or lysates. Since glycerol-P dehydrogenase (NAD independent) has the same K(m), this enzyme reaction rather than entry of glycerol-P into vesicles represents the rate-limiting step for glycerol-phosphate oxidation. The K(m) for amino acid uptake by vesicles in the presence of NADH corresponds to the high K(m) for NADH oxidation, indicating that NADH energizes transport only if it enters closed, normally oriented vesicles. Studies with rotenone and proteolytic enzymes support this interpretation. The apparent efficiency of NADH in energizing uptake seems to be lower than that of glycerol-P because, under the experimental conditions usually employed, open or inverted vesicles that do not participate in amino acid uptake are responsible for the major portion of NADH oxidation. When the results are corrected for this effect, the efficiency of NADH is essentially the same as that of l-alpha-glycerol-P.  相似文献   

9.
Yang X  Ma K 《Analytical biochemistry》2005,344(1):130-134
Hydrogen peroxide can be conveniently determined using horseradish peroxidase (HRP) and 2,2'-azino-bis(3-ethylbenzthiazoline-6-sulfonic acid). However, interference occurs among assay components in the presence of reduced nicotinamide adenine dinucleotide (NADH) that is also a substrate of NADH oxidase. So, depletion of NADH is required before using the HRP method. Here, we report simple and rapid procedures to accurately determine hydrogen peroxide generated by NADH oxidase. All procedures developed were based on the extreme acid lability of NADH and the stability of hydrogen peroxide, because NADH was decomposed at pH 2.0 or 3.0 for 10 min, while hydrogen peroxide was stable at pH 2.0 or 3.0 for at least 60 min. Acidification and neutralization were carried out by adjusting sample containing NADH up to 30 microM to pH 2.0 for 10 min before neutralizing it back to pH 7.0. Then, hydrogen peroxide in the sample was measured using the HRP method and its determination limit was found to be about 0.3 microM. Alternatively, hydrogen peroxide in samples containing NADH up to 100 microM could be quantitated using a modified HRP method that required an acidification step only, which was found to have a determination limit of about 3 microM hydrogen peroxide in original samples.  相似文献   

10.
The L-(+)-lactate dehydrogenase (L-lactate:NAD+ oxidoreductase, EC 1.1.1.27) of Streptococcus lactis C10, like that of other streptococci, was activated by fructose 1,6-diphosphate (FDP). The enzyme showed some activity in the absence of FDP, with a pH optimum of 8.2; FDP decreased the Km for both pyruvate and reduced nicotinamide adenine dinucleotide (NADH) and shifted the pH optimum to 6.9. Enzyme activity showed a hyperbolic response to both NADH and pyruvate in all the buffers tried except phosphate buffer, in which the response to increasing NADH was sigmoidal. The FDP concentration required for half-maximal velocity (FDP0.5V) was markedly influenced by the nature of the assay buffer used. Thus the FDP0.5V was 0.002 mM in 90 mM triethanolamine buffer, 0.2 mM in 90 mM tris(hydroxymethyl)aminomethanemaleate buffer, and 4.4 mM in 90 mM phosphate buffer. Phosphate inhibition of FDP binding is not a general property of streptococcal lactate dehydrogenase, since the FDP0.5V value for S. faecalis 8043 lactate dehydrogenase was not increased by phosphate. The S. faecalis and S. lactis lactate dehydrogenases also differed in that Mn2+ enhanced FDP binding in S. faecalis but had no effect on the S. lactis dehydrogenase. The FDP concentration (12 to 15 mM) found in S. lactis cells during logarithmic growth on a high-carbohydrate (3% lactose) medium would be adequate to give almost complete activation of the lactate dehydrogenase even if the high FDP0.5V value found in 90 mM phosphate were similar to the FDP requirement in vivo.  相似文献   

11.
The 13C spectra of β-NADH, NADHX, and the primary acid product of NADH were obtained and assigned. The conversion of the NADHX isomers to the two isomers of NADH acid product is demonstrated through the use of 13C-enriched compounds. The structure of NADHX is assigned as β-6-hydroxy-1,4,5,6-tetrahydronicotinamide adenine dinucleotide and the structures of the primary acid products of NADH are assigned as α-O2′-6B-cyclotetrahydronicotinamide adenine dinucleotide and α-O2′-6A-cyclotetrahydronicotinamide adenine dinucleotide.The structures of NADHX and the major isomer of the primary acid product, derived from studies of model compounds, are consistent with those proposed by Oppenheimer and Kaplan [Biochemistry (1974) 13, 4675, 4685]. However, the spectra of 13C-enriched primary acid product also demonstrated the existence of the A isomer which was not observed in the latter 1H study. The A and B isomers were found to exist in the same ratio even when the primary acid product was formed directly from NADHX. This observation is discussed in terms of the previously proposed mechanism for the acid decomposition of NADH.  相似文献   

12.
Nicotinamide adenine dinucleotide (NAD) dependent urocanase (4'-imidazolone-5'-propionate hydro-lyase, EC 4.2.1.49) from Pseudomonas putida was found to catalyze an exchange reaction between solvent and the 4'-hydrogen of urocanate or imidazolepropionate at a rate faster than that of overall deuterium was compared to unlabeled urocanate as a substrate, no isotope rate effect was noted. For examination of the possibility of an NAD+-mediated intramolecular hydride transfer of the 4'-hydrogen to a position on the side chain of oxoimidazolepropionate, the origins of hydrogen at positions 2 and 3 in the propionate chain were studied as a function of reaction time and extent of exchange of the 4'-hydrogen. No transfer of hydrogen from the 4' position to the side chain was observed, thereby eliminating mechanisms requiring hydride transfer via NADH between these positions. Catalytic rates in 1H2O vs. 2H2O revealed a 3-fold difference which was ascribed to a rate-limiting proton addition step. Similarly, a 5-fold decrease in Vmax was found for the reverse reaction when oxoimidazole[2,3-2H2]propionate was compared to unlabeled oxoimidazolepropionate. These data support a mechanism involving water addition across the conjugated double bond system of urocanate, rather than an internal oxidation--reduction process, yet NAD+ is required. A mechanism is proposed which uses electron delocalization in the imidazole nucleus, via an imidazole--NAD adduct, to facilitate water attack and subsequent formation of oxoimidazolepropionate.  相似文献   

13.
3alpha-hydroxysteroid dehydrogenase/carbonyl reductase from Comamonas testosteroni catalyzes the oxidation of androsterone with NAD(+) to form androstanedione and NADH with a concomitant releasing of protons to bulk solvent. To probe the proton transfer during the enzyme reaction, we used mutagenesis, chemical rescue, and kinetic isotope effects to investigate the release of protons. The kinetic isotope effects of (D)V and (D(2)O)V for wild-type enzyme are 1 and 2.1 at pL 10.4 (where L represents H, (2)H), respectively, and suggest a rate-limiting step in the intramolecular proton transfer. Substitution of alanine for Lys(159) changes the rate-limiting step to the hydride transfer, evidenced by an equal deuterium isotope effect of 1.8 on V(max) and V/K(androsterone) and no solvent kinetic isotope effect at saturating 3-(cyclohexylamino)propanesulfonic acid (CAPS). However, a value of 4.4 on V(max) is observed at 10 mm CAPS at pL 10.4, indicating a rate-limiting proton transfer. The rate of the proton transfer is blocked in the K159A and K159M mutants but can be rescued using exogenous proton acceptors, such as buffers, small primary amines, and azide. The Br?nsted relationship between the log(V/K(d)(-base)Et) of the external amine (corrected for molecular size effects) and pK(a) is linear for the K159A mutant-catalyzed reaction at pH 10.4 (beta = 0.85 +/- 0.09) at 5 mm CAPS. These results show that proton transfer to the external base with a late transition state occurred in a rate-limiting step. Furthermore, a proton inventory on V/Et is bowl-shaped for both the wild-type and K159A mutant enzymes and indicates a two-proton transfer in the transition state from Tyr(155) to Lys(159) via 2'-OH of ribose.  相似文献   

14.
M Fujioka  Y Takata 《Biochemistry》1981,20(3):468-472
The baker's yeast saccharopine dehydrogenase (EC 1.5.1.7) was inactivated by 2,3-butanedione following pseudo-first-order reaction kinetics. The pseudo-first-order rate constant for inactivation was linearly related to the butanedione concentration, and a value of 7.5 M-1 min-1 was obtained for the second-order rate constant at pH 8.0 and 25 degrees C. Amino acid analysis of the inactivated enzyme revealed that arginine was the only amino acid residue affected. Although as many as eight arginine residues were lost on prolonged incubation with butanedione, only one residue appears to be essential for activity. The modification resulted in the change in Vmax, but not in Km, values for substrates. The inactivation by butanedione was substantially protected by L-leucine, a competitive analogue of substrate lysine, in the presence of reduced nicotinamide adenine dinucleotide (NADH) and alpha-ketoglutarate. Since leucine binds only to the enzyme-NADH-alpha-ketoglutarate complex, the result suggests that an arginine residue located near the binding site for the amino acid substrate is modified. Titration with leucine showed that the reaction of butanedione also took place with the enzyme-NADH-alpha-ketoglutarate-leucine complex more slowly than with the free enzyme. The binding study indicated that the inactivated enzyme still retained the capacity to bind leucine, although the affinity appeared to be somewhat decreased. From these results it is concluded that an arginine residue essential for activity is involved in the catalytic reaction rather than in the binding of the coenzyme and substrates.  相似文献   

15.
Nitric oxide reductase cytochrome P450nor catalyzes an unusual reaction, direct electron transfer from NAD(P)H to bound heme. Here, we succeeded in determining the crystal structure of P450nor in a complex with an NADH analogue, nicotinic acid adenine dinucleotide, which provides conclusive evidence for the mechanism of the unprecedented electron transfer. Comparison of the structure with those of dinucleotide-free forms revealed a global conformational change accompanied by intriguing local movements caused by the binding of the pyridine nucleotide. Arg64 and Arg174 fix the pyrophosphate moiety upon the dinucleotide binding. Stereo-selective hydride transfer from NADH to NO-bound heme was suggested from the structure, the nicotinic acid ring being fixed near the heme by the conserved Thr residue in the I-helix and the upward-shifted propionate side-chain of the heme. A proton channel near the NADH channel is formed upon the dinucleotide binding, which should direct continuous transfer of the hydride and proton. A salt-bridge network (Glu71-Arg64-Asp88) was shown to be crucial for a high catalytic turnover.  相似文献   

16.
Saccharopine dehydrogenase (SDH) catalyzes the final reaction in the α-aminoadipate pathway, the conversion of l-saccharopine to l-lysine (Lys) and α-ketoglutarate (α-kg) using NAD? as an oxidant. The enzyme utilizes a general acid-base mechanism to conduct its reaction with a base proposed to accept a proton from the secondary amine of saccharopine in the oxidation step and a group proposed to activate water to hydrolyze the resulting imine. Crystal structures of an open apo form and a closed form of the enzyme with saccharopine and NADH bound have been determined at 2.0 and 2.2 ? resolution, respectively. In the ternary complex, a significant movement of domain I relative to domain II that closes the active site cleft between the two domains and brings H96 and K77 into the proximity of the substrate binding site is observed. The hydride transfer distance is 3.6 ?, and the side chains of H96 and K77 are properly positioned to act as acid-base catalysts. Preparation of the K77M and H96Q single-mutant and K77M/H96Q double-mutant enzymes provides data consistent with their role as the general acid-base catalysts in the SDH reaction. The side chain of K77 initially accepts a proton from the ε-amine of the substrate Lys and eventually donates it to the imino nitrogen as it is reduced to a secondary amine in the hydride transfer step, and H96 protonates the carbonyl oxygen as the carbinolamine is formed. The K77M, H976Q, and K77M/H96Q mutant enzymes give 145-, 28-, and 700-fold decreases in V/E(t) and >103-fold increases in V?/K(Lys)E(t) and V?/K(α-kg)E(t) (the double mutation gives >10?-fold decreases in the second-order rate constants). In addition, the K77M mutant enzyme exhibits a primary deuterium kinetic isotope effect of 2.0 and an inverse solvent deuterium isotope effect of 0.77 on V?/K(Lys). A value of 2.0 was also observed for (D)(V?/K(Lys))(D?O) when the primary deuterium kinetic isotope effect was repeated in D?O, consistent with a rate-limiting hydride transfer step. A viscosity effect of 0.8 was observed on V?/K(Lys), indicating the solvent deuterium isotope effect resulted from stabilization of an enzyme form prior to hydride transfer. A small normal solvent isotope effect is observed on V, which decreases slightly when repeated with NADD, consistent with a contribution from product release to rate limitation. In addition, V?/K(Lys)E(t) is pH-independent, which is consistent with the loss of an acid-base catalyst and perturbation of the pK(a) of the second catalytic group to a higher pH, likely a result of a change in the overall charge of the active site. The primary deuterium kinetic isotope effect for H96Q, measured in H?O or D?O, is within error equal to 1. A solvent deuterium isotope effect of 2.4 is observed with NADH or NADD as the dinucleotide substrate. Data suggest rate-limiting imine formation, consistent with the proposed role of H96 in protonating the leaving hydroxyl as the imine is formed. The pH-rate profile for V?/K(Lys)E(t) exhibits the pK(a) for K77, perturbed to a value of ~9, which must be unprotonated to accept a proton from the ε-amine of the substrate Lys so that it can act as a nucleophile. Overall, data are consistent with a role for K77 acting as the base that accepts a proton from the ε-amine of the substrate lysine prior to nucleophilic attack on the α-oxo group of α-ketoglutarate, and finally donating a proton to the imine nitrogen as it is reduced to give saccharopine. In addition, data indicate a role for H96 acting as a general acid-base catalyst in the formation of the imine between the ε-amine of lysine and the α-oxo group of α-ketoglutarate.  相似文献   

17.
Xu H  West AH  Cook PF 《Biochemistry》2006,45(39):12156-12166
Kinetic data have been measured for the histidine-tagged saccharopine dehydrogenase from Saccharomyces cerevisiae, suggesting the ordered addition of nicotinamide adenine dinucleotide (NAD) followed by saccharopine in the physiologic reaction direction. In the opposite direction, the reduced nicotinamide adenine dinucleotide (NADH) adds to the enzyme first, while there is no preference for the order of binding of alpha-ketoglutarate (alpha-Kg) and lysine. In the direction of saccharopine formation, data also suggest that, at high concentrations, lysine inhibits the reaction by binding to free enzyme. In addition, uncompetitive substrate inhibition by alpha-Kg and double inhibition by NAD and alpha-Kg suggest the existence of an abortive E:NAD:alpha-Kg complex. Product inhibition by saccharopine is uncompetitive versus NADH, suggesting a practical irreversibility of the reaction at pH 7.0 in agreement with the overall K(eq). Saccharopine is noncompetitive versus lysine or alpha-Kg, suggesting the existence of both E:NADH:saccharopine and E:NAD:saccharopine complexes. NAD is competitive versus NADH, and noncompetitive versus lysine and alpha-Kg, indicating the combination of the dinucleotides with free enzyme. Dead-end inhibition studies are also consistent with the random addition of alpha-Kg and lysine. Leucine and oxalylglycine serve as lysine and alpha-Kg dead-end analogues, respectively, and are uncompetitive against NADH and noncompetitive against alpha-Kg and lysine, respectively. Oxaloacetate (OAA), pyruvate, and glutarate behave as dead-end analogues of lysine, which suggests that the lysine-binding site has a higher affinity for keto acid analogues than does the alpha-Kg site or that dicarboxylic acids have more than one binding mode on the enzyme. In addition, OAA and glutarate also bind to free enzyme as does lysine at high concentrations. Glutarate gives S-parabolic noncompetitive inhibition versus NADH, indicating the formation of a E:(glutarate)2 complex as a result of occupying both the lysine- and alpha-Kg-binding sites. Pyruvate, a slow alternative keto acid substrate, exhibits competitive inhibition versus both lysine and alpha-Kg, suggesting the combination to the E:NADH:alpha-Kg and E:NADH:lysine enzyme forms. The equilibrium constant for the reaction has been measured at pH 7.0 as 3.9 x 10(-7) M by monitoring the change in NADH upon the addition of the enzyme. The Haldane relationship is in very good agreement with the directly measured value.  相似文献   

18.
The kinetic characteristics of the diaphorase activities associated with the NADH-dependent nitrite reductase (EC 1.6.6.4) from Escherichia coli have been determined. The values of the apparent maximum velocity are similar for the reduction of Fe(CN)6(3)-and mammalian cytochrome c by NADH. These reactions may therefore have the same rate-limiting step. NAD+ activates NADH-dependent reduction of cytochrome c, and the apparent maximum velocity for this substrate increases more sharply with the concentration of NAD+ than for hydroxylamine. The simplest explanation is that NAD+ activation of hydroxylamine reduction derives solely from activation of steps involved in the reduction of cytochrome c, a flavin-mediated reaction, but these steps are only partly rate-limiting for the reduction of hydroxylamine. At 0.5 mM-NAD+, the apparent maximum velocity was 2.3 times higher for 0.1 mM-cytochrome c as substrate than for 100 mM-hydroxylamine, suggesting that the rate-limiting step during hydroxylamine reduction is a step that is not involved in cytochrome c reduction. A scheme is proposed that can account for the pattern of variation with [NAD+] of the Michaelis-Menten parameters for hydroxylamine and for NADH with hydroxylamine or cytochrome c as oxidized substrate.  相似文献   

19.
Micrococcus aerogenes grown in media containing glutamate has high levels of glutamate dehydrogenase and alpha-ketoglutarate reductase. The latter enzyme catalyzes the reversible reduction of alpha-ketoglutarate to alpha-hydroxyglutarate in the presence of reduced nicotinamide adenine dinucleotide (NADH). The enzyme has a high specificity for both substrates in either direction and displays Michaelis-Menten kinetics at moderate substrate concentrations. K(m) values of 0.12 to 0.17 mm alpha-ketoglutarate and 0.3 mm NADH for the forward reaction were calculated from data obtained at low substrate concentrations. At high concentrations, this reaction was inhibited by both substrates. The reverse reaction, which proceeded at 0.1 to 0.2 times the rate of the forward reactions, was inhibited by one of the products, alpha-ketoglutarate. K(m) values for the substrates of this reaction were 10 mm for alpha-hydroxyglutarate and 1 mm for nicotinamide adenine dinucleotide. alpha-Ketoglutarate reductase has a molecular weight of 7.5 x 10(4) to 8.2 x 10(4) and is composed of identical polypeptide chains with a molecular weight of 3.6 x 10(4) to 3.8 x 10(4).  相似文献   

20.
A reduced nicotinamide adenine dinucleotide (NADH)-dependent glutamate synthase has been detected and partially purified from crude extracts of Saccharomyces cerevisiae. The enzyme is specific for NADH, glutamine, and alpha-ketoglutarate (K(m) values of 2.6 muM, 1.0 mM, and 140 muM, respectively) and has a pH optimum between 7.1 and 7.7. The stoichiometry of the reaction has been determined as 2 mol of glutamate synthesized per mol of glutamine consumed. Glutamate synthase can be distinguished from either of the glutamate dehydrogenases of yeast on the basis of its substrate requirements and behavior during agarose gel and ion exchange chromatography. Variations in the specific activity of glutamate synthase, which occur in response to changes in the growth medium, are similar in character to those observed with the nicotinamide adenine dinucleotide phosphate-dependent (anabolic) glutamate dehydrogenase.  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