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.
The cDNA for the alpha-isoenzyme from rhesus monkey (Macaca mulatta) liver was cloned and expressed in yeast. The alpha-isoenzymes of human and monkey liver alcohol dehydrogenase differ from the other human and horse liver enzymes in having Met57, Ala93, and Val116 instead of Leu57, Phe93, and Leu116 in the substrate binding pocket and Gly47 instead of Arg47 near the pyrophosphate moiety of the coenzyme. The effects of these differences on the kinetic mechanism, substrate specificity, and coenzyme binding were studied with the purified, recombinant monkey alpha-isoenzyme (MmADH alpha) and mutated enzymes with Gly47 substituted with His or Arg. The mechanism appears to be random for the binding of NAD+ and ethanol and ordered for NADH and acetaldehyde, with formation of a dead-end enzyme-NADH-ethanol complex. MmADH alpha reacts 130-fold slower (V/K) with ethanol and 3-25-fold slower with 2-methyl alcohols but 20-fold faster with cyclohexanol, as compared with horse (Equus caballus) liver EE isoenzyme (EqADH). MmADH alpha is stereoselective for the R isomer of 2-butanol, whereas EqADH favors the S isomer. Both enzymes have comparable reactivity with larger primary alcohols. MmADH alpha is more reactive with secondary alcohols and has highest activity with cyclohexanol. However, it does not react with steroids such as 5 beta-androstane-17 beta-ol-3-one. Molecular modeling suggests that the differences between MmADH alpha and EqADH are a result of the substitution of Ala for Phe93 and Thr for Ser48. MmADH alpha binds NAD+ most rapidly when a group with a pK of 7.4 is unprotonated, implicating His51 in this reaction. The G47R substitution decreased the dissociation constants for NAD+ and NADH and turnover numbers only about 2-fold, whereas the G47H substitution increased dissociation constants 7-14-fold and turnover numbers 4-fold. A basic residue at position 47 is not crucial for activity, as multiple interactions determine coenzyme affinity.  相似文献   

2.
Aspartic acid 244 that occurs at the putative NAD(+)-binding site of rat liver S-adenosylhomocysteinase was replaced by glutamic acid by oligonucleotide-directed mutagenesis. The mutant enzyme was purified to homogeneity as judged by sodium dodecyl sulfate-polyacrylamide gel electrophoresis. Gel permeation chromatography showed that the purified mutant enzyme was a tetramer as is the wild-type enzyme. In contrast to the wild-type enzyme, which possesses 1 mol of tightly bound NAD+ per mol of enzyme subunit, the mutant enzyme had only 0.05 mol of NAD+ but contained about 0.6 mol each of NADH and adenine per mol of subunit. The mutant enzyme, after removal of the bound compounds by acid-ammonium sulfate treatment, exhibited S-adenosylhomocysteinase activity when assayed in the presence of NAD+. From the appearance of activity as a function of NAD+ concentration, the enzyme was shown to bind NAD+ with a Kd of 23.0 microM at 25 degrees C, a value greater than 280-fold greater than that of the wild-type enzyme. In the presence of a saturating concentration of NAD+, the mutant enzyme showed apparent Km values for substrates similar to those of the wild-type enzyme. Moderate decreases of 8- and 15-fold were observed in Vmax values for the synthetic and hydrolytic directions, respectively. These results indicate the importance of Asp-244 in binding NAD+, and are consistent with the idea that the region of S-adenosylhomocysteinase from residues 213 to 244 is part of the NAD+ binding site. This region has structural features characteristic of the dinucleotide-binding domains of NAD(+)- and FAD-binding proteins (Ogawa, H., Gomi, T., Mueckler, M. M., Fujioka, M., Backlund, P.S., Jr., Aksamit, R.R., Unson, C.G., and Cantoni, G.L. (1987) Proc. Natl. Acad. Sci. U.S.A. 84, 719-723).  相似文献   

3.
Marohnic CC  Bewley MC  Barber MJ 《Biochemistry》2003,42(38):11170-11182
Microsomal cytochrome b(5) reductase (EC 1.6.2.2) catalyzes the reduction of ferricytochrome b(5) using NADH as the physiological electron donor. Site-directed mutagenesis has been used to engineer the soluble rat cytochrome b(5) reductase diaphorase domain to utilize NADPH as the preferred electron donor. Single and double mutations at residues D239 and F251 were made in a recombinant expression system that corresponded to D239E, S and T, F251R, and Y, D239S/F251R, D239S/F251Y, and D239T/F251R, respectively. Steady-state turnover measurements indicated that D239S/F251Y was bispecific while D239T, D239S/F251R, and D239T/F251R were each NADPH-specific. Wild-type (WT) cytochrome b(5) reductase showed a 3700-fold preference for NADH whereas the mutant with the highest NADPH efficiency, D239T, showed an 11-fold preference for NADPH, a 39200-fold increase. Wild-type cytochrome b(5) reductase only formed a stable charge-transfer complex with NADH while D239T formed complexes with both NADH and NADPH. The rates of hydride ion transfer, determined by stopped-flow kinetics, were k(NADH-WT) = 130 s(-1), k(NADPH-WT) = 5 s(-1), k(NADH-D239T) = 180 s(-1), and k(NADPH-D239T) = 73 s(-1). K(s) determinations by differential spectroscopy demonstrated that D239T could bind nonreducing pyridine nucleotides with a phosphate or a hydroxyl substituent at the 2' position, whereas wild-type cytochrome b(5) reductase would only bind 2' hydroxylated molecules. Oxidation-reduction potentials (E degrees ', n = 2) for the flavin cofactor were WT = -268 mV, D239T = -272 mV, WT+NAD(+) = -190 mV, D239T+NAD(+) = -206 mV, WT+NADP(+) = -253 mV, and D239T+NADP(+) = -215 mV, which demonstrated the thermodynamic contribution of NADP(+) binding to D239T. The crystal structures of D239T and D239T in complex with NAD(+) indicated that the loss of the negative electrostatic surface that precluded 2' phosphate binding in the wild-type enzyme was primarily responsible for the observed improvement in the use of NADPH by the D239T mutant.  相似文献   

4.
An NAD(+)-dependent CDP-D-glucose oxidoreductase which catalyzes the first step of the biosynthesis of CDP-ascarylose (CDP-3,6-dideoxy-L-arabino-hexose), converting CDP-D-glucose to CDP-4-keto-6-deoxy-D-glucose, was isolated from Yersinia pseudotuberculosis. A protocol consisting of DEAE-cellulose, Matrex Blue-A, hydroxylapatite, DEAE-Sephadex, Sephadex G-100, and NAD(+)-agarose column chromatography was used to purify this enzyme 6000-fold to homogeneity. This enzyme consists of two identical subunits, each with a molecular weight of 42,500. Using CDP-D-glucose as the substrate, the Km and Vmax of this catalysis were determined to be 222 microM and 8.3 mumols mg-1 min-1, respectively. Unlike most other oxidoreductases of its class which have a tightly bound NAD+, this highly purified CDP-D-glucose oxidoreductase showed an absolute requirement of NAD+ for its activity. Using chemically synthesized (6S)- and (6R)-CDP-D-[4-2H,6-3H]glucose as substrates, a stereochemical analysis showed this enzymatic reaction involves an intramolecular hydrogen migration from C-4 to C-6, and the displacement of C-6 hydroxyl group by the C-4 hydrogen occurs with inversion. Thus, despite the low cofactor affinity, this enzyme undergoes a mechanism consistent with that followed by other members of its type. Such a mechanistic and stereochemical convergency found for all sugar oxidoreductases so far characterized suggests the presence of a common progenitor of this class of enzyme.  相似文献   

5.
The tetrameric glyceraldehyde-3-phosphate dehydrogenase from rabbit muscle binds NAD+ and some of its analogues in a negatively cooperative manner, whereas other NAD+ analogues bind non-cooperatively to this enzyme. Subsequent to alkylation of a fraction of the active sites of the enzyme with the fluorescent SH reagent N-iodoacetyl-N'-(5-sulfo-1-naphthyl)-ethylenediamine, it was found that the alkylated sites bind NAD+ and NAD+ analogues with a markedly reduced affinity as compared with non-alkylated sites. It was therefore feasible to measure the fluorescence and the circular polarization of the luminescence of the enzyme-bound alkyl groups as a function of binding of NAD+ and of NAD+ analogues to the non-alkylated sites. The changes observed indicate that ligand binding to the non-alkylated sites induces changes in the fluorescence properties of the alkyl groups bound to neighbouring subunits, most likely through the protein moiety. The nature of these changes appears to depend on the structure of the coenzyme analogue. The binding of the non-cooperative binders acetyl-pyridine--adenine dinucleotide, ATP and ADP-ribose induce different conformational changes in the neighbouring vacant subunit, as monitored by the spectroscopic properties of the bound alkyl group. These results in conjunction with other data support the view that the negative cooperativity in NAD+ binding to glyceraldehyde-3-phosphate dehydrogenase results from ligand-induced conformational changes. Furthermore, these results further support the view that subtle structural changes in the coenzyme molecule determine the nature of the conformational changes induced within the enzyme tetramer.  相似文献   

6.
The bicyclic 4-nitroimidazoles PA-824 and OPC-67683 represent a promising novel class of therapeutics for tuberculosis and are currently in phase II clinical development. Both compounds are pro-drugs that are reductively activated by a deazaflavin (F(420)) dependent nitroreductase (Ddn). Herein we describe the biochemical properties of Ddn including the optimal enzymatic turnover conditions and substrate specificity. The preference of the enzyme for the (S) isomer of PA-824 over the (R) isomer is directed by the presence of a long hydrophobic tail. Nitroimidazo-oxazoles bearing only short alkyl substituents at the C-7 position of the oxazole were reduced by Ddn without any stereochemical preference. However, with bulkier substitutions on the tail of the oxazole, Ddn displayed stereospecificity. Ddn mediated metabolism of PA-824 results in the release of reactive nitrogen species. We have employed a direct chemiluminescence based nitric oxide (NO) detection assay to measure the kinetics of NO production by Ddn. Binding affinity of PA-824 to Ddn was monitored through intrinsic fluorescence quenching of the protein facilitating a turnover-independent assessment of affinity. Our results indicate that (R)-PA-824, despite not being turned over by Ddn, binds to the enzyme with the same affinity as the active (S) isomer. This result, in combination with docking studies in the active site, suggests that the (R) isomer probably has a different binding mode than the (S) with the C-3 of the imidazole ring orienting in a non-productive position with respect to the incoming hydride from F(420). The results presented provide insight into the biochemical mechanism of reduction and elucidate structural features important for understanding substrate binding.  相似文献   

7.
Clift MD  Ji H  Deniau GP  O'Hagan D  Silverman RB 《Biochemistry》2007,46(48):13819-13828
Gamma-aminobutyric acid aminotransferase (GABA-AT), a pyridoxal 5'-phosphate dependent enzyme, catalyzes the degradation of the inhibitory neurotransmitter gamma-aminobutyric acid (GABA) to succinic semialdehyde with concomitant conversion of pyridoxal 5'-phosphate (PLP) to pyridoxamine 5'-phosphate (PMP). The enzyme then catalyzes the conversion of alpha-ketoglutarate to the excitatory neurotransmitter L-glutamate. Racemic 4-amino-3-fluorobutanoic acid (3-F-GABA) was shown previously to act as a substrate for GABA-AT, not for transamination, but for HF elimination. Here we report studies of the reaction catalyzed by GABA-AT on (R)- and (S)-3-F-GABA. Neither enantiomer is a substrate for transamination. Very little elimination from the (S)-enantiomer was detected using a coupled enzyme assay; The rate of elimination of HF from the (R)-enantiomer is at least 10 times greater than that for the (S)-enantiomer. The (R)-enantiomer is about 20 times more efficient as a substrate for GABA-AT catalyzed HF elimination than GABA is a substrate for transamination. The (R)-enantiomer also inhibits the transamination of GABA 10 times more effectively than the (S)-enantiomer. Using a combination of computer modeling and the knowledge that vicinal C-F and C-NH3+ bonds have a strong preference to align gauche rather than anti to each other, it is concluded that on binding of free 3-F-GABA to GABA-AT the optimal conformation places the C-NH3+ and C-F bonds gauche in the (R)-enantiomer but anti in the (S)-enantiomer. Furthermore, the dynamic binding process and the bioactive conformation of GABA bound to GABA-AT have been inferred on the basis of the different biological behavior of the two enantiomers of 3-F-GABA when they bind to the enzyme. The present study suggests that the C-F bond can be utilized as a conformational probe to explore the dynamic binding process and provide insight into the bioactive conformation of substrates, which cannot be easily determined by other biophysical approaches.  相似文献   

8.
ATP pyrophosphohydrolase was partially purified from fetal bovine epiphyseal cartilage. The purification was about 10- and 100-fold over the enzyme activities of matrix vesicle fraction and cell homogenate, respectively. The pyrophosphohydrolase and alkaline phosphatase were separated by a sequential application of Sepharose CL-6B and DEAE-cellulose column chromatographies. The purified enzyme migrated as a single band corresponding to the molecular weight of 230,000 in sodium dodecyl sulfate-polyacrylamide disc gel by electrophoresis. The enzyme absolutely required Zn2+ for its activity and appeared to bind Zn2+ strongly with an apparent affinity of p[Zn2+]0.5 = 13.4. The apparent Km for ATP was 0.18 mM. The enzyme was also reactive toward various nucleoside triphosphates including GTP, CTP, and UTP. In contrast, various phosphodiesters including RNA, UDP-glucose, NAD, and bis-p-nitrophenylphosphate were 5% or less as reactive as the nucleoside triphosphates. The pyrophosphohydrolase was inactive toward adenosine 3':5'-monophosphate or various phosphonates. UDP-glucose (1 mM), NAD (1 mM), or RNA (1 mg/ml) failed to inhibit the ATP pyrophosphohydrolase activity. These observations suggest that the ATP pyrophosphohydrolase of the cartilage is probably not a phosphodiesterase I. The matrix vesicle fraction, which probably also included some plasma membrane vesiculated during collagenase digestion, contained the highest specific activity of the enzyme as compared to other subcellular fractions of either epiphyseal or articular cartilage.  相似文献   

9.
Vaccinia DNA topoisomerase forms a covalent DNA-(3'-phosphotyrosyl)-enzyme intermediate at a pentapyrimidine target site 5'-C(+5)C(+4)C(+3)T(+2)T(+1)p downward arrow in duplex DNA. The enzyme engages the target site within a C-shaped protein clamp. Here we mapped the interface of topoisomerase with the DNA minor groove by introducing chiral C-10 R and S 7,8-diol 9,10-epoxide adducts of benzo[a]pyrene (BP) at single N(2)-deoxyguanosine (dG) positions within the nonscissile DNA strand. These trans opened BPdG adducts fit into the minor groove without perturbing helix conformation or base pairing, and the R and S diastereomers are oriented in opposite directions within the minor groove. We measured the effects of the BPdG adducts on the rate and extent of single-turnover DNA transesterification. We observed a sharp margin of interference effects, whereby +5 and -2 BPdG modifications were well tolerated but +4, +3, and -1 BPdG adducts were severely deleterious. Stereoselective effects at the -1 nucleoside (the R isomer interfered, whereas the S isomer did not) delineated at high resolution the downstream border of the minor groove interface. BPdG inhibition of transesterification is likely caused by steric exclusion of constituents of the topoisomerase from the minor groove. We also applied the BPdG interference method to probe the interactions of exonuclease III with the minor groove. DNAs containing these BPdG adducts were protected from digestion by exonuclease III, which was consistently arrested at positions 2-4 nucleotides prior to the BP-modified guanosine.  相似文献   

10.
Experimental conditions favouring the dissociation of tetrameric rabbit muscle D-glyceraldehyde-3-phosphate dehydrogenase into active monomers were elaborated. The urea-induced dissociation of the tetramer was shown to be a stepwise process (in 2 M urea only dimers are formed; an increase in urea concentration up to 3 M causes the splitting of the dimers into monomers). The specific activity of immobilized monomers in the glyceraldehyde-3-phosphate oxidation reaction does not differ from that of the parent immobilized tetrameric form. The tetrameric enzyme molecule binds the coenzyme with a negative cooperativity (the first two NAD+ molecules bind with KD below 0.1 microM; for the third and fourth molecules the dissociation constant was determined to be equal to 5.5 +/- 1.5 microM (50 mM medinal buffer, 10 mM sodium phosphate, pH 8.2). The cooperativity of NAD+ binding is preserved in the immobilized preparation of tetrameric dehydrogenase. The immobilized monomers bind NAD+ with KD of 1.6 +/- 1.0 microM. The experimental results are consistent with the hypothesis according to which the association of catalytically active subunits into a tetramer changes their coenzyme-binding properties in such a way that the first two NAD+ molecules bind more firmly to a tetramer than to a monomer, whereas the third and the fourth NAD+ molecules bind less firmly.  相似文献   

11.
The analogue of NAD+, 4-chloroacetylpyridine-adenine dinucleotide (clac4PdAD+), inactivated the glyceraldehyde-3-phosphate dehydrogenase from sturgeon at a high rate. An affinity labeling was shown to occur with clac4PdAD+. The mononucleotide 4-chloroacetylpyridine 1-beta-D-ribose 5'-phosphate (clac4PdMN+) reacted with the enzyme in a second-order reaction whose rate was much smaller than that calculated for clac4PdAD+ taken as a second-order rate reagent. The rate of the reaction of clac4PdAD+ with the enzyme was determined by stopped flow, using as a probe the long-wavelength absorption maximum (430 nm) formed concomitantly with inactivation of the enzyme. Computer-assisted graphic simulation showed that the clac4PdAD+ analogue could bind to the active site of the enzyme from Bacillus stearothermophilus in a similar manner to that of NAD+, and that the reactive carbon and the reactive thiolate of Cys-149 were within bonding distance. The absorption at 430 nm was linearly proportional to the substoichiometric concentration of clac4PdAD+/mole subunit. Thiol titration suggested the modification of one thiol residue per subunit. The modified thiol was identified by degradation as Cys-149. In contrast to the absorption band generated during the reaction of the 3-chloroacetylpyridine-adenine dinucleotide (clac3PdAD+) with the same enzyme [Eur. J. Biochem. (1982) 127, 519-524; 129, 437-446], enzyme inactivated with clac4PdAD+ and clac4PdMN+ exhibited an absorption maximum at long wavelength which was still present after denaturation. The chromophore is proposed to be the enol form of the alpha-thioether ketone produced by alkylation of the thiolate of Cys-149 by the chloroacetyl group.  相似文献   

12.
Indanol dehydrogenase was purified to apparent homogeneity from monkey liver cytosol. The enzyme was a monomer with a molecular weight of 36,000 and pI of 8.7. The amino acid composition was determined. The enzyme oxidized alicyclic alcohols including transdihydrodiols of benzene and naphthalene in the presence of both NADP+ and NAD+, and reduced several xenobiotic carbonyl compounds in the presence of NADPH, the 4-pro-R hydrogen atom of which was transferred to the substrate. The results of fluorometric binding and kinetic studies are consistent with an ordered sequential mechanism with NADP+ binding first. The enzyme was inhibited competitively versus NADP+ and uncompetitively versus 1-indanol not only by chelating agents such as 1,10-phenanthroline and 2,2'-bipyridine but also by a nonchelating isomer, 4,4'-bipyridine, which suggests hydrophobic interaction of the aromatic compounds with the enzyme, which did not contain zinc. The enzyme was also inhibited by Cibacron blue dye, synthetic estrogens, and delta 4-3-ketosteroids. The inhibition by Cibacron blue was competitive versus NADP+ and noncompetitive versus 1-indanol, whereas those by hexestrol, medroxyprogesterone acetate, and progesterone were uncompetitive versus NADP+ and competitive versus 1-indanol, corraborating the ordered addition of the coenzyme prior to 1-indanol.  相似文献   

13.
An oxidized nicotinamide adenine dinucleotide phosphate/oxidized nicotinamide adenine dinucleotide (NADP+/NAD+) nonspecific L-glutamate dehydrogenase from Bacteroides thetaiotaomicron was purified 40-fold (NADP+ or NAD+ activity) over crude cell extract by heat treatment, (NH4)2SO2 fractionation, diethylaminoethyl-cellulose, Bio-Gel A 1.5m, and hydroxylapatite chromatography. Both NADP+- and NAD+-dependent activities coeluted from all chromatographic treatments. Moreover, a constant ratio of NADP+/NAD+ specific activities was demonstrated at each purification step. Both activities also comigrated in 6% nondenaturing polyacrylamide gels. Affinity chromatography of the 40-fold-purified enzyme using Procion RED HE-3B gave a preparation containing both NADP+- and NAD+-linked activities which showed a single protein band of 48,5000 molecular weight after sodium dodecyl sulfate-polyacrylamide gradient gel electrophoresis. The dual pyridine nucleotide nature of the enzyme was most readily apparent in the oxidative direction. Reductively, the enzyme was 30-fold more active with reduced NADP than with reduced NAD. Nonlinear concave 1/V versus 1/S plots were observed for reduced NADP and NH4Cl. Salts (0.1 M) stimulated the NADP+-linked reaction, inhibited the NAD+-linked reaction, and had little effect on the reduced NADP-dependent reaction. The stimulatory effect of salts (NADP+) was nonspecific, regardless of the anion or cation, whereas the degree of NAD+-linked inhibition decreased in the order to I- greater than Br- greater than Cl- greater than F-. Both NADP+ and NAD+ glutamate dehydrogenase activities were also detected in cell extracts from representative strains of other bacteroides deoxyribonucleic acid homology groups.  相似文献   

14.
We have previously reported that cytochrome P-450LTB in the microsomes of human polymorphonuclear leukocytes (PMN) catalyzes three omega-oxidations of leukotriene B4 (LTB4), leading to the sequential formation of 20-OH-LTB4, 20-CHO-LTB4, and 20-COOH-LTB4 (Soberman, R.J., Sutyak, J.P., Okita, R.T., Wendelborn, D.F., Roberts, L.J., II, and Austen, K. F. (1988) J. Biol. Chem. 263, 7996-8002). The identification of the novel final intermediate, 20-CHO-LTB4, allowed direct analysis of its metabolism by PMN microsomes in the presence of adenine nucleotide cofactors. Microsomes in the presence of 100 microM NAD+ or 100 microM NADP+ converted 1.0 microM 20-CHO-LTB4 to 20-COOH-LTB4 with a Km of 2.4 +/- 0.8 microM (mean +/- S.E., n = 4) and a Vmax of 813.9 +/- 136.6 pmol.min-1.mg-1, for NAD+, as compared to 0.12 microM and 5.0 pmol.min-1.mg-1 (n = 2) for NADPH as a cofactor. The conversion of 1.0 microM of 20-CHO-LTB4 to 20-COOH-LTB4 in the presence of saturating concentrations (1.0 mM) of both NAD+ and NADP+ was not greater than the reaction in the presence of 1.0 mM of each cofactor separately, indicating that NAD+ and NADP+ were cofactors for the same enzyme. Antibody to cytochrome P-450 reductase did not inhibit the conversion of 20-CHO-LTB4 to 20-COOH-LTB4. When 1.0 microM 20-OH-LTB4 was added to microsomes in the presence of NADPH, approximately three-fourths of the product formed (63.7 +/- 5.1 pmol; mean +/- S.E., n = 3) was 20-CHO-LTB4 and approximately one-fourth (21.3 +/- 3.9 pmol; mean +/- S.E., n = 3) was 20-COOH-LTB4. In the presence of both NADPH and NAD+, only 20-COOH-LTB4 (85.5 +/- 9.9 pmol; mean +/- S.E., n = 3) was formed. PMN microsomes also contain an NADH-dependent aldehyde reductase which converts 20-CHO-LTB4 to 20-OH-LTB4, a member of the LTB4 family of molecules with biological activity. Based upon kinetic, cofactor and inhibition data, microsomal aldehyde dehydrogenase preferentially regulates the final and irreversible inactivation step in the LTB4 metabolic sequence.  相似文献   

15.
Geraniol dehydrogenase (GeDH), which plays an important role in the biosynthesis of neral, an alarm pheromone, was purified from the astigmatid mite Carpoglyphus lactis. The enzyme was obtained in an apparently homogeneous and active form after 1879-fold purification through seven steps of chromatography. Car. lactis GeDH was determined to be a monomer in its active form with a relative molecular mass of 42 800, which is a unique subunit structure in comparison with already established alcohol dehydrogenases. Car. lactis GeDH oxidized geraniol into geranial in the presence of NAD+. NADP+ was ineffective as a cofactor, suggesting that Car. lactis GeDH is an NAD+-dependent alcohol dehydrogenase. The optimal pH and temperature for geraniol oxidation were determined to be pH 9.0 and 25 degrees C, respectively. The Km values for geraniol and NAD+ were 51.0 microm and 59.5 microm, respectively. Car. lactis GeDH was shown to selectively oxidize geraniol, whereas its geometrical isomer, nerol, was inert as a substrate. The high specificity for geraniol suggests that Car. lactis GeDH specializes in the alarm pheromone biosynthesis of Car. lactis. Car. lactis GeDH is composed of 378 amino acids. Structurally, Car. lactis GeDH showed homology with zinc-dependent alcohol dehydrogenases found in mammals and a mosquito (36.6-37.6% identical), and the enzyme was considered to be a member of the medium-chain dehydrogenase/reductase family, in view of the highly conserved sequences of zinc-binding and NAD+-binding sites. Phylogenetic analyses indicate that Car. lactis GeDH could be categorized as a new class, different from other established alcohol dehydrogenases.  相似文献   

16.
1. The seven isomeric optically inactive forms of methylcyclohexanol (i.e. 1-, and cis- and trans-2-, -3- and -4-) are excreted by rabbits mainly as glucuronides of the thermodynamically more stable forms of the alcohols. The eighth isomer, cyclohexylmethanol, however, undergoes aromatization in vivo, giving rise to benzoic acid and hippuric acid. The (±)-2-, (±-3- and 4-methylcyclohexanones are reduced in the rabbit and excreted mainly as the glucuronides of the thermodynamically more stable forms of the corresponding methylcyclohexanols. 2. Racemic cis- and trans-2-methylcyclohexanol and 2-methylcyclohexanone are all excreted as conjugated trans-2-methylcyclohexanol. However, when the (±)-cis-alcohol or the (±)-ketone is fed, (+)-trans-2-methylcyclohexanol is excreted, whereas when the (±)-trans-alcohol is fed it is excreted as the (±)-trans-alcohol. 3. Racemic cis- and trans-3-methylcyclohexanol and 3-methylcyclohexanone are all excreted as conjugated racemic cis-3-methylcyclohexanol. cis- and trans-4-Methylcyclohexanol and 4-methylcyclohexanone are all excreted as conjugated trans-4-methylcyclohexanol. 4. The metabolic differences between the various methylcyclohexanols are explicable in terms of their conformations and of Vennesland's (1958) hypothesis of the role of NADH in dehydrogenation reactions.  相似文献   

17.
Karsten WE  Cook PF 《Biochemistry》2007,46(50):14578-14588
The NAD-malic enzyme catalyzes the oxidative decarboxylation of l-malate. Structures of the enzyme indicate that arginine 181 (R181) is within hydrogen bonding distance of the 1-carboxylate of malate in the active site of the enzyme and interacts with the carboxamide side chain of the nicotinamide ring of NADH, but not with NAD+. Data suggested R181 might play a central role in binding and catalysis in malic enzyme, and it was thus changed to lysine and glutamine to probe its potential function. A nearly 100-fold increase in the Km for malate and a 30-fold increase in the Ki for oxalate, an analogue of the enolpyruvate intermediate, in the R181Q and R181K mutants are consistent with a role for R181 in binding substrates. The mutant enzymes also exhibit a >10-fold increase in KiNADH, but only a slight or no change in KNAD, consistent with rotation of the nicotinamide ring into the malate binding site upon reduction of NAD+ to NADH. The activity of the R181Q mutant can be rescued by ammonium ion likely by binding in the pocket vacated by the guanidinium group of R181. Results suggest 2 mol of ammonia bind per mole of active sites with a high-affinity KNH4 of 0.7 +/- 0.1 mM and a low-affinity KNH4 of approximately 420 mM. Occupancy of the high-affinity site, likely by NH4+, results in an increase in the affinity of malate, oxalate, and NADH (with no change in NAD affinity), consistent with the above-proposed roles for R181. The second molecule to bind is likely neutral NH3, and its binding increases V/Et approximately 20-fold. Primary deuterium and 13C isotope effects measured in the absence and presence of ammonium ion suggest R181Q predominantly affects the rate of the reaction by changing the rate of the precatalytic conformational change. The isotope effects do not change upon binding the second mole of ammonia in spite of the 20-fold increase in V/Et. Thus, the R181Q mutant enzyme exists as an equilibrium mixture between active and less active forms, and NH3 stabilizes the more active conformation of the enzyme.  相似文献   

18.
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   

19.
S Chen  T D Lee  K Legesse  J E Shively 《Biochemistry》1986,25(19):5391-5395
We have identified the site labeled by arylazido-beta-alanyl-NAD+ (A3'-O-(3-[N-(4-azido-2-nitrophenyl)amino]propionyl)NAD+) in rabbit muscle glyceraldehyde-3-phosphate dehydrogenase by microsequencing and fast atom bombardment mass spectrometry. This NAD+ photoaffinity analogue has been previously demonstrated to modify glyceraldehyde-3-phosphate dehydrogenase in a very specific manner and probably at the active site of the enzyme [Chen, S., Davis, H., Vierra, J. R., & Guillory, R. J. (1984) Biochem. Biophys. Stud. Proteins Nucleic Acids, Proc. Int. Symp., 3rd, 407-425]. The label is associated exclusively with a tryptic peptide that has the sequence Ile-Val-Ser-Asn-Ala-Ser-Cys-Thr-Thr-Asn. In comparison to the amino acid sequence of glyceraldehyde-3-phosphate dehydrogenase from other species, this peptide is in a highly conserved region and is part of the active site of the enzyme. The cysteine residue at position seven was predominantly labeled and suggested to be the site modified by arylazido-beta-alanyl-NAD+. This cysteine residue corresponds to the Cys-149 in the pig muscle enzyme, which has been shown to be an essential residue for the enzyme activity. The present investigation clearly demonstrates that arylazido-beta-alanyl-NAD+ is a useful photoaffinity probe to characterize the active sites of NAD(H)-dependent enzymes.  相似文献   

20.
R(-) isomer of diniconazole [S-3308L, (E)-1-(2,4-dichlorophenyl)-4,4-dimethyl-2-(1,2,4-triazol-l-yl)-1-+ ++penten-3-ol], a newly developed fungicide strongly inhibited lanosterol 14 alpha-demethylation catalyzed by a yeast cytochrome P-450 (P-450/14DM). On the other hand, S(+) isomer of diniconazole was a weaker inhibitor for P-450/14DM. The R(-) isomer combined with both ferric and ferrous P-450/14DM and interfered binding of CO to the cytochrome. The S(+) isomer also interacted with both forms of P-450/14DM but the absorption spectra of the S(+)-diniconazole complexes were different from those of the R(-)-diniconazole complexes. Furthermore, S(+) isomer did not significantly interfere the binding of CO to P-450/14DM. These observations suggest that P-450/14DM discriminates enantiomers of diniconazole and the R(-) isomer is more favorably fit for the active site of the cytochrome.  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