首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 15 毫秒
1.
The influence of coenzyme structure on the transient chemical intermediate formed in the reaction between the horse-liver alcohol dehydrogenase-NADH complex and an aromatic aldehyde such as 4-trans-(N,N-dimethylamino)cinnamaldehyde or 4-(N,N-dimethylamino)benzaldehyde was investigated by substituting various adenylic dinucleotides for NADH. Two classes of dinucleotide were studied. (a) Dinucleotides which, in the presence of horse-liver alcohol dehydrogenase and either 4-(N,N-dimethylamino)benzaldehyde or 4-trans-(N,N-dimethylamino)cinnamaldehyde, lead to a chromophore structurally analogous to the transient chemical intermediate formed with NADH under the same experimental conditions. This includes dinucleotides with a neutral 1,4-dihydropyridine ring, analogues of NADH and adducts of NAD+ (or analogues) with enolizable carbonyl compounds. (b) Dinucleotides which, under the same experimental conditions, do not form any new chromophores when mixed with horse-liver alcohol dehydrogenase and either 4-trans-(N,N-dimethylamino)cinnamaldehyde or 4-trans-(N,N-dimethylamino)benzaldehyde. This includes oxidized coenzyme analogues, NADPH and NADP+ adducts. Our data suggest that a neutral 1,4-dihydropyridine ring is crucial for the formation of the transient chemical intermediate. When the NAD+-sulphite complex, which has a 1,4-dihydronicotinamide structure and a positive charge at position 4 neutralized by sulphite ions, was substituted for NADH, the transient chemical intermediate chromophore was observed. The implications of this phenomenon are examined by assuming the existence of intermediate-activated forms of substrates and coenzymes during the horse-liver alcohol dehydrogenase catalytic reduction of aldehydes.  相似文献   

2.
Abstract: Alcohol dehydrogenase (ADH) is a key enzyme for the production of butanol, ethanol, and isopropanol by the solvent-producing clostridia. Initial studies of ADH in extracts of several strains of Clostridium acetobutylicum and C. beijerinckii gave conflicting molecular properties. A more coherent picture has emerged because of the following results: (i) identification of ADHs with different coenzyme specificities in these species; (ii) discovery of structurally conserved ADHs (type 3) in three solvent-producing species; (iii) isolation of mutants with deficiencies in butanol production and restoration of butanol production with a cloned alcohol/aldehyde dehydrogenase gene; and (iv) resolution of various ' C. acetobutylicum ' cultures into four species. The three ADH isozymes of C. beijerinckii NRRL B592 have high sequence similarities to ADH-1 of Clostridium sp. NCP 262 (formerly C. acetobutylicum P262) and to the ADH domain of the alcohol/aldehyde dehydrogenase of C. acetobutylicum ATCC 824/DSM 792. The NADH-dependent activity of the ADHs from C. beijerinckii NRRL B592 and the BDHs from C. acetobutylicum ATCC 824 is profoundly affected by the pH of the assay, and the relative importance of NADH and NADPH to butanol production may be misappraised when NAD(P)H-dependent activities were measured at different pH values. The primary/secondary ADH of isopropanol-producing C. beijerinckii is a type-1 enzyme and is highly conserved in Thermoanaerobacter brockii (formerly Thermoanaerobium brockii ) and Entamoeba histolytica . Several solvent-forming enzymes (primary ADH, aldehyde dehydrogenase, and 3-hydroxybutyryl-CoA dehydrogenase) are very similar between C. beijerinckii and the species represented by Clostridium sp. NCP 262 and NRRL B643. The realization of such relationships will facilitate the elucidation of the roles of different ADHs because each type of ADH can now be studied in an organism most amenable to experimental manipulations.  相似文献   

3.
The proton magnetic resonance spectra of the dihydronicotinamide ring of αNADH3 and the nicotinamide ring of αNAD+ are reported and the proton absorptions assigned. The absolute assignment of the C4 methylene protons of αNADH is based on the generation of specifically deuterium-labeled (pro-S) B-deuterio-αNADH from enzymatically prepared B-deuterio-βNADH. The C4 proton absorption of αNAD+ is assigned by oxidation of B-deuterio-αNADH by the A specific, yeast alcohol dehydrogenase to yield 4-deuterio-αNAD+.The epimerization of either αNADH or βNADH yields an equilibrium ratio of approximately 9:1 βNADH to αNADH. The rate of epimerization of αNADH to βNADH at 38 °C in 0.05, pH 7.5, phosphate buffer is 3.1 × 10?3 min?1, corresponding to a half-life of 4 hr. Four related dehydrogenases, yeast and horse liver alcohol dehydrogenase and chicken M4 and H4 lactate dehydrogenase, are shown to oxidize αNADH to αNAD+ at rates three to four orders of magnitude slower than for βNADH. By using specifically labeled B-deuterio-αNADH the enzymatic oxidation by yeast alcohol dehydrogenase has been shown to occur with the identical stereospecificity as the oxidation of βNADH. The nonenzymatic epimerization of αNADH to βNADH and the enzymatic oxidation αNADH are discussed as a possible source of αNAD+in vivo.  相似文献   

4.
The structure of mouse class II alcohol dehydrogenase (ADH2) has been determined in a binary complex with the coenzyme NADH and in a ternary complex with both NADH and the inhibitor N-cyclohexylformamide to 2.2 A and 2.1 A resolution, respectively. The ADH2 dimer is asymmetric in the crystal with different orientations of the catalytic domains relative to the coenzyme-binding domains in the two subunits, resulting in a slightly different closure of the active-site cleft. Both conformations are about half way between the open apo structure and the closed holo structure of horse ADH1, thus resembling that of ADH3. The semi-open conformation and structural differences around the active-site cleft contribute to a substantially different substrate-binding pocket architecture as compared to other classes of alcohol dehydrogenase, and provide the structural basis for recognition and selectivity of alcohols and quinones. The active-site cleft is more voluminous than that of ADH1 but not as open and funnel-shaped as that of ADH3. The loop with residues 296-301 from the coenzyme-binding domain is short, thus opening up the pocket towards the coenzyme. On the opposite side, the loop with residues 114-121 stretches out over the inter-domain cleft. A cavity is formed below this loop and adds an appendix to the substrate-binding pocket. Asp301 is positioned at the entrance of the pocket and may control the binding of omega-hydroxy fatty acids, which act as inhibitors rather than substrates. Mouse ADH2 is known as an inefficient ADH with a slow hydrogen-transfer step. By replacing Pro47 with His, the alcohol dehydrogenase activity is restored. Here, the structure of this P47H mutant was determined in complex with NADH to 2.5 A resolution. His47 is suitably positioned to act as a catalytic base in the deprotonation of the substrate. Moreover, in the more closed subunit, the coenzyme is allowed a position closer to the catalytic zinc. This is consistent with hydrogen transfer from an alcoholate intermediate where the Pro/His replacement focuses on the function of the enzyme.  相似文献   

5.
Dihydrolipoamide dehydrogenase is a flavoenzyme that reversibly catalyzes the oxidation of reduced lipoyl substrates with the reduction of NAD+ to NADH. In vivo, the dihydrolipoamide dehydrogenase component (E3) is associated with the pyruvate, α-ketoglutarate, and glycine dehydrogenase complexes. The pyruvate dehydrogenase (PDH) complex connects the glycolytic flux to the tricarboxylic acid cycle and is central to the regulation of primary metabolism. Regulation of PDH via regulation of the E3 component by the NAD+/NADH ratio represents one of the important physiological control mechanisms of PDH activity. Furthermore, previous experiments with the isolated E3 component have demonstrated the importance of pH in dictating NAD+/NADH ratio effects on enzymatic activity. Here, we show that a three-state mechanism that represents the major redox states of the enzyme and includes a detailed representation of the active-site chemistry constrained by both equilibrium and thermodynamic loop constraints can be used to model regulatory NAD+/NADH ratio and pH effects demonstrated in progress-curve and initial-velocity data sets from rat, human, Escherichia coli, and spinach enzymes. Global fitting of the model provides stable predictions to the steady-state distributions of enzyme redox states as a function of lipoamide/dihydrolipoamide, NAD+/NADH, and pH. These distributions were calculated using physiological NAD+/NADH ratios representative of the diverse organismal sources of E3 analyzed in this study. This mechanistically detailed, thermodynamically constrained, pH-dependent model of E3 provides a stable platform on which to accurately model multicomponent enzyme complexes that implement E3 from a variety of organisms.  相似文献   

6.
Glyceraldehyde-3-phosphate dehydrogenase binds to homologous and heterologous single-stranded but not double-stranded DNA. Binding to RNA, poly(A) and poly(dA-dT) has also been observed. Enzyme binding to these nucleic acids leads to the formation of an insoluble complex which can be sedimented at low speed.The interaction of glyceraldehyde-3-phosphate dehydrogenase with DNA is strongly inhibited by NAD and NADH but not by NADP. Adenine nucleotides, which inhibit the dehydrogenase activity by competing with NAD for its binding site (Yang, S.T. and Deal, W.C., Jr. (1969) Biochemistry 8, 2806–2813), also inhibit enzyme binding to DNA, whereas glyceraldehyde-3-phosphate and inorganic phosphate are non-inhibitory. These results suggest that DNA interacts through the NAD binding sites of glyceraldehyde-3-phosphate dehydrogenase. In accordance with this idea, it was found that DNA also binds to lactate dehydrogenase, an enzyme containing a similar dinucleotide binding domain, and that this binding is inhibited by NADH.A study of the base specificity of the DNA-glyceraldehyde-3-phosphate dehydrogenase interaction using dinucleoside monophosphates shows that inhibition of DNA binding by the dinucleotides requires the presence of a 3′-terminal adenosine and is greater when the 5′-terminus contains a pyrimidine instead of a purine. These results suggest that the dinucleotides bind at the NAD site of the dehydrogenase and that the enzyme would interact preferentially with PypA dinucleotides present in the nucleic acid.  相似文献   

7.
Cell-free extracts of a streptomycin-bleached strain of Euglena gracilis var. bacillaris have been examined for enzyme systems primarily responsible for the oxidation of reduced pyridine nucelotides. NADH lipoyl dehydrogenase, NADH and NADPH oxidase, NADH and NADPH diaphorase, and NADH and NADPH cytochrome c reductase have been demonstrated. The NADPH-linked enzymes had lower activity rates and were less sensitive to N-ethyl maleimide and p-hydroxymercuribenzoate than their NADH-linked counterparts. NADH cytochrome c reductase was the most sensitive to antimycin A. Michaelis-Menten constants (Km) determined were as follows: NADH diaphorase, 350 muM; NADPH oxidase 150 muM ; NADH lipoyl dehydrogenase, 0.35 muM. Enzyme activities after storage at -5 C indicate that the diaphorases are less labile than the other tested enzymes, and the differential activities of the NADH and NADPH linked enzymes suggest that functionally they may have different roles.  相似文献   

8.
Nigrostriatal cell death in 1-methyl-4-phenyl-1,2,3,6-tetrahydropyridine (MPTP)-induced Parkinson's disease results from the inhibition of mitochondrial respiration by 1-methyl-4-phenylpyridinium (MPP+). MPP+ blocks electron flow from NADH dehydrogenase to coenzyme Q at or near the same site as do rotenone and piericidin and protects against binding of and loss of activity due to these inhibitors. The 4'-analogs of MPP+ showed increasing affinity for the site with increasing length of alkyl chain, with the lowest Ki, for 4'-heptyl-MPP+, being 6 microM. The 4'-analogs compete with rotenone for the binding site in a concentration-dependent manner. They protect the activity of the enzyme from inhibition by piericidin in parallel to preventing its binding, indicating that the analogs and piericidin bind at the same inhibitory site(s). The optimum protection, however, was afforded by 4'-propyl-MPP+. The lesser protection by the more lipophilic MPP+ analogs with longer alkyl chains may involve a different orientation in the hydrophobic cleft, allowing rotenone and piericidin to still bind even when the pyridinium cation is in a position to interrupt electron flow from NADH to coenzyme Q.  相似文献   

9.
Anaerobically grown cells of Saccharomyces cerevisiae entrapped in polyacrylamide gel have been shown to provide a stable source of alcohol dehydrogenase [(ADH) alcohol:NAD+ oxidoreductase, EC 1.1.1.1] for effective regeneration of NAD(H). This system was able to provide the coenzyme required for the operation of other dehydrogenases, such as lactate dehydrogenase [(LDH) l-lactate: NAD+ oxidoreductase, EC 1.1.1.27] and malate dehydrogenase [(MDH) l-malate:NAD+ oxidoreductase, EC 1.1.1.37]. Yeast cells coimmobilized with a dehydrogenase are capable of the reversible regeneration of the reduced or oxidized coenzyme, depending on the additions made. A two-cell system can also be constituted using the same strain of yeast, adapted differently. Cells grown anaerobically and aerobically as sources of ADH and MDH, respectively, can operate efficiently on coimmobilization. The system can be used repeatedly without measurable loss of efficiency.  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.
Alcohol dehydrogenase SS, free from other isoenzymes, has been purified from horse livers. The enzyme has high activity with NADP(H) as coenzyme. With NADPH its activity is 3 times more than with NADH. While its affinity for NADPH is less than for NADH, in comparison with the classical ADH its affinity for NADP(H) is increased. In its activity with NADP(H) and inhibition with barbiturates, ADH SS resembles aldehyde reductases.  相似文献   

12.
The proposal that the lipoate acetyltransferase component (E2) of the pyruvate dehydrogenase multienzyme (PD) complex from Escherichia coli contains three covalently bound lipoyl residues, one of which acts to pass reducing equivalents to lipoamide dehydrogenase (E3), has been tested. The PD complex was incubated with pyruvate and N-ethylmaleimide, to yield an inactive PD complex containing lipoyl groups on E2 with the S6 acetylated and the S8H irreversibly alkylated with N-ethylmaleimide. This chemically modified form would be expected to exist only on two of the three proposed lipoyl groups. The third nonacetylatable lipoyl group, which is proposed to interact with E3, would remain in its oxidized form. Reaction of the N-ethylmaleimide-modified PD complex with excess NADH should generate the reduced form of the proposed third nonacetylatable lipoyl group and thereby make it susceptible to cyclic dithioarsinite formation with bifunctional arsenicals (BrCH2CONHPhAsCl2; BrCH2[14C]CONHPhAsO). Once "anchored" to the reduced third lipoyl group via the--AsO moiety, these reagents would be delivered into the active site of E3 by the normal catalytic process of the PD complex where the BrCH2CONH--group inactivates E3. Whereas the E3 component of native PD complex is inactivated by the bifunctional reagents in the presence of excess NADH (owing to the above delivery process), the E3 component of the PD complex modified with N-ethylmaleimide in the presence of pyruvate is not inhibited. The results indicate that acetylatable lipoyl residues interact directly with E3 and do not support a functional role for a proposed third lipoyl residue.  相似文献   

13.
1. The reaction of the pyruvate dehydrogenase multienzyme complex of Escherichia coli with maleimides was examined. In the absence of substrates, the complex showed little or no reaction with N-ethylmaleimide. However, in the presence of pyruvate and N-ethylmaleimide, inhibition of the pyruvate dehydrogenase complex was rapid. Modification of the enzyme was restricted to the transacetylase component and the inactivation was proportional to the extent of modification. The lipoamide dehydrogenase activity of the complex was unaffected by the treatment. The simplest explanation is that the lipoyl groups on the transacetylase are reductively acetylated by following the initial stages of the normal catalytic cycle, but are thereby made susceptible to modification. Attempts to characterize the reaction product strongly support this conclusion. 2. Similarly, in the presence of N-ethylmaleimide and NADH, much of the pyruvate dehydrogenase activity was lost within seconds, whereas the lipoamide dehydrogenase activity of the complex disappeared more slowly: the initial site of the reaction with the complex was found to be in the lipoyl transacetylase component. The simplest interpretation of these experiments is that NADH reduces the covalently bound lipoyl groups on the transacetylase by means of the associated lipoamide dehydrogenase component, thereby rendering them susceptible to modification. However, the dependence of the rate and extent of inactivation on NADH concentration was complex and it proved impossible to inhibit the pyruvate dehydrogenase activity completely without unacceptable modification of the other component enzymes. 3. The catalytic reduction of 5,5'-dithiobis-(2-nitrobenzoic acid) by NADH in the presence of the pyruvate dehydrogenase complex was demonstrated. A new mechanism for this reaction is proposed in which NADH causes reduction of the enzyme-bound lipoic acid by means of the associated lipoamide dehydrogenase component and the dihydrolipoamide is then oxidized back to the disulphide form by reaction with 5,5'-dithiobis-(2-nitrobenzoic acid). 4. A maleimide with a relatively bulky N-substituent, N-(4-diemthylamino-3,5-dinitrophenyl)maleimide, was an effective replacement for N-ethylmaleimide in these reactions with the pyruvate dehydrogenase complex. 5. The 2-oxoglutarate dehydrogenase complex of E. coli behaved very similarly to the pyruvate dehydrogenase complex, in accord with the generally accepted mechanisms of the two enzymes. 6. The treatment of the 2-oxo acid dehydrogenase complexes with maleimides in the presence of the appropriate 2-oxo acid substrate provides a simple method for selectively inhibiting the transacylase components and for introducing reporter groups on to the lipoyl groups covalently bound to those components.  相似文献   

14.
Fatty acid biosynthesis from Na[1-14C]acetate was characterized in plastids isolated from primary roots of 7-day-old germinating pea (Pisum sativum L.) seeds. Fatty acid synthesis was maximum at 82 nanomoles per hour per milligram protein in the presence of 200 micromolar acetate, 0.5 millimolar each of NADH, NADPH, and coenzyme A, 6 millimolar each of ATP and MgCl2, 1 millimolar each of MnCl2 and glycerol-3-phosphate, 15 millimolar KHCO3, 0.31 molar sucrose, and 0.1 molar Bis-Tris-propane, pH 8.0, incubated at 35°C. At the standard incubation temperature of 25°C, fatty acid synthesis was essentially linear for up to 6 hours with 80 to 120 micrograms per milliliter plastid protein. ATP and coenzyme A were absolute requirements, whereas divalent cations, potassium bicarbonate, and reduced nucleotides all variously improved activity two- to 10-fold. Mg2+ and NADH were the preferred cation and nucleotide, respectively. Glycerol-3-phosphate had little effect, whereas dithiothreitol and detergents generally inhibited the incorporation of [14C]acetate into fatty acids. On the average, the principal radioactive products of fatty acid biosynthesis were approximately 39% palmitic, 9% stearic, and 52% oleic acid. The proportions of these fatty acids synthesized depended on the experimental conditions.  相似文献   

15.
Xu H  West AH  Cook PF 《Biochemistry》2007,46(25):7625-7636
A survey of NADH, alpha-Kg, and lysine analogues has been undertaken in an attempt to define the substrate specificity of saccharopine dehydrogenase and to identify functional groups on all substrates and dinucleotides important for substrate binding. A number of NAD analogues, including NADP, 3-acetylpyridine adenine dinucleotide (3-APAD), 3-pyridinealdehyde adenine dinucleotide (3-PAAD), and thionicotinamide adenine dinucleotide (thio-NAD), can serve as a substrate in the oxidative deamination reaction, as can a number of alpha-keto analogues, including glyoxylate, pyruvate, alpha-ketobutyrate, alpha-ketovalerate, alpha-ketomalonate, and alpha-ketoadipate. Inhibition studies using nucleotide analogues suggest that the majority of the binding energy of the dinucleotides comes from the AMP portion and that distinctly different conformations are generated upon binding of the oxidized and reduced dinucleotides. Addition of the 2'-phosphate as in NADPH causes poor binding of subsequent substrates but has little effect on coenzyme binding and catalysis. In addition, the 10-fold decrease in affinity of 3-APAD in comparison to NAD suggests that the nicotinamide ring binding pocket is hydrophilic. Extensive inhibition studies using aliphatic and aromatic keto acid analogues have been carried out to gain insight into the keto acid binding pocket. Data suggest that a side chain with three carbons (from the alpha-keto group up to and including the side chain carboxylate) is optimal. In addition, the distance between the C1-C2 unit and the C5 carboxylate of the alpha-keto acid is also important for binding; the alpha-oxo group contributes a factor of 10 to affinity. The keto acid binding pocket is relatively large and flexible and can accommodate the bulky aromatic ring of a pyridine dicarboxylic acid and a negative charge at the C3 but not the C4 position. However, the amino acid binding site is hydrophobic, and the optimal length of the hydrophobic portion of the amino acid carbon side chain is three or four carbons. In addition, the amino acid binding pocket can accommodate a branch at the gamma-carbon, but not at the beta-carbon.  相似文献   

16.
The activity of alpha-ketoglutarate dehydrogenase complex from pigeon breast muscle is controlled by ADP and the reaction products, i. e. succinyl-CoA and NADH. ADP activates the alpha-ketoglutarate dehydrogenase component of the complex, whereas NADH inhibits alpha-ketoglutarate dehydrogenase and lipoyl dehydrogenase. In the presence of NADH the kinetic curve of the complex with respect to alpha-ketoglutarate and NAD and the dependence of upsilon versus [NAD] and upsilon versus [Lip (SH)2] in the lipoyl dehydrogenase reaction are S-shaped. In the absence of inhibitor ADP had no activating effect on lipoyl dehydrogenase; however, in the presence of NADH ADP decreases the cooperativity for NAD. The cooperative kinetics of the constituent enzymes of the complex are indicative of its allosteric properties. Isolation of the alpha-ketoglutarate dehydrogenase complex and its lipoyl dehydrogenase and alpha-ketoglutarate dehydrogenase components in a desensitized state confirms their allosteric nature. It is assumed that NADH effects of isolated alpha-ketoglutarate dehydrogenase is due to a shift in the equilibrium between different oligomeric forms of the enzyme.  相似文献   

17.
The stereochemistry of the hydrogen transfer to NAD catalyzed by ribitol dehydrogenase (ribitol:NAD 2-oxidoreductase, EC 1.1.1.56) from Klebsiella pneumoniae and D-mannitol-1-phosphate dehydrogenase (D-mannitol-1-phosphate:NAD 2-oxidoreductase, EC 1.1.1.17) from Escherichia coli was investigated. [4-3H]NAD was enzymatically reduced with nonlabelled ribitol in the presence of ribitol dehydrogenase and with nonlabelled D-mannitol 1-phosphate and D-mannitol 1-phosphate dehydrogenase, respectively. In both cases the [4-3H]-NADH produced was isolated and the chirality at the C-4 position determined. It was found that after the transfer of hydride, the label was in both reactions exclusively confined to the (4R) position of the newly formed [4-3H]NADH. In order to explain these results, the hydrogen transferred from the nonlabelled substrates to [4-3H]NAD must have entered the (4S) position of the nicotinamide ring. These data indicate for both investigated inducible dehydrogenases a classification as B or (S) type enzymes. Ribitol also can be dehydrogenated by the constitutive A-type L-iditol dehydrogenase (L-iditol:NAD 5-oxidoreductase, EC 1.1.1.14) from sheep liver. When L-iditol dehydrogenase utilizes ribitol as hydrogen donor, the same A-type classification for this oxidoreductase, as expected, holds true. For the first time, opposite chirality of hydrogen transfer to NAD in one organic reaction--ribitol + NAD = D-ribu + NADH + H--is observed when two different dehydrogenases, the inducible ribitol dehydrogenase from K. pneumoniae and the constitutive L-iditol dehydrogenase from sheep liver, are used as enzymes. This result contradicts the previous generalization that the chirality of hydrogen transfer to the coenzyme for the same reaction is independent of the source of the catalyzing enzyme.  相似文献   

18.
The vinylogue of NAD, 3-pyridylacryloamide adenine dinucleotide, was prepared from NAD and 3-pyridylacryloamide through the snake venom NADase-catalyzed transglycosidation reaction. The analog, purified by ion-exchange chromatography, was obtained in a 55% yield. The cyanide adduct and reduced form of the analog exhibited absorbance maxima at 358 nm and 378 nm, respectively, with extinction coefficients in each case being 2.3-times higher than those reported for the corresponding NAD derivatives. 3-Pyridylacryloamide adenine dinucleotide served as a coenzyme with bovine liver glutamic dehydrogenase and to a lesser extent with malate and lactate dehydrogenases. The analog was not reduced in reactions catalyzed by yeast and horse liver alcohol dehydrogenases, sheep liver sorbitol dehydrogenase, and rabbit muscle glycerophosphate dehydrogenase. Substitution of the pyridylacryloamide analogs for NAD and NADH in the assay of substrates for glutamic dehydrogenase was demonstrated.  相似文献   

19.
Purified NAD-linked d-lactate dehydrogenase from the depressor muscle of the giant barnacle, Balanus nubilus Darwin, is inactivated when incubated with the metal chelators o-phenanthroline and EDTA. M-Phenanthroline and p-phenanthroline, which lack metal chelating ability, are ineffective in inactivating the enzyme. Inactivated enzyme can be reactivated by the addition of zinc ions to the assay mixture. Atomic absorption spectrophotometric analysis of purified B. nubilusd-lactate dehydrogenase revealed that this enzyme contains stoichiometric amounts of zinc (2 g-atoms per mol of subunit), unlike other lactate dehydrogenases, which lack zinc. Zinc appears to be required for maximal catalytic activity. Aromatic, nitrogen-containing metal chelators and their nonchelating analogs are effective instantaneous inhibitors of B. nubilusd-lactate dehydrogenase. These compounds bind at the coenzyme binding site, as the mode of inhibition is distinctly competitive with respect to NADH. The different effects of metal chelators and their nonchelator analogs suggest that time-dependent inactivation (chelation of the enzyme zinc ions) and instantaneous inhibition (competition with NADH binding) have independent mechanisms.  相似文献   

20.
Sorbitol dehydrogenase (SDH) is a distant relative to the alcohol dehydrogenases (ADHs) with sequence identities around 20%. SDH is a tetramer with one zinc ion per subunit. We have crystallized rat SDH and determined the structure by molecular replacement using a tetrameric bacterial ADH as search object. The conformation of the bound coenzyme is extended and similar to NADH bound to mammalian ADH but the interactions with the NMN-part have several differences with those of ADH. The active site zinc coordination in SDH is significantly different than in mammalian ADH but similar to the one found in the bacterial tetrameric NADP(H)-dependent ADH of Clostridiim beijerinckii. The substrate cleft is significantly more polar than for mammalian ADH and a number of residues are ideally located to position the sorbitol molecule in the active site. The SDH molecule can be considered to be a dimer of dimers, with subunits A–B and C–D, where the dimer interactions are similar to those in mammalian ADH. The tetramers are composed of two of these dimers, which interact with their surfaces opposite the active site clefts, which are accessible on the opposite side. In contrast to the dimer interactions, the tetramer-forming interactions are small with only few hydrogen bonds between side-chains.  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