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 46 毫秒
1.
The kinetic mechanism of betaine aldehyde dehydrogenase from leaves of the plant Amaranthus hypochondriacus is ordered with NAD(+) adding first. NADH is a noncompetitive inhibitor against NAD(+), which was interpreted before as evidence of an iso mechanism, in which NAD(+) and NADH binds to different forms of free enzyme. With the aim of testing the proposed kinetic mechanism, we have now investigated the ability of NADH to form different complexes with the enzyme. By initial velocity and equilibrium binding studies, we found that the steady-state levels of E.glycine betaine are negligible, ruling out binding of NADH to this complex. However, NADH readily bind to E.betaine aldehyde, whose levels most likely are kinetically significant given its low dissociation constant. Also, NADH combined with E.NADH and E.NAD(+). Finally, NADH was not able to revert the hydride transfer step, what suggest that there is no acyl-enzyme intermediate, i.e. the release of the reduced dinucleotide takes place after the deacylation step. Although formation of the complex E.NAD(+).NADH would produce an uncompetitive effect in the inhibition of NADH against NAD(+), the iso mechanism cannot be conclusively discarded.  相似文献   

2.
Renal hyperosmotic conditions may produce reactive oxygen species, which could have a deleterious effect on the enzymes involved in osmoregulation. Hydrogen peroxide was used to provoke oxidative stress in the environment of betaine aldehyde dehydrogenase in vitro. Enzyme activity was reduced as hydrogen peroxide concentration was increased. Over 50% of the enzyme activity was lost at 100 μM hydrogen peroxide at two temperatures tested. At pH 8.0, under physiological ionic strength conditions, peroxide inhibited the enzyme. Initial velocity assays of betaine aldehyde dehydrogenase in the presence of hydrogen peroxide (0-200 μM) showed noncompetitive inhibition with respect to NAD(+) or to betaine aldehyde at saturating concentrations of the other substrate at pH 7.0 or 8.0. Inhibition data showed that apparent V(max) decreased 40% and 26% under betaine aldehyde and NAD(+) saturating concentrations at pH 8.0, while at pH 7.0 V(max) decreased 40% and 29% at betaine aldehyde and NAD(+) saturating concentrations. There was little change in apparent Km(NAD) at either pH, while Km(BA) increased at pH 7.0. K(i) values at pH 8 and 7 were calculated. Our results suggest that porcine kidney betaine aldehyde dehydrogenase could be inhibited by hydrogen peroxide in vivo, thus compromising the synthesis of glycine betaine.  相似文献   

3.
Intra- and intermolecular electron transfer processes in redox proteins   总被引:2,自引:0,他引:2  
Initial velocity and product inhibition experiments were performed to characterize the kinetic mechanism of branched chain ketoacid dehydrogenase (the branched chain complex) activity. The results were directly compared to predicted patterns for a three-site ping-pong mechanism. Product inhibition experiments confirmed that NADH is competitive versus NAD+ and isovaleryl CoA is competitive versus CoA. Furthermore, both NADH and isovaleryl CoA were uncompetitive versus ketoisovaleric acid. These results are consistent with a ping-pong mechanism and are similar to pyruvate dehydrogenase and alpha-ketoglutarate dehydrogenase. However, inhibition patterns for isovaleryl CoA versus NAD+ and NADH versus CoA are not consistent with a ping-pong mechanism. These patterns may result from a steric interaction between the flavoprotein and transacetylase subunits of the complex. To determine the kinetic mechanism of the substrates and feedback inhibitors (NADH and isovaleryl CoA) of the branched chain complex, it was necessary to define the interaction of the inhibitors at nonsaturating fixed substrate (CoA and NAD+) concentrations. While the competitive inhibition patterns were maintained, slope replots for NADH versus NAD+ at nonsaturating CoA concentrations were parabolic. This unexpected finding resembles a linear mixed type of inhibition where the inhibition is a combination of pure competitive and noncompetitive inhibition.  相似文献   

4.
L(+) lactate dehydrogenase (LDH) activity from the electric organ of Electrophorus electricus was measured in the presence of ATP in the forward (substrate lactate) and reverse (substrate pyruvate) enzymatic reactions. The I50 for ATP was first determined and then the kinetics of the reactions were investigated with either constant coenzyme (NAD or NADH) concentration and varying substrate (lactate or pyruvate) concentration, or, constant substrate and varying coenzyme concentration. The kinetic data showed that ATP inhibits LDH uncompetitively with respect to the reduced and the oxidized coenzyme. As for the substrates, ATP gives a mixed type inhibition for lactate and a noncompetitive inhibition for pyruvate.  相似文献   

5.
1. Mitochondrial aldehyde dehydrogenase is purified to near homogeneity by hydroxylapatite-, affinity- and hydrophobic interaction-chromatography. 2. The enzyme is an oligomeric protein and its molecular weight, as determined by gel-filtration, is 117,000 +/- 5000. 3. Active only in the presence of exogenous sulfhydryl compounds and NAD(+)-dependent, aldehyde dehydrogenase works optimally with linear-chain aliphatic aldehydes and is practically inactive with benzaldehyde. The pH-optimum is at about pH 8.5. 4. Km-Values for aliphatic aldehydes (C2-C6) range between 0.17 and 0.32 microM. The Km for NAD+ increases from 16 microM with acetaldehyde to 71 microM with capronaldehyde. 5. Millimolar concentrations of Mg2+ promote high increases of both V and Km for NAD+. At the same time, saturation curves with C4-C6 aldehydes can be simulated with a substrate inhibition model. 6. Inhibition by NADH is competitive: with capronaldehyde, the inhibition constant for NADH is 52 microM in the absence of Mg2+ and 14 microM in the presence of 4 mM Mg2+; with acetaldehyde, the inhibition constant is about three times higher (36 and 159 microM, respectively).  相似文献   

6.
N6022 is a novel, first-in-class drug with potent inhibitory activity against S-nitrosoglutathione reductase (GSNOR), an enzyme important in the metabolism of S-nitrosoglutathione (GSNO) and in the maintenance of nitric oxide (NO) homeostasis. Inhibition of GSNOR by N6022 and related compounds has shown safety and efficacy in animal models of asthma, chronic obstructive pulmonary disease, and inflammatory bowel disease [Sun, X., et al. (2011) ACS Med. Chem. Lett. 2, 402-406]. N6022 is currently in early phase clinical studies in humans. We show here that N6022 is a tight-binding, specific, and fully reversible inhibitor of GSNOR with an IC(50) of 8 nM and a K(i) of 2.5 nM. We accounted for the fact that the NAD(+)- and NADH-dependent oxidation and reduction reactions, catalyzed by GSNOR are bisubstrate in nature in our calculations. N6022 binds in the GSNO substrate binding pocket like a competitive inhibitor, although in kinetic assays it behaves with a mixed uncompetitive mode of inhibition (MOI) toward the GSNO substrate and a mixed competitive MOI toward the formaldehyde adduct, S-hydroxymethylglutathione (HMGSH). N6022 is uncompetitive with cofactors NAD(+) and NADH. The potency, specificity, and MOI of related GSNOR inhibitor compounds are also reported.  相似文献   

7.
The kinetic mechanisms of the 2-oxoglutarate and pyruvate dehydrogenease complexes from pig heart mitochondria were studied at pH 7.5 and 25 degrees. A three-site ping-pong mechanism for the actin of both complexes was proposed on the basis of the parallel lines obtained when 1/v was plotted against 2-oxoglutarate or pyruvate concentration for various levels of CoA and a level of NAD+ near its Michaelis constant value. Rate equations were derived from the proposed mechanism. Michaelis constants for the reactants of the 2-oxoglutarate dehydrogenase complex reaction are: 2-oxoglutarate, 0.220 mM; CoA, 0.025 mM; NAD+, 0.050 mM. Those of the pyruvate dehydrogenase complex are: pyruvate, 0.015 mM; CoA, 0.021 mM; NAD+, 0.079 mM. Product inhibition studies showed that succinyl-CoA or acetyl-CoA was competitive with respect to CoA, and NADH was competitive with respect to NAD+ in both overall reactions, and that succinyl-CoA or acetyl-CoA and NADH were uncompetitive with respect to 2-oxoglutarate or pyruvate, respectively. However, noncompetitive (rather than uncompetitive) inhibition patterns were observed for succinyl-CoA or acetyl-CoA versus NAD+ and for NADH versus CoA. These results are consistent with the proposed mechanisms.  相似文献   

8.
Abstract— The overall steady state kinetic mechanism of pyruvate dehydrogenase multienzyme complex purified from rat brain has been investigated. Initial rate patterns were a series of parallel lines regardless of which substrate was varied at several fixed concentrations of other substrates. Product inhibition patterns showed that acetyl CoA is competitive vs CoA, that NADH is competitive vs NAD, and that both acetyl CoA and NADH are uncompetitive vs pyruvate. Both acetyl CoA and NADH are noncompetitive vs NAD and CoASH, respectively. These results are inconsistent with classical 'hexa uni' ping-pong mechanisms, but are consistent with a non-classical 3-site ping-pong mechanism.  相似文献   

9.
Fan F  Germann MW  Gadda G 《Biochemistry》2006,45(6):1979-1986
Choline oxidase catalyzes the four-electron oxidation of choline to glycine betaine via two sequential FAD-dependent reactions in which betaine aldehyde is formed as an intermediate. The chemical mechanism for the oxidation of choline catalyzed by choline oxidase was recently elucidated by using kinetic isotope effects [Fan, F., and Gadda, G. (2005) J. Am. Chem. Soc. 127, 2067-2074]. In this study, the oxidation of betaine aldehyde has been investigated by using spectroscopic and kinetic analyses with betaine aldehyde and its isosteric analogue 3,3-dimethylbutyraldehyde. The pH dependence of the kcat/Km and kcat values with betaine aldehyde showed that a catalytic base with a pKa of approximately 6.7 is required for betaine aldehyde oxidation. Complete reduction of the enzyme-bound flavin was observed in a stopped-flow spectrophotometer upon anaerobic mixing with betaine aldehyde or choline at pH 8, with similar k(red) values > or = 48 s(-1). In contrast, only 10-26% of the enzyme-bound flavin was reduced by 3,3-dimethylbutyraldehyde between pH 6 and 10. Furthermore, this compound acted as a competitive inhibitor versus choline. NMR spectroscopic analyses indicated that betaine aldehyde exists predominantly (99%) as a diol form in aqueous solution. In contrast, the thermodynamic equilibrium for 3,3-dimethylbutyraldehyde favors the aldehyde (> or = 65%) over the hydrated form in the pH range from 6 to 10. The keto species of 3,3-dimethylbutyraldehyde is reactive toward enzymic nucleophiles, as suggested by the kinetic data with NAD+-dependent yeast aldehyde dehydrogenase. The data presented suggest that choline oxidase utilizes the hydrated species of the aldehyde as substrate in a mechanism for aldehyde oxidation in which hydride transfer is triggered by an active site base.  相似文献   

10.
Initial velocity and product inhibition studies were carried out on UDP-glucose dehydrogenase (UDPglucose: NAD+ 6-oxidoreductase, EC 1.1.1.22) from beef liver to determine if the kinetics of the reaction are compatible with the established mechanism. An intersecting initial velocity pattern was observed with NAD+ as the variable substrate and UDPG as the changing fixed substrate. UDPglucuronic acid gave competitive inhibition of UDPG and non-competitive inhibition of NAD+. Inhibition by NADH gave complex patterns.Lineweaver-Burk plots of 1/upsilon versus 1/NAD+ at varied levels of NADH gave highly non-linear curves. At levels of NAD+ below 0.05 mM, non-competitive inhibition patterns were observed giving parabolic curves. Extrapolation to saturation with NAD+ showed NADH gave linear uncompetitive inhibition of UDPG if NAD+ was saturating. However, at levels of NAD+ above 0.10 mM, NADH became a competitive inhibitor of NAD+ (parabolic curves) and when NAD+ was saturating NADH gave no inhibition of UDPG. NADH was non-competitive versus UDPG when NAD+ was not saturating. These results are compatible with a mechanism in which UDPG binds first, followed by NAD+, which is reduced and released. A second mol of NAD+ is then bound, reduced, and released. The irreversible step in the reaction must occur after the release of the second mol of NADH but before the release of UDPglucuronic acid. This is apparently caused by the hydrolysis of a thiol ester between UDPglucoronic acid and the essential thiol group of the enzyme. Examination of rate equations indicated that this hydrolysis is the rate-limiting step in the overall reaction. The discontinuity in the velocities observed at high NAD+ concentrations is apparently caused by the binding of NAD+ in the active site after the release of the second mol of NADH, eliminating the NADH inhibition when NAD+ becomes saturating.  相似文献   

11.
Inosine 5'-monophosphate dehydrogenase (IMPDH) catalyzes the oxidation of inosine 5'-monophosphate (IMP) to xanthosine 5'-monophosphate with the concomitant reduction of NAD to NADH. Escherichia coli IMPDH is activated by K(+), Rb(+), NH(+)(4), and Cs(+). K(+) activation is inhibited by Li(+), Na(+), Ca(2+), and Mg(2+). This inhibition is competitive versus K(+) at high K(+) concentrations, noncompetitive versus IMP, and competitive versus NAD. Thus monovalent cation activation is linked to the NAD site. K(+) increases the rate constant for the pre-steady-state burst of NADH production, possibly by increasing the affinity of NAD. Three mutant IMPDHs have been identified which increase the value of K(m) for K(+): Asp13Ala, Asp50Ala, and Glu469Ala. In contrast to wild type, both Asp13Ala and Glu469Ala are activated by all cations tested. Thus these mutations eliminate cation selectivity. Both Asp13 and Glu469 appear to interact with the K(+) binding site identified in Chinese hamster IMPDH. Like wild-type IMPDH, K(+) activation of Asp50Ala is inhibited by Li(+), Na(+), Ca(2+), and Mg(2+). However, this inhibition is noncompetitive with respect to K(+) and competitive with respect to both IMP and NAD. Asp50 interacts with residues that form a rigid wall in the IMP site; disruption of this wall would be expected to decrease IMP binding, and the defect could propagate to the proposed K(+) site. Alternatively, this mutation could uncover a second monovalent cation binding site.  相似文献   

12.
A gene having high sequence homology (45-49%) with the glycerol-1-phosphate dehydrogenase gene from Methanobacterium thermoautotrophicum was cloned from the aerobic hyperthermophilic archaeon Aeropyrum pernix K1 (JCM 9820). This gene expressed in Escherichia coli with the pET vector system consists of 1113 nucleotides with an ATG initiation codon and a TAG termination codon. The molecular mass of the purified enzyme was estimated to be 38 kDa by SDS/PAGE and 72.4 kDa by gel column chromatography, indicating presence as a dimer. The optimum reaction temperature of this enzyme was observed to be 94-96 degrees C at near neutral pH. This enzyme was subjected to two-substrate kinetic analysis. The enzyme showed substrate specificity for NAD(P)H-dependent dihydroxyacetone phosphate reduction and NAD(+)-dependent glycerol-1-phosphate (Gro1P) oxidation. NADP(+)-dependent Gro1P oxidation was not observed with this enzyme. For the production of Gro1P in A. pernix cells, NADPH is the preferred coenzyme rather than NADH. Gro1P acted as a noncompetitive inhibitor against dihydroxyacetone phosphate and NAD(P)H. However, NAD(P)(+) acted as a competitive inhibitor against NAD(P)H and as a noncompetitive inhibitor against dihydroxyacetone phosphate. This kinetic data indicates that the catalytic reaction by glycerol- 1-phosphate dehydrogenase from A. pernix follows a ordered bi-bi mechanism.  相似文献   

13.
Steady state initial velocity studies were carried out to determine the kinetic mechanism of human liver aldehyde dehydrogenase. Intersecting double reciprocal plots obtained in the absence of inhibitors demonstrated that the dehydrogenase reaction proceeded by sequential addition of both substrates prior to release of products. Dead end inhibition patterns obtained with coenzyme and substrate analogues (e.g. thionicotinamide-AD+ and chloral hydrate) indicated that NAD+ and aldehyde can bind in random fashion. The patterns of inhibition by the product NADH and of substrate inhibition by glyceraldehyde were also consistent with this mechanism. However, comparisons between kinetic constants associated with the dehydrogenase and esterase activities of this enzyme suggested that most of the dehydrogenase reaction flux proceeds via formation of an initial binary NAD+-enzyme complex over a wide range of substrate and coenzyme concentrations.  相似文献   

14.
The haem-distal pocket of nitric oxide reductase cytochrome P450 contains many Arg and Lys residues that are clustered to form a putative access channel for NADH. Asp88 is the sole negatively charged amino acid in this positive charge cluster, and thus it would be interesting to know its functional role. Here we found the intriguing phenomenon that mutation at this site of P450nor (D88A or D88V) considerably decreased the overall nitric oxide reductase activity without blocking the reducing half reaction in which the ferric enzyme-NO complex is reduced with NADH to yield a specific intermediate (I). The results indicate that the catalytic turnover subsequent to the I formation was blocked by such mutation. This property of the mutants made it possible to perform kinetic analysis of the reduction step, which is impossible with the wild-type P450nor. These results are the first kinetic evidence for direct complex formation between P450nor and an electron donor (NADH or NADPH). The kinetic analysis also showed that the inhibition by chloride ions (Cl(-)) is competitive with respect to NAD(P)H, which highlights the importance of the binding site for Cl(-) (the anion hole) in the interaction with NAD(P)H. We also characterized another mutant (D393A) of P450nor. The results demonstrated that both Asp residues play important roles in the interaction with NADH, whereas the role of Asp88 is unique in that it must be essential for the release of NAD(+) rather than binding to NADH.  相似文献   

15.
The effects of quercetin and resveratrol (substances found in red wine) on the activity of cytosolic aldehyde dehydrogenase in vitro are compared with those of the synthetic hormone diethylstilbestrol. It is proposed that quercetin inhibits the enzyme by binding competitively in both the aldehyde substrate binding-pocket and the NAD(+)-binding site, whereas resveratrol and diethylstilbestrol can only bind in the aldehyde site. When inhibition is overcome by high aldehyde and NAD(+) concentrations (1 mM of each), the modifiers enhance the activity of the enzyme; we hypothesise that this occurs through binding to the enzyme-NADH complex and consequent acceleration of the rate of dissociation of NADH. The proposed ability of quercetin to bind in both enzyme sites is supported by gel filtration experiments with and without NAD(+), by studies of the esterase activity of the enzyme, and by modelling the quercetin molecule into the known three-dimensional structure of the enzyme. The possibility that interaction between aldehyde dehydrogenase and quercetin may be of physiological significance is discussed.  相似文献   

16.
The kinetic mechanism of betaine aldehyde dehydrogenase from leaves of the plant Amaranthus hypochondriacus is ordered with NAD+ adding first. NADH is a noncompetitive inhibitor against NAD+, which was interpreted before as evidence of an iso mechanism, in which NAD+ and NADH binds to different forms of free enzyme. With the aim of testing the proposed kinetic mechanism, we have now investigated the ability of NADH to form different complexes with the enzyme. By initial velocity and equilibrium binding studies, we found that the steady-state levels of E.glycine betaine are negligible, ruling out binding of NADH to this complex. However, NADH readily bind to E.betaine aldehyde, whose levels most likely are kinetically significant given its low dissociation constant. Also, NADH combined with E.NADH and E.NAD+. Finally, NADH was not able to revert the hydride transfer step, what suggest that there is no acyl-enzyme intermediate, i.e. the release of the reduced dinucleotide takes place after the deacylation step. Although formation of the complex E.NAD+.NADH would produce an uncompetitive effect in the inhibition of NADH against NAD+, the iso mechanism cannot be conclusively discarded.  相似文献   

17.
Initial velocity studies and product inhibition studies were conducted for the forward and reverse reactions of formaldehyde dehydrogenase (formaldehyde: NAD oxidoreductase, EC 1.2.1.1) isolated from a methanol-utilizing yeast Candida boidinii. The data were consistent with an ordered Bi-Bi mechanism for this reaction in which NAD+ is bound first to the enzyme and NADH released last. Kinetic studies indicated that the nucleoside phosphates ATP, ADP and AMP are competitive inhibitors with respect to NAD and noncompetitive inhibitors with respect to S-hydroxymethylglutathione. The inhibitions of the enzyme activity by ATP and ADP are greater at pH 6.0 and 6.5 than at neutral or alkaline pH values. The kinetic studies of formate dehydrogenase (formate:NAD oxidoreductase, EC 1.2.1.2) from the methanol grown C. boidinii suggested also an ordered Bi-Bi mechanism with NAD being the first substrate and NADH the last product. Formate dehydrogenase the last enzyme of the dissimilatory pathway of the methanol metabolism is also inhibited by adenosine phosphates. Since the intracellular concentrations of NADH and ATP are in the range of the Ki values for formaldehyde dehydrogenase and formate dehydrogenase the activities of these main enzymes of the dissimilatory pathway of methanol metabolism in this yeast may be regulated by these compounds.  相似文献   

18.
The Michaelis constant values for the highly purified pyruvate dehydrogenase complex (PDC) from human heart are 25, 13 and 50 microM for pyruvate, CoA and NAD, respectively. Acetyl-CoA produces a competitive inhibition of PDC (Ki = 35 microM) with respect to CoA, whereas NADH produces the same type of inhibition with respect to NAD (Ki = 36 microM). The oxoglutarate dehydrogenase complex (OGDC) from human heart has active sites with two different affinities for 2-oxoglutarate ([S]0.5 of 30 and 120 microM). ADP (1 mM) decreases the [S]0.5 values by a half. The inhibition of OGDC (Ki = 81 microM) by succinyl-CoA is of a competitive type with respect to CoA (Km = 2.5 microM), whereas that of NADH (Ki = 25 microM) is of a mixed type with respect to NAD (Km = 170 microM).  相似文献   

19.
The kinetic properties of highly purified preparations of sheep liver cytoplasmic aldehyde dehydrogenase (preparations that had been shown to be free from contamination with the corresponding mitochondrial enzyme) were investigated with both propionaldehyde and butyraldehyde as substrates. At low aldehyde concentrations, double-reciprocal plots with aldehyde as the variable substrate are linear, and the mechanism appears to be ordered, with NAD+ as the first substrate to bind. Stopped-flow experiments following absorbance and fluorescence changes show bursts of NADH production in the pre-steady state, but the observed course of reaction depends on the pre-mixing conditions. Pre-mixing enzyme with NAD+ activates the enzyme in the pre-steady state and we suggest that the reaction mechanism may involve isomeric enzyme--NAD+ complexes. High concentrations of aldehyde in steady-state experiments produce significant activation (about 3-fold) at high concentrations of NAD+, but inhibition at low concentrations of NAD+. Such behaviour may be explained by postulating the participation of an abortive complex in product release. Stopped-flow measurements at high aldehyde concentrations indicate that the mechanism of reaction under these conditions is complex.  相似文献   

20.
The reduction of both NO2- and hydroxylamine by the NADH-dependent nitrite reductase of Escherichia coli K 12 (EC 1.6.6.4) appears to follow Michaelis-Menten kinetics over a wide range of NADH concentrations. Substrate inhibition can, however, be detected at low concentrations of the product NAD+. In addition, NAD+ displays mixed product inhibition with respect to NADH and mixed or uncompetitive inhibition with respect to hydroxylamine. These inhibition characteristics are consistent with a mechanism in which hydroxylamine binds during catalysis to a different enzyme form from that generated when NAD+ is released. The apparent maximum velocity with NADH as varied substrate increases as the NAD+ concentration increases from 0.05 to 0.7 mM with 1 mM-NO2- or 100 mM-hydroxylamine as oxidized substrate. This increase is more marked for hydroxylamine reduction than for NO2- reduction. Models incorporating only one binding site for NAD can account for the variation in the Michaelis-Menten parameters for both NADH and hydroxylamine with [NAD+] for hydroxylamine reduction. According to these models, activation of the reaction occurs by reversal of an over-reduction of the enzyme by NADH. If the observed activation of the enzyme by NAD+ derives both from activation of the generation of the enzyme-hydroxylamine complex from the enzyme-NO2- complex during NO2- reduction and from activation of the reduction of the enzyme-hydroxylamine complex to form NH4+, then the variation of Vapp. for NO2- or hydroxylamine with [NAD+] is consistent with the occurrence of the same enzyme-hydroxylamine complex as an intermediate in both reactions.  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