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.
An effective means of relieving the toxicity of furan aldehydes, furfural (FFA) and 5-hydroxymethylfurfural (HMF), on fermenting organisms is essential for achieving efficient fermentation of lignocellulosic biomass to ethanol and other products. Ari1p, an aldehyde reductase from Saccharomyces cerevisiae, has been shown to mitigate the toxicity of FFA and HMF by catalyzing the NADPH-dependent conversion to corresponding alcohols, furfuryl alcohol (FFOH) and 5-hydroxymethylfurfuryl alcohol (HMFOH). At pH 7.0 and 25°C, purified Ari1p catalyzes the NADPH-dependent reduction of substrates with the following values (k(cat) (s(-1)), k(cat)/K(m) (s(-1)mM(-1)), K(m) (mM)): FFA (23.3, 1.82, 12.8), HMF (4.08, 0.173, 23.6), and dl-glyceraldehyde (2.40, 0.0650, 37.0). When acting on HMF and dl-glyceraldehyde, the enzyme operates through an equilibrium ordered kinetic mechanism. In the physiological direction of the reaction, NADPH binds first and NADP(+) dissociates from the enzyme last, demonstrated by k(cat) of HMF and dl-glyceraldehyde that are independent of [NADPH] and (K(ia)(NADPH)/k(cat)) that extrapolate to zero at saturating HMF or dl-glyceraldehyde concentration. Microscopic kinetic parameters were determined for the HMF reaction (HMF+NADPH?HMFOH+NADP(+)), by applying steady-state, presteady-state, kinetic isotope effects, and dynamic modeling methods. Release of products, HMFOH and NADP(+), is 84% rate limiting to k(cat) in the forward direction. Equilibrium constants, [NADP(+)][FFOH]/[NADPH][FFA][H(+)]=5600×10(7)M(-1) and [NADP(+)][HMFOH]/[NADPH][HMF][H(+)]=4200×10(7)M(-1), favor the physiological direction mirrored by the slowness of hydride transfer in the non-physiological direction, NADP(+)-dependent oxidation of alcohols (k(cat) (s(-1)), k(cat)/K(m) (s(-1)mM(-1)), K(m) (mM)): FFOH (0.221, 0.00158, 140) and HMFOH (0.0105, 0.000104, 101).  相似文献   

2.
从杜氏盐藻分离得到的二羟丙酮还原酶能专一性地催化二羟丙酮和甘油之间的可逆反应。酶催化二羟丙酮还原及甘油氧化的最适 PH分别为7.5和9.0;藻细胞经高渗处理,其甘油含量增加,酶催化甘油合成的活性比处理前提高120%,且大于其催化甘油转化的活性;藻细胞经低渗处理,其甘油含量减少,酶催化甘油转化的速率比处理前提高32%,暗示二羟丙酮还原酶在杜氏盐藻渗透调节过程中是甘油合成或转化的一个关键酶。  相似文献   

3.
Fungal metabolism of biphenyl.   总被引:9,自引:0,他引:9       下载免费PDF全文
gamma-Glutamyl phosphate reductase, the second enzyme of proline biosynthesis, catalyses the formation of l-glutamic acid 5-semialdehyde from gamma-glutamyl phosphate with NAD(P)H as cofactor. It was purified 150-fold from crude extracts of Pseudomonas aeruginosa PAO 1 by DEAE-cellulose chromatography and hydroxyapatite adsorption chromatography. The partially purified preparation, when assayed in the reverse of the biosynthetic direction, utilized l-1-pyrroline-5-carboxylic acid as substrate and reduced NAD(P)(+). The apparent K(m) values were: NAD(+), 0.36mm; NADP(+), 0.31mm; l-1-pyrroline-5-carboxylic acid, 4mm with NADP(+) and 8mm with NAD(+); P(i), 28mm. 3-(Phosphonoacetylamido)-l-alanine, a structural analogue of gamma-glutamyl phosphate, inhibited this enzyme competitively (K(i)=7mm). 1-Pyrroline-5-carboxylate reductase (EC 1.5.1.2), the third enzyme of proline biosynthesis, was purified 56-fold by (NH(4))(2)SO(4) fractionation, Sephadex G-150 gel filtration and DEAE-cellulose chromatography. It reduced l-1-pyrroline-5-carboxylate with NAD(P)H as a cofactor to l-proline. NADH (K(m)=0.05mm) was a better substrate than NADPH (K(m)=0.02mm). The apparent K(m) values for l-1-pyrroline-5-carboxylate were 0.12mm with NADPH and 0.09mm with NADH. The 3-acetylpyridine analogue of NAD(+) at 2mm caused 95% inhibition of the enzyme, which was also inhibited by thio-NAD(P)(+), heavy-metal ions and thiol-blocking reagents. In cells of strain PAO 1 grown on a proline-medium the activity of gamma-glutamyl kinase and gamma-glutamyl phosphate reductase was about 40% lower than in cells grown on a glutamate medium. No repressive effect of proline on 1-pyrroline-5-carboxylate reductase was observed.  相似文献   

4.
Nitrate reductase (NR) from the yeast, Rhodotorula glutinis var. salinaria was composed of two enzymatic components, diaphorase and terminal nitrate reducing moieties. The enzyme used NADPH as electron donor and FAD as cofactor. The synthesis of nitrate reductase was promoted specifically by nitrate and repressed by ammonium and amino acids. Nitrate reductase from this yeast had an inactive as well as an active form. Inactive enzyme was reactivated by oxidation with ferricyanide in vitro. Hydroxylamine promoted the formation of inactive enzyme in vivo. Ammonium could neither promote the inactivation nor reduce the total level of nitrate reductase activity. Nitrate could protect nitrate reductase from inactivation caused by nitrogen starvation or hydroxylamine.  相似文献   

5.
Nitrate reductase of the salt-tolerant alga Dunaliella parva could utilize NADPH as well as NADH as an electron donor. The two pyridine nucleotide-dependent activities could not be separated by either ion exchange chromatography on DEAE-cellulose or gel filtration on Sepharose 4B. The NADPH-dependent activity was not inhibited by phosphatase inhibitors. NADPH was not hydrolyzed to NADH and inorganic phosphate in the course of nitrate reduction. Reduction of nitrate in vitro could be coupled to a NADPH-regenerating system of glycerol and NADP-dependent glycerol dehydrogenase. It is concluded that the nitrate reductase of D. parva will function with NADPH as well as NADH. This is a unique characteristic not common to most algae.  相似文献   

6.
A d-galacturonic acid reductase and the corresponding gene were identified from the mold Hypocrea jecorina (Trichoderma reesei). We hypothesize that the enzyme is part of a fungal d-galacturonic acid catabolic pathway which has not been described previously and which is distinctly different from the bacterial pathway. H. jecorina grown on d-galacturonic acid exhibits an NADPH-dependent d-galacturonic acid reductase activity. This activity is absent when the mold is grown on other carbon sources. The d-galacturonic acid reductase was purified, and tryptic digests of the purified protein were sequenced. The open reading frame of the corresponding gene was then cloned from a cDNA library. The open reading frame was functionally expressed in the yeast Saccharomyces cerevisiae. A histidine-tagged protein was purified, and the enzyme kinetics were characterized. The enzyme converts in a reversible reaction from d-galacturonic acid and NADPH to l-galactonic acid and NADP. The enzyme also exhibits activity with d-glucuronic acid and dl-glyceraldehyde.  相似文献   

7.
Glycerol can be oxidized by rat liver microsomes to formaldehyde in a reaction that requires the production of reactive oxygen intermediates. Studies with inhibitors, antibodies, and reconstituted systems with purified cytochrome P4502E1 were carried out to evaluate whether P450 was required for glycerol oxidation. A purified system containing phospholipid, NADPH-cytochrome P450 reductase, P4502E1, and NADPH oxidized glycerol to formaldehyde. Formaldehyde production was dependent on NADPH, reductase, and P450, but not phospholipid. Formaldehyde production was inhibited by substrates and ligands for P4502E1, as well as by anti-pyrazole P4502E1 IgG. The oxidation of glycerol by the reconstituted system was sensitive to catalase, desferrioxamine, and EDTA but not to superoxide dismutase or mannitol, indicating a role for H2O2 plus non-heme iron, but not superoxide or hydroxyl radical in the overall glycerol oxidation pathway. The requirement for reactive oxygen intermediates for glycerol oxidation is in contrast to the oxidation of typical substrates for P450. In microsomes from pyrazole-treated, but not phenobarbital-treated rats, glycerol oxidation was inhibited by anti-pyrazole P450 IgG, anti-hamster ethanol-induced P450 IgG, and monoclonal antibody to ethanol-induced P450, although to a lesser extent than inhibition of dimethylnitrosamine oxidation. Anti-rabbit P4503a IgG did not inhibit glycerol oxidation at concentrations that inhibited oxidation of dimethylnitrosamine. Inhibition of glycerol oxidation by antibodies and by aminotriazole and miconazole was closely associated with inhibition of H2O2 production. These results indicate that P450 is required for glycerol oxidation to formaldehyde; however, glycerol is not a direct substrate for oxidation to formaldehyde by P450 but is a substrate for an oxidant derived from interaction of iron with H2O2 generated by cytochrome P450.  相似文献   

8.
Oxidative deformylation of 4-hydroxy[14C]methylene-5alpha-cholest-7-en-3-one and oxidative demethylation of [30,31-14C]4,4-dimethyl-5alpha-cholest-7-en-3beta-ol by rat liver microsomes have been compared with regard to the manner in which electrons are introduced from both NADH and NADPH. Evidence suggests that NADH and NADPH support oxidation of both substrates via separate routes of electron transfer. Thus, 10 micron cytochrome c will inhibit NADPH-supported oxidation to 40 to 50% of control activity leaving NADH-supported oxidation unaffected. Also, treatment of microsomes with subtilisin diminishes NADPH-supported oxidation to 10 to 30% of control activity for either substrate to 70 to 90% of control activity while NADH-supported oxidative activity is virtually unaffected. Studies on the oxidase activities and NADPH-cytochrome c reductase as well as NADH-ferricyanide reductase have shown marked differences in activity in the presence of inhibitors. Thus, 9 mM 2'-AMP inhibits NADPH-cytochrome c reductase to 10 to 20% of control activity while NADPH-supported oxidative demethyl ation and deformylation are essentially unchanged. Mersalyl at 15 to 25 nmol/mg of microsomal protein inhibits both reductases to 20 to 40% of control activity; oxidative demethylation is unaffected and oxidative deformylation stimulated slightly when NADPH is used. Finally, antibody to NADPH-cytochrome c reductase inhibits oxidase activity for either substrate to 70 to 90% of control activity while reductase activity is inhibited to 10 to 30% of control activity.  相似文献   

9.
1. NADPH-dependent nitrite reductase from the leaves of higher plants was purified at least 70-fold and separated into two enzyme fractions. The first enzyme, a diaphorase with ferredoxin-NADP-reductase activity, is required only to transfer electrons from NADPH to a suitable electron acceptor, which then donates electrons to nitrite reductase proper. 2. Purified nitrite reductase accepted electrons from ferredoxin (the natural donor) or from reduced dyes. Ferredoxin was reduced by illuminated chloroplasts or dithionite, or by NADPH when diaphorase was present. The purified enzyme did not accept electrons directly from NADPH. 3. Ferredoxins purified from maize, spinach or Clostridium were interchangeable in the nitrite-reductase system. 4. Nitrite reductase had K(m) 0.15mm for nitrite. The pH optimum varied with plant and method of assay. The preparation had low sulphite-reductase activity. Ammonia was the product of nitrite reduction. 5. For some plants, the assay of crude preparations with NADPH was limited by diaphorase and the addition of diaphorase gave a better estimate of nitrite-reductase activity. A simple method of assay is described that uses dithionite with benzyl viologen as electron donor.  相似文献   

10.
Pig kidney aldehyde reductase is inactivated by 2,3-butanedione, phenylglyoxal, methylglyoxal, and 1,2-cyclohexanedione. 2,3-Butanedione caused the most rapid loss in enzyme activity, the rate of loss being proportional to the concentration of 2,3-butanedione. Neither D-glyceraldehyde nor pyridine 3-aldehyde, both substrates for this broadly specific enzyme, protected the enzyme from inactivation but 1 mM NADPH or NADP completely prevented the loss of activity by 2,3-butanedione suggesting the involvement of arginine in the binding of cofactor. Nicotinamide mononucleotide (NMN) (reduced form) offered no protection to inactivation whereas ADP-ribose phosphate gave complete protection indicating that it is the latter portion of NADPH which interacts with the essential arginine. Both NMN and ADP-ribose phosphate are competitive inhibitors of aldehyde reductase with respect to NADPH. Butanedione-modified aldehyde reductase could still bind to a blue dextran-Sepharose 4B column suggesting that the modified arginine did not bind NADPH. This was confirmed by fluorescence spectra which showed that chemically modified aldehyde reductase caused the same blue shift of NADPH fluorescence as did native aldehyde reductase. Of additional interest was the quenching of NADPH fluorescence by aldehyde reductase which, with one exception, is in contrast to the fluorescence behavior of all other oxidoreductases.  相似文献   

11.
Individual flowers from inflorescences of Dahlia variabilis (cv Scarlet Star) in young developmental stages contained relatively high activity of (+)-dihydroflavonol (DHF) 4-reductase. The DHF reductase was purified from such flowers to apparent homogeneity by a five-step procedure. This included affinity adsorption on Blue Sepharose and elution of the enzyme with NADP+. By gel filtration and by sodium dodecyl sulfate-polyacrylamide gel electrophoresis it was shown that DHF reductase contains only one polypeptide chain with a Mr of about 41,000. The reductase required NADPH as cofactor and catalyzed transfer of the pro-S hydrogen of NADPH to the substrate. Flavanones and dihydroflavonols (3-hydroxyflavanones) were substrates for DHF reductase with pH optima of about 6.0 for flavanones and of about 6.8 for dihydroflavonols. Flavanones were reduced to the corresponding flavan-4-ols and (+)-dihydroflavonols to flavan-3,4-cis-diols. Apparent Michaelis constants determined for (2S)-naringenin, (2S)-eriodicytol, (+)-dihydrokaempferol, (+)-dihydroquercetin, and NADPH were, respectively, 2.3, 2, 10, 15, and 42 microM. V/Km values were higher for dihydroflavonols than for flavanones. Conversion of dihydromyricetin to leucodelphinidin was also catalyzed by the enzyme at a low rate, whereas flavones and flavonols were not accepted as substrates. DHF reductase was not inhibited by metal chelators.  相似文献   

12.
β-Ketoacyl-[acyl-carrier-protein] (ACP) reductase, β-hydroxyacyl-ACP dehydrase, and enoyl-ACP reductase have been purified to homogeneity from extracts of spinach leaves. Based on sodium dodecyl sulfate-polyacrylamide gel eletrophoresis studies, the monomeric molecular weights of the β-ketoacyl-ACP reductase, β-hydroxyacyl-ACP dehydrase, and enoyl-ACP reductase were 24,200, 19,000, and 32,500, respectively, and by gel filtration, their molecular weights were 97,000, 85,000, and 115,000, respectively, suggesting that these three enzymes exist as tetramers. The β-ketoacyl-ACP reductase, the β-hydroxyacyl-ACP dehydrase, and the enoyl-ACP reductase contained two, one, and two cystein residues per monomer. β-Ketoacyl-ACP reductase preferably utilized NADPH as the reductant, whereas enoyl-ACP reductase was absolutely specific to NADH. β-Ketoacyl-ACP reductase reversibly catalyzed the reduction of acetoacetyl-ACP to d-β-hydroxybutyryl-ACP and β-hydroxyacyl-ACP dehydrase catalyzed the dehydration of d-β-hydroxyacyl-ACP to 2-enoyl-ACP. Both β-hydroxyacyl-ACP dehydrase and enoyl-ACP reductase were active with 2-enoyl-ACPs having chain lengths from C4 to C16, with 2-hexenoyl-ACP and 2-octenoyl-ACP being the most effective substrate. CoA esters served as substrates with the β-ketoacyl-ACP reductase and the enoyl-ACP reductase but were inert with β-hydroxyacyl-ACP dehydrase. These enzymes were inhibited by p-chloromercuribenzoate but not by N-ethylmaleimide.  相似文献   

13.
delta 1-Pyrroline-5-carboxylate reductase (L-proline:NAD(P)+ 5-oxidoreductase, EC 1.5.1.2) has been purified from rat lens and biochemically characterized. Purification steps included ammonium sulfate fractionation, affinity chromatography on Amicon Matrex Orange A, and gel filtration with Sephadex G-200. These steps were carried out at ambient temperature (22 degrees C) in 20 mM sodium phosphate/potassium phosphate buffer (pH 7.5) containing 10% glycerol, 7 mM mercaptoethanol and 0.5 mM EDTA. The enzyme, purified to apparent homogeneity, displayed a molecular weight of 240 000 by gel chromatography and 30 000 by SDS-polyacrylamide gel electrophoresis. This suggests that the enzyme is composed of eight subunits. The purified enzyme displays a pH optimum between 6.5 and 7.1 and is inhibited by heavy metal ions and p-chloromercuribenzoate. Kinetic studies indicated Km values of 0.62 mM and 0.051 mM for DL-pyrroline-5-carboxylate as substrate when NADH and NADPH respectively were employed as cofactors. The Km values for the cofactors NADH and NADPH with DL-pyrroline-5-carboxylate as substrate were 0.37 mM and 0.006 mM, respectively. With L-pyrroline-5-carboxylate as substrate, Km values of 0.21 mM and 0.022 mM were obtained for NADH and NADPH, respectively. Enzyme activity is potentially inhibited by NADP+ and ATP, suggesting that delta 1-pyrroline-5-carboxylate reductase may be regulated by the energy level and redox state of the lens.  相似文献   

14.
Pyridine nucleotide specificity of barley nitrate reductase   总被引:6,自引:4,他引:2       下载免费PDF全文
Dailey FA  Kuo T  Warner RL 《Plant physiology》1982,69(5):1196-1199
NADPH nitrate reductase activity in higher plants has been attributed to the presence of NAD(P)H bispecific nitrate reductases and to the presence of phosphatases capable of hydrolyzing NADPH to NADH. To determine which of these conditions exist in barley (Hordeum vulgare L. cv. Steptoe), we characterized the NADH and NADPH nitrate reductase activities in crude and affinity-chromatography-purified enzyme preparations. The pH optima were 7.5 for NADH and 6 to 6.5 for the NADPH nitrate reductase activities. The ratio of NADPH to NADH nitrate reductase activities was much greater in crude extracts than it was in a purified enzyme preparation. However, this difference was eliminated when the NADPH assays were conducted in the presence of lactate dehydrogenase and pyruvate to eliminate NADH competitively. The addition of lactate dehydrogenase and pyruvate to NADPH nitrate reductase assay media eliminated 80 to 95% of the NADPH nitrate reductase activity in crude extracts. These results suggest that a substantial portion of the NADPH nitrate reductase activity in barley crude extracts results from enzyme(s) capable of converting NADPH to NADH. This conversion may be due to a phosphatase, since phosphate and fluoride inhibited NADPH nitrate reductase activity to a greater extent than the NADH activity. The NADPH activity of the purified nitrate reductase appears to be an inherent property of the barley enzyme, because it was not affected by lactate dehydrogenase and pyruvate. Furthermore, inorganic phosphate did not accumulate in the assay media, indicating that NADPH was not converted to NADH. The wild type barley nitrate reductase is a NADH-specific enzyme with a slight capacity to use NADPH.  相似文献   

15.
Two dihydroxyacetone kinases (DHAKs), DHAK I and DHAK II, were purified to homogeneity from Schizosaccharomyces pombe IFO 0354. They were immunologically different from each other. Although both of the enzymes had some affinity for glycerol and dl-glyceraldehyde in addition to dihydroxyacetone and glyceraldehyde, V(infmax) values for dihydroxyacetone were much higher than those for glycerol and dl-glyceraldehyde. On the basis of the K(infm) values of both enzymes for dihydroxyacetone, DHAK II plays a more important role than DHAK I in dissimilation of glycerol via dihydroxyacetone.  相似文献   

16.
The reaction of pentaerythritol tetranitrate reductase with reducing and oxidizing substrates has been studied by stopped-flow spectrophotometry, redox potentiometry, and X-ray crystallography. We show in the reductive half-reaction of pentaerythritol tetranitrate (PETN) reductase that NADPH binds to form an enzyme-NADPH charge transfer intermediate prior to hydride transfer from the nicotinamide coenzyme to FMN. In the oxidative half-reaction, the two-electron-reduced enzyme reacts with several substrates including nitroester explosives (glycerol trinitrate and PETN), nitroaromatic explosives (trinitrotoluene (TNT) and picric acid), and alpha,beta-unsaturated carbonyl compounds (2-cyclohexenone). Oxidation of the flavin by the nitroaromatic substrate TNT is kinetically indistinguishable from formation of its hydride-Meisenheimer complex, consistent with a mechanism involving direct nucleophilic attack by hydride from the flavin N5 atom at the electron-deficient aromatic nucleus of the substrate. The crystal structures of complexes of the oxidized enzyme bound to picric acid and TNT are consistent with direct hydride transfer from the reduced flavin to nitroaromatic substrates. The mode of binding the inhibitor 2,4-dinitrophenol (2,4-DNP) is similar to that observed with picric acid and TNT. In this position, however, the aromatic nucleus is not activated for hydride transfer from the flavin N5 atom, thus accounting for the lack of reactivity with 2,4-DNP. Our work with PETN reductase establishes further a close relationship to the Old Yellow Enzyme family of proteins but at the same time highlights important differences compared with the reactivity of Old Yellow Enzyme. Our studies provide a structural and mechanistic rationale for the ability of PETN reductase to react with the nitroaromatic explosive compounds TNT and picric acid and for the inhibition of enzyme activity with 2,4-DNP.  相似文献   

17.
NADPH cytochrome c (P-450) reductase was purified from human placental microsomes using a combination of affinity and gel filtration chromatography. Affinity chromatography using agarose-hexane-adenosine 2'5 diphosphate resulted in two protein bands being detected by SDS-PAGE of approximate MwS 68 and 75 kDa. Fractions containing the two proteins were pooled, and then resolved using Sephacryl S-200. Both of the purified proteins displayed enzyme activity, measured by their ability to reduce cytochrome c. The 75 kDa protein obtained was used to immunize three female New Zealand white rabbits. The IgG fraction was partly purified from rabbit sera which suppressed placental microsomal NADPH cytochrome c reductase activity by > 80% using 33% ammonium sulphate. The procured antibody suppressed androstenedione aromatase activity in microsomal preparations of human placental and breast adipose tissue, and NADPH cytochrome c reductase activity in prostate (benign and malignant), MDA-MB-231 breast cancer cells, breast adipose, Hep G2 hepatoma cells and placental microsomal preparations. The extent of NADPH cytochrome c reductase inhibition varied in the order of malignant prostate < benign prostate < MDA < breast adipose < Hep G2 < placenta. The results suggest that human placental NADPH cytochrome c (P-450) reductase shares common antigenic epitopes pertinent to its capability of reducing cytochrome c in all of the above-mentioned tissues. In attempting to associate possible changes in NADPH cytochrome c reductase activity imposed by neoplasia to the obtained immunochemical cross reactivity and enzyme activity results, it was noted that microsomes obtained from MDA cells exhibited enzyme activity significantly less than that of breast adipose microsomes (1.6 and 8.1 nmol/min/mg protein, respectively) and by comparison showed 6% less homology towards the placental antibody. The results obtained for benign and malignant prostate showed no significant difference between the neoplastic states as adjudged by enzyme activity and immunochemical assays.  相似文献   

18.
Glutathione reductase from rat liver has been purified greater than 5000-fold in a yield of 20%. The molecular weights of the enzyme and its subunits were estimated to be 125,000 and 60,000, respectively, indicating that the native enzyme is a dimer. The enzyme molecular contains 2 FAD molecules, which are reducible by NADPH, GSH or dithioerythritol. The reduced flavin is instantaneously reoxidized by addition of GSSG. The steady state kinetic data are consistent with a branching reaction mechanism previously proposed for glutathione reductase from yeast (MANNERVIK, B. (1973) Biochem. Biophy. Res. Commun. 53, 1151-1158). This mechanism is also favored by the nonlinear inhibition pattern produced by NADP-+. However, at low GSSG concentrations the rate equation can be approximated by that of a simple ping pong mechanism. NADPH and the mixed disulfide of coenzyme A and GSH were about 10% as active as NADPH and GSSG, respectively, whereas some sulfenyl derivatives related to GSSG were less active as substrates. The pH activity profiles of these substrates differed from that of the NADPH-GSSG substrate pair.  相似文献   

19.
Effects of twelve flavonoids and five catechins as well as gallic acid on two kinds of glutathione-related enzymes were investigated. Glutathione 5-transferase (EC 2.5.1.18) activity was measured by S-2,4-dinitrophenyl glutathione formation from 1-chloro-2,4-dinitrobenzene and reduced glutathione. Glutathione reductase (EC 1.6.4.2) activity was followed by NADPH dehydrogenation. Fisetin and myricetin were potent inhibitors of glutathione S-transferase, while kaempferol, quercetin, baicalein, and quercitrin were medium inhibitors. Epicatechin gallate and epigallocatechin gallate also showed medium inhibition. Kinetic analyses indicated that fisetin was a mixed type inhibitor of glutathione S-transferase with respect to both substrates, while myricetin was a competitive inhibitor of the same enzyme with both substrates. Fisetin and myricetin were noncompetive inhibitors of glutathione reductase with both NADPH and oxidized glutathione. The inhibition patterns of GT and GR as well as the results of kinetic analyses indicated a possibility that inhibitory flavonoids might have some influence on the glutathione recognition sites of the two enzymes.  相似文献   

20.
An NAD-dependent glycerol 3-phosphate dehydrogenase (sn-glycerol 3-phosphate: NAD oxidoreductase; EC 1.1.1.8) has been purified from spinach leaves by a three-step procedure involving ion-exchange, gel filtration, and affinity chromatography. The enzyme has been purified over 10,000-fold to a specific activity of 38. It has a molecular weight of approximately 63,500. The pH optimum for the reduction of dihydroxyacetone phosphate is 6.8 and for glycerol 3-phosphate oxidation it is 9.5. During dihydroxyacetone phosphate reduction hyperbolic kinetics were observed when either NADH or dihydroxyacetone phosphate was the variable substrate, but concentrations of NADH greater than 150 μm were inhibitory. Michaelis constants were 0.30–0.35 mm for dihydroxyacetone phosphate and 0.01 mm for NADH. Glycerol 3-phosphate oxidation obeyed Michaelis-Menten kinetics with a Km of 0.19 mm for NAD and 1.6 mm for glycerol 3-phosphate. The enzyme was specific for those substrates, and dihydroxyacetone, glyceraldehyde, glyceraldehyde 3-phosphate, NADPH, NADP, and glycerol were not utilized. The spinach leaf enzyme appears to be in the cytoplasm and probably functions for the production of glycerol 3-phosphate from dihydroxyacetone phosphate.  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