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 0 毫秒
1.
In a microemulsion made with Triton X-100, the stability of the enzymatic activity was higher than in ionic microemulsions. The stability increased with water content. The kinetic constants (Michaelis constant of NAD+ and maximum velocity) were close to those found in the previously studied microemulsions. The Michaelis constant of NAD+ expressed with respect to the buffer volume was higher than in water. The pH dependence of the kinetic constants in this microemulsion was determined. The activity determined by NAD+ reduction decreased with water content, whereas the redox activity determined via butanol oxidation coupled to retinal reduction was only slightly reduced.  相似文献   

2.
Horse liver alcohol dehydrogenase is a homodimer, the protomer having a coenzyme-binding domain and a catalytic domain. Using all available x-ray structures and 50 ns of molecular dynamics simulations, we investigated the mechanism of NAD+-induced domain closure. When the well-known loop at the domain interface was modeled to its conformation in the closed structure, the NAD+-induced domain closure from the open structure could be simulated with remarkable accuracy. Native interactions in the closed structure between Arg369, Arg47, His51, Ala317, Phe319, and NAD+ were seen to form at different stages during domain closure. Removal of the Arg369 side-chain charge resulted in the loss of the tendency to close, verifying that specific interactions do help drive the domains closed. Further simulations and a careful analysis of x-ray structures suggest that the loop prevents domain closure in the absence of NAD+, and a cooperative mechanism operates between the subunits for domain closure. This cooperative mechanism explains the role of the loop as a block to closure because in the absence of NAD+ it would prevent the occurrence of an unliganded closed subunit when the other subunit closes on NAD+. Simulations that started with one subunit open and one closed supported this.  相似文献   

3.
In the present study we show that the enzymatic activity of the coenzyme nicotinamide adenine dinucleotide (NAD+) and its analogues (C(O)NH2 replaced by C(S)NH2, C(O)CH3, C(O)H and CN) with horse liver alcohol dehydrogenase (LADH) (alcohol:NAD+ oxidoreductase, EC 1.1.1.1) can be rationalized by their conformation in the active site determined with molecular mechanics (AMBER, assisted model building with energy refinement). In order to establish the relation between the hydride transfer rate and the conformation of the NAD+ and its analogues, kinetic experiments with the poor substrate isopropanol were carried out. It appears that the enzymatic activity can be readily explained by the geometry of the pyridinium ring, in particular the magnitude of the 'out-of-plane' rotation of the carboxamide side chain (or analogues). The latter is nicely illustrated in the case of 3-cyanopyridine adenine dinucleotide which lacks any 'out-of-plane' rotation and concomitantly exhibits no significant enzymatic activity.  相似文献   

4.
Excitation transfer in complexes of horse liver alcohol dehydrogenase   总被引:3,自引:0,他引:3  
The protein fluorescence of LADH1 was quenched upon coupling with NADH, NAD+, o-phenanthroline, or thyroxine and its related compounds, while AMP, ADP, ADPR, or NMN did not quench the fluorescence. Addition of isobutyramide or pyrazole to E2R2 or E2O2 did not alter the degree of quenching. The coupling of two molecules of NADH to one molecule of E2I2 caused an equal fluorescence enhancement for both molecules of NADH when excited in its 340-mμ absorption band. However, with excitation in the protein absorption range, it was found that the binding of the first NADH molecule to LADH caused a larger fluorescence change than the binding of the second one. This was ascertained by following the increase of the fluorescence caused by addition of excess E2 to E2I2R2, whereby the complexes E2I2R and E2I2 were formed. This seemed to indicate that excitation energy could be transferred from one subunit to the other in the same LADH molecule.  相似文献   

5.
6.
Acetone was found to form a dead-end ternary complex with horse liver alcohol dehydrogenase and oxidized nicotinamide adenine dinucleotide (NAD+) when the reactants were incubated for a long time at relatively high concentrations. The complex formation was demonstrated by measuring the increase in absorbance at 320 nm, the quenching of protein fluorescence, and the loss of enzyme activity. Since acetone is a substrate of liver alcohol dehydrogenase, and the presence of acetaldehyde or pyrazole prevents acetone from forming the dead-end complex with liver alcohol dehydrogenase and NAD+, the acetone molecule in the complex may be bound to the substrate binding site of liver alcohol dehydrogenase. The dissociation of the complex was demonstrated by prolonged dialysis or by addition of reduced nicotinamide adenine dinucleotide (NADH) and iso-butyramide. A modified nicotinamide adenine dinucleotide was obtained as a main product from the dead-end complex after dissociation of the complex or denaturation of the apoenzyme. The modified nicotinamide adenine dinucleotide was found to exhibit an absorption spectrum similar to that of NADH; however, it was not oxidizable by liver alcohol dehydrogenase in the presence of acetaldehyde and exhibited no fluorescence.  相似文献   

7.
8.
9.
Structure and activity of methylated horse liver alcohol dehydrogenase   总被引:2,自引:0,他引:2  
  相似文献   

10.
A new, acyclic NAD-analog, acycloNAD+ has been synthesized where the nicotinamide ribosyl moiety has been replaced by the nicotinamide (2-hydroxyethoxy)methyl moiety. The chemical properties of this analog are comparable to those of β-NAD+ with a redox potential of −324 mV and a 341 nm λmax for the reduced form. Both yeast alcohol dehydrogenase (YADH) and horse liver alcohol dehydrogenase (HLADH) catalyze the reduction of acycloNAD+ by primary alcohols. With HLADH 1-butanol has the highest Vmax at 49% that of β-NAD+. The primary deuterium kinetic isotope effect is greater than 3 indicating a significant contribution to the rate limiting step from cleavage of the carbon–hydrogen bond. The stereochemistry of the hydride transfer in the oxidation of stereospecifically deuterium labeled n-butanol is identical to that for the reaction with β-NAD+. In contrast to the activity toward primary alcohols there is no detectable reduction of acycloNAD+ by secondary alcohols with HLADH although these alcohols serve as competitive inhibitors. The net effect is that acycloNAD+ has converted horse liver ADH from a broad spectrum alcohol dehydrogenase, capable of utilizing either primary or secondary alcohols, into an exclusively primary alcohol dehydrogenase. This is the first example of an NAD analog that alters the substrate specificity of a dehydrogenase and, like site-directed mutagenesis of proteins, establishes that modifications of the coenzyme distance from the active site can be used to alter enzyme function and substrate specificity. These and other results, including the activity with α-NADH, clearly demonstrate the promiscuity of the binding interactions between dehydrogenases and the riboside phosphate of the nicotinamide moiety, thus greatly expanding the possibilities for the design of analogs and inhibitors of specific dehydrogenases.  相似文献   

11.
12.
Examination of the model of the fixation site of the adenosine phosphate part of NAD+ on horse liver alcohol dehydrogenase led us to synthesize a NAD+ analogue N6-[N-(8-amino-3,6-dioxaoctyl)carbamoylmethyl]-NAD+ in order to alkylate the carboxylic acid group of Asp-273 and to convert the normally dissociable coenzyme into a permanently bound prosthetic group. This NAD+ analogue is coupled to the horse liver alcohol dehydrogenase in the ternary complex formed with pyrazole. In these conditions the degree of fixation varies between 0.4 and 0.58 coenzyme molecule/enzyme subunit molecule. The N6-[N-(8-amino-3,6-dioxaoctyl)carbamoylmethyl]NAD+ acts as a true prosthetic group which can be reduced and reoxidized by a coupled substrate reaction and the internal activity of this holoenzyme corresponds to the amount of analogue incorporated.  相似文献   

13.
Alcohol dehydrogenase from horse liver was immobilized by covalent attachment to CNBr-Sepharose and by adsorption to octyl-Sepharose CL-4B, a hydrophobic analog of Sepharose. In each case, rate constants for the binding and release of coenzyme and for the oxidation of substrates were measured based on the concentration of accessible active-site zinc atoms determined by titration with a paramagnetic inhibitor. All rate constants were substantially reduced upon immobilization; however, the rate constant of immobilized enzyme for ethanol oxidation was independent of the immobilization method, whereas the rate constant for cyclohexanol oxidation was lower for enzyme immobilized to octyl-Sepharose. Consequently, the substrate specificity of the two immobilized enzyme samples differed by an order of magnitude. Moreover, EPR spectroscopy studies and computer graphic analyses of spin labels occupying three defined regions of the active-site domain indicated that the active-site conformation adjacent to the catalytic zinc atom was similar in the two samples while the conformation slightly further from the zinc atom was different. This result may explain why the two immobilized enzyme preparations exhibited the same rate constant toward a small substrate (ethanol) yet different rate constants toward a larger substrate (cyclohexanol), whose rate constant is expected to be sensitive to a larger portion of the active site.  相似文献   

14.
15.
16.
Activity and flexibility of alcohol dehydrogenase in organic solvents   总被引:2,自引:0,他引:2  
The oxidation of cinnamyl alcohol to cinnamaldehyde by horse liver alcohol dehydrogenase (LADH) was carried out in nearly anhydrous organic solvents and in solvents containing from 0.1 to 10% added water. In nearly anhydrous solvents containing less than 0.02% water, the oxidation rate increased as the water solubility in the solvent decreased, but the reaction did not require active LADH. Moreover, the highest activity in nearly anhydrous heptane was obtained by lyophilizing the enzyme from a solution of pH 2.0, even though LADH exhibits virtually no enzymatic activity in water at this pH. The catalytic activity of LADH was restored and increased dramatically as small amounts of water were added to each solvent. In conjunction with the activity measurements, electron paramagnetic resonance (EPR) spectroscopy and two active-site directed spin labels were used to examine solvent-dependent structural features of LADH. The EPR spectra indicated that LADH became more rigid as the dielectric constant of the solvent decreased. The degree of rigidity also depended on the pH from which the enzyme was lyophilized, indicating that the ionization state of the enzyme can have an important influence on its dynamics in organic solvents. Finally, adding 1% water to organic solvents had no apparent effect on the enzyme's conformation or flexibility near the spin label, even though enzyme activity was an order of magnitude higher when 1% water was present.  相似文献   

17.
18.
The geometry of seven NAD+ analogues bound to horse liver alcohol dehydrogenase (LADH) modified only in their nicotinamide group, have been studied using AMBER molecular mechanics energy-minimization procedures. Starting geometries were taken from X-ray crystallographic data for NAD+/Me2SO/LADH reported by Eklund and co-workers. In this study the NAD+ analogues were encaged by the constituent amino acids of the enzyme within a range of 0.6 nm from the initial NAD+/Me2SO/Zn2+ complex. The calculational method used is able to rationalize individual substituent effects and to evaluate the essential interactions between NAD+ analogue, enzyme, Me2SO and Zn2+ without the necessity of additional X-ray data. The results presented here demonstrate that the reactivity of NAD+ derivatives as reported in literature can be qualitatively related to the position of the pyridine moiety in the active site.  相似文献   

19.
Changes in reduced viscosity of nuclear lysates from rat liver cells have been studied, in conditions of very low shear stress by the use of an oscillating viscometer, as a function of incubation time in alkaline (pH 12.5) and neutral (pH 8.0) solutions. In non-denaturing conditions, nuclear DNA showed a stepwise, time-dependent increase of reduced viscosity, which suggests that it behaves as a single hydrodynamic unit that progressively changes its radius and viscoelastic properties because of a very slow unfolding, through discrete successive transitions, from a highly superpacked structure toward a linear relaxed B-form fiber. Experimental conditions shown to reduce chromatin-DNA superpacking without changing DNA length (e.g. G1 cycling versus G0 non-cycling liver cells, or young versus old rat liver cells) dramatically increased the initial value of reduced viscosity and its time-dependent increment. Conversely, in denaturing conditions, reduced viscosity increased in the initial phase (probably because DNA unfolding prevails on DNA unwinding), then exhibited a plateau level (when unfolding balances unwinding), and subsequently decreased progressively to the value of sheared DNA (when unwinding becomes more rapid due to the progressive breakage of phosphodiester bridges in alkali). Experimental conditions known to induce DNA single- or double-strand breaks (i.e. the use of liver cells from rats treated with dimethylnitrosamine or 2-acetylaminofluorene, or of liver cells exposed to X-rays) caused in both neutral and alkaline solution an increment in the initial reduced viscosity and in the slope of its time-dependent increase, which may be related to a reduction of chromatin-DNA superpacking. Moreover, it became evident in denaturing conditions that a decrease of the maximum viscosity and of the time taken to reach it both related to a reduced DNA length. These viscoelastic properties are constantly correlated with independent DNA structural measurements on the same nuclear lysates, to discriminate the effect due to mere aggregation and disaggregation.  相似文献   

20.
The crystals of holoenzyme from native and cross-linked alcohol dehydrogenase exhibit electron transfer from NADH to phenazinium methosulfate (PMS), and then to the tetrazolium salt sodium 3,3'-{1-[(phenylamino)carbonyl]-3,4-tetrazolium}-bis(4-methoxy-6-nitro)benzenesulfonate (XXT). The slow dissociation of the cofactor and/or the conformational change associated can now be bypassed. The reduction product, formazan, did not diffuse out of the crystals in buffer and the crystals turned colored. In the presence of dimethyl sulfoxide or dimethoxyethane, the formazan diffused out to the solution. The reaction rates were found to be, respectively, 18% and 15% of the redox reaction rate of ethanol with cinnamaldehyde, close to the activity determined for the enzyme in solution in the presence of dimethoxyethane. The use of system PMS-tetrazolium salt is a useful tool to visualize the activity of dehydrogenases and other electron transferring systems in the crystalline state. The adsorption of formazan by the alcohol dehydrogenase crystals occurs in solution.  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