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.
Some physical and chemical properties of the monomeric NADP+-dependent aldehyde reductase (previously called TPN-L-hexonate dehydrogenase or D-glucuronate reductase) from pig kidney have been examined. The amino acid composition has been determined. Four of the five thiol groups react with p-mercuribenzoate at pH 7, with no resulting loss of catalytic activity. High concentrations of p-mercuribenzoate cause complete enzyme inhibition, which can be partly reversed by addition of aldehyde reductase is low (9%, estimated from the ellipticity at 208 nm), and 70 to 80% of the tyrosine and tryptophan residues aare buried within the molecule. One molecule of NADPH binds to the enzyme (Kp equal 25 muM), causing a blue shift and enhancement of the coenzyme fluorescence, and suggesting that the environment of the active site is hydrophobic. In the reduction of D-glyceraldehyde, catalyzed by aldehyde reductase, the pro-4R "A" hydrogen of NADPH attacks the re face of the carbonyl group. This stereospecificity is the same as in the reductions of D-glyceraldehyde and acetaldehyde effected by rabbit muscle dehydrogenase and liver alcohol dehydrogenase, respectively.  相似文献   

2.
Biliverdin reductase (molecular form 1, EC 1.3.1.24, bilirubin:NAD(P)+ oxidoreductase) carries three thiol residues. Only one of them could be alkylated when a ratio N-ethylmaleimide (NEM)/mol enzyme's SH = 90 was used. The alkylation of this thiol group inhibited the conversion of molecular form 1 to its dimer, molecular form 3; however, it did not inhibit the enzymatic activity. At a ratio of NEM/enzyme's SH = 300, two thiol residues were alkylated and the activity of the enzyme was totally inhibited. The third thiol group could not be alkylated either by NEM or by iodoacetamide. Biliverdin as well as the co-substrate NADPH protected the thiol residue essential for the enzymatic activity from alkylation. Spectroscopic evidence was obtained that this thiol group binds covalently to the C-10 of biliverdin to form a rubinoid adduct. The presence of a lysine residue, which is also essential for the enzymatic activity, could be inferred from the fact that by reduction of the Schiff base formed by the enzyme with pyridoxal phosphate the catalytic activity was irreversibly abolished. The location of a lysine residue in the vicinity of the thiol group involved in the catalytic activity was evident when the enzyme was treated with o-phthalaldehyde. The inactivation of the enzymatic activity was coincident with the formation of the fluorescent isoindole derivative which originates when the thiol and epsilon-NH2 groups are located about 3 A apart. The presence of a positively charged ammonium ion in the vicinity of the NADPH binding site was inferred from the shifts in the UVmax of NADPH from 340 nm to 327 nm and of 3-acetyl NADPH from 360 nm to 348 nm when the pyridine nucleotides bind to the reductase. The involvement of arginine residues in the enzymatic activity was established by inhibition of the latter after reaction with butanedione. This inhibition was totally protected by NADPH but not by biliverdin. The similarity of the structural features of biliverdin reductase with those of several dehydrogenases is discussed.  相似文献   

3.
Aldose reductase (EC 1.1.1.21) and aldehyde reductase II (L-hexonate dehydrogenase, EC 1.1.1.2) have been purified to homogeneity from human erythrocytes by using ion-exchange chromatography, chromatofocusing, affinity chromatography, and Sephadex gel filtration. Both enzymes are monomeric, Mr 32,500, by the criteria of the Sephadex gel filtration and polyacrylamide slab gel electrophoresis under denaturing conditions. The isoelectric pH's for aldose reductase and aldehyde reductase II were determined to be 5.47 and 5.06, respectively. Substrate specificity studies showed that aldose reductase, besides catalyzing the reduction of various aldehydes such as propionaldehyde, pyridine-3-aldehyde and glyceraldehyde, utilizes aldo-sugars such as glucose and galactose. Aldehyde reductase II, however, did not use aldo-sugars as substrate. Aldose reductase activity is expressed with either NADH or NADPH as cofactors, whereas aldehyde reductase II can utilize only NADPH. The pH optima for aldose reductase and aldehyde reductase II are 6.2 and 7.0, respectively. Both enzymes are susceptible to the inhibition by p-hydroxymercuribenzoate and N-ethylmaleimide. They are also inhibited to varying degrees by aldose reductase inhibitors such as sorbinil, alrestatin, quercetrin, tetramethylene glutaric acid, and sodium phenobarbital. The presence of 0.4 M lithium sulfate in the assay mixture is essential for the full expression of aldose reductase activity whereas it completely inhibits aldehyde reductase II. Amino acid compositions and immunological studies further show that erythrocyte aldose reductase is similar to human and bovine lens aldose reductase, and that aldehyde reductase II is similar to human liver and brain aldehyde reductase II.  相似文献   

4.
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   

5.
Some properties of aldose reductase isolated from various sources and possible ways of regulation of the enzyme catalytic activity are reviewed. Mammalian aldose reductases are monomeric enzymes with M(r) of 30-40 kDa and a broad substrate specificity towards aldoses. The physiological role of this enzyme consists, apparently, in providing an additional pathway for utilization of glucose and removing toxic compounds carrying an aldehyde group from the cell. Aldose reductase is thought to play a key role in various hyperglycemic states, including diabetic cataract. The kinetics of the aldose reductase reaction is hyperbolic with NADPH and nonhyperbolic with glucose. The rate of the enzyme-catalyzed reaction is determined by the effector binding in the active of inhibitory center of the enzyme. Incubation with substrates leads to the activation of the enzyme which is accompanied by a decrease of the effector binding in the enzyme inhibitory center with a sharp decrease in the sensitivity of the activated enzyme to NADPH concentration changes in the presence of glucose excess. A mechanism underlying the catalytic effect of both native and activated forms of the enzyme is proposed.  相似文献   

6.
Dimeric and monomeric proteins containing dihydrodiol dehydrogenase and aldehyde reductase activities were purified from pig lens. The dimeric enzyme of Mr 65,000 specifically oxidized the trans-dihydrodiols of naphthalene and benzene with NADP+ as a strict cofactor, and reduced alpha-diketones, aromatic aldehydes and glyceraldehyde with NADPH as a cofactor. The monomeric enzyme of Mr 35,000, although identical with aldose reductase, oxidized the trans-dihydrodiol of naphthalene at a pH optimum of 7.6. These results suggest that the two enzymes are involved in the pathogenesis of naphthalene cataract.  相似文献   

7.
A cytosolic aldo-keto reductase was purified from Saccharomyces cerevisiae ATCC 26602 to homogeneity by affinity chromatography, chromatofocusing, and hydroxylapatite chromatography. The relative molecular weights of the aldo-keto reductase as determined by sodium dodecyl sulfate-polyacrylamide gel electrophoresis and size exclusion chromatography were 36,800 and 35,000, respectively, indicating that the enzyme is monomeric. Amino acid composition and N-terminal sequence analysis revealed that the enzyme is closely related to the aldose reductases of xylose-fermenting yeasts and mammalian tissues. The enzyme was apparently immunologically unrelated to the aldose reductases of other xylose-fermenting yeasts. The aldo-keto reductase is NADPH specific and catalyzes the reduction of a variety of aldehydes. The best substrate for the enzyme is the aromatic aldehyde p-nitrobenzaldehyde (Km = 46 microM; kcat/Km = 52,100 s-1 M-1), whereas among the aldoses, DL-glyceraldehyde was the preferred substrate (Km = 1.44 mM; kcat/Km = 1,790 s-1 M-1). The enzyme failed to catalyze the reduction of menadione and p-benzoquinone, substrates for carbonyl reductase. The enzyme was inhibited only slightly by 2 mM sodium valproate and was activated by pyridoxal 5'-phosphate. The optimum pH of the enzyme is 5. These data indicate that the S. cerevisiae aldo-keto reductase is a monomeric NADPH-specific reductase with strong similarities to the aldose reductases.  相似文献   

8.
Kinetic and physicochemical properties of hamster liver diacetyl reductase have been examined. The results of kinetic studies on the reduction of diacetyl and NADPH to acetoin and NADP+ suggest that the reaction follows an Ordered Bi Bi mechanism in which NADPH binds first before diacetyl. The enzyme is a tetrameric glycoprotein of single subunits of a molecular weight of 23,500 with a sedimentation coefficient of 6.0S. The enzyme does not contain Zn, Cu, or Fe. The amino acid composition revealed an unusually low proportion of proline residues (0.9%). p-Chloromercuriphenylsulfonate and phenylglyoxal inactivated the enzyme, but the presence of NADPH prevented the loss of activity due to thiol and arginine modification. The enzyme transferred the pro 4S hydrogen atom of NADPH to the substrate and the binding of the enzyme to NADPH resulted in a red shift of the ultraviolet absorption spectrum of the cofactor.  相似文献   

9.
An improved procedure for purifying aldehyde reductase is described. Utilization of Blue Dextran--Sepharose 4B and elimination of hydroxyapatite chromatography greatly improves the yield and ease of purification. Starting with 340 g of kidney tissue (two pig kidneys) approx. 50 mg of purified reductase may be routinely and reproducibly obtained. The purified reductase was used to establish the kinetic reaction mechanism of the enzyme. Initial-velocity analysis and product-inhibition data revealed that pig kidney aldehyde reductase follows an Ordered Bi Bi reaction mechanism in which NADPH binds first before D-glyceraldehyde. The limiting Michaelis constants for D-glyceraldehyde and NADPH were 4.8 +/- 0.7 mM and 9.1 +/- 2.1 micrometer respectively. The mechanism is similar to that of another monomeric oxidoreductase, octopine dehydrogenase, towards which aldehyde reductase exhibits several similarities, but differs from that of other aldehyde reductases. Phenobarbital is a potent inhibitor of aldehyde reductase, inhibiting both substrate and cofactor non-competitively (Ki = 80.4 +/- 10.5 micrometer and 66.9 +/- 1.6 micrometer respectively). Barbiturate inhibition seems to be a common property of NADPH-dependent aldehyde reductases.  相似文献   

10.
Glutathione reductase (NAD(P)H:GSSG oxidoreductase EC 1.6.4.2.) was purified 1160-fold to homogeneity from the nonsulfurous purple bacteria Rhodospirillum rubrum (wild type). Specific activity of the pure preparation was 102 U/mg. The enzyme displayed a typical flavoprotein absorption spectrum with maxima at 274,365, and 459 nm and an absorbance ratio A280/A459 of 7.6. The amino acid analysis revealed an unusually high content of glycine and arginine residues. Titration of the enzyme with 5,5'-dithiobis(2-nitrobenzoic acid) showed a total of two free thiol groups per subunit, one of which is made accessible only under denaturing conditions. An isoelectric point of 5.2 was found for the native enzyme. Km values, determined at pH 7.5, were 6.1 and 90 microM for NADPH and GSSG, respectively. NADH was about 2% as active as NADPH as an electron donor. The enzyme's second choice in disulfide substrate was the mixed disulfide of coenzyme A and glutathione, for which the specific activity and Km values were 5.1 U/mg and 3.4 mM, respectively. A native molecular weight of 118,000 was found, while denaturing electrophoresis gave a value of 54,400 per subunit, thus suggesting that R. rubrum glutathione reductase exists as a dimeric protein. Other physicochemical constants of the enzyme, such as Stokes radius (4.2 nm) and sedimentation coefficient (5.71 S), were also consistent with a particle of 110,000.  相似文献   

11.
Amidination of human liver aldehyde reductase (alcohol:NADP+ oxidoreductase, EC 1.1.1.2) with monofunctional n-alkane methylimidates increased the enzymic activity by 10--30%, whereas analogous bifunctional imidoesters caused a loss of activity of about 80%. Both effects were prevented in the presence of the coenzyme NADPH or NADP+, but not of the substrate 4-nitrobenzaldehyde. Amidination increased the apparent Michaelis constant of both the coenzyme (up to 20-fold) and the substrate (about 5-fold). Bifunctional imidoesters with at least 4 carbon atoms between the functional groups (approx. 0.7 nm) crosslinked the enzyme intramolecularly. This reaction was retarded in the presence of the coenzyme, whereas 4-nitrobenzaldehyde had no effect. The results suggest the presence of reactive amino groups at the coenzyme binding site of aldehyde reductase.  相似文献   

12.
Purification and properties of low-Km aldehyde reductase from ox brain   总被引:1,自引:0,他引:1  
A low-Km aldehyde reductase (alcohol:NADP+ oxidoreductase, EC 1.1.1.2), which may be identical with aldose reductase (alditol:NADP+ 1-oxidoreductase, EC 1.1.1.21), has been purified from ox brain to homogeneity. It was shown to be a monomer with Mr values of 31 000 and 35 100 being obtained by gel filtration and polyacrylamide gel electrophoresis in the presence of sodium dodecyl sulphate, respectively. The enzyme catalyses the NADPH-dependent reduction of a number of aromatic and sugar aldehydes. The activity of the enzyme with 133 microM NADH was about one-third of that with 120 microM NADPH. Activity with both these coenzymes was optimum at pH 6.2 and was inhibited by increasing the ionic strength with KCl, NaCl or NaNO3. In contrast, the activity was stimulated by sodium phosphate. The activity with NADH as the coenzyme was more sensitive to stimulation by phosphate and to inhibition by increasing ionic strength than that determined with NADPH.  相似文献   

13.
3-Hydroxy-3-methylglutaryl-CoA reductase (NADPH) was solubilized with polyoxyethylene ether (Brij) W-1 from a heavy-membrane fraction, sedimented at 16000 X g from a cell-free homogenate of four-day-old, dark-grown radish seedlings (Raphanus sativus L.). Approximately 350-fold purification of the solubilized enzyme activity was achieved by (NH4)2SO4 precipitation followed by column chromatography on DEAE-Sephadex A-50, blue-dextran-agarose and HMG-CoA-hexane-agarose. The presence of detergent, which was required at all times to maintain activity, did not interfere with the chromatographic procedures used. Sucrose density centrifugation suggested an apparent molecular mass of 180 kDa with subunits of 45 kDa (polyacrylamide gel electrophoresis in the presence of sodium dodecylsulphate). The enzyme was stable at 67.5 degrees C for 30 min in the presence of glycerol, dithioerythritol and detergent. Studies of enzyme stability and activation indicate that the enzyme is a hydrophobic protein with free thiol groups that are essential for full activity. The activation energy was estimated to be 92 kJ (Arrhenius plot). Antibodies raised against rat liver and yeast hydroxymethylglutaryl-CoA (HMG-CoA) reductase failed to bind or inactivate the radish enzyme. When both HMG-CoA and NADPH concentrations were varied, intersecting patterns were obtained with double-reciprocal plots. The apparent Km values determined in this way are 1.5 microM [(S)-HMG-CoA], and 27 microM (NADPH). Concentrations of NADPH greater than 150 microM caused substrate inhibition at low HMG-CoA concentrations resulting in deviations from linearity in secondary plots. Analysis of these data and the product inhibition pattern suggest a sequential mechanism for the reduction of HMG-CoA to mevalonic acid with HMG-CoA being the first substrate binding to the enzyme, followed by NADPH.  相似文献   

14.
The reduced nicotinamide adenine dinucleotide phosphate (NADPH)-nitrate oxidoreductase (EC 1.6.6.2) from Aspergillus nidulans was purified over 200-fold by use of salt fractionation, gel filtration, and ion-exchange chromatography. The purified enzyme was specific for NADPH and catalyzed reduction of nitrate, cytochrome c from isolated mitochondria of Aspergillus, and mammalian cytochrome c. An S(0.725) (20, w) of 7.8 was derived with sucrose density gradient centrifugation, and a Stokes radius of 6.4 nm was derived by gel filtration on Sephadex G-200. From these values, a molecular weight of 197,000 was computed, assuming v = 0.725 cm(3)/g. The spectral properties of the purified enzyme suggested a flavine component was present but revealed no pattern indicative of a hemoprotein. A cytochrome c, similar to the cytochrome c from isolated mitochondria, was found unassociated with the nitrate reductase after ion-exchange chromatography. No NADPH-nitrate reductase activity was detected in isolated mitochondria. Spectrally discernable reduction of the flavine component of the enzyme at 450 nm was noted after reaction with NADPH. This reduction was inhibited by p-chloromercuribenzoate but not by KCN. The addition of nitrate to NADPH reduced enzyme caused a reoxidation of the flavine component via a reaction which was inhibited by KCN but not by p-chloromercuribenzoate. The half-life of the purified enzyme at 37 C was 20 min for NADPH-nitrate reductase and 35 min for NADPH-cytochrome c reductase.  相似文献   

15.
A soluble cytochrome P450 from the yeast Trichosporon cutaneum was purified to homogeneity, using ammonium sulfate fractionation followed by fast protein liquid chromatography (FPLC) with DEAE-cellulose and phenyl-Sepharose columns. This procedure resulted in a 45-fold increase in specific activity with an activity yield of 6.8%. One- and two-dimensional sodium dodecyl sulfate-polyacrylamide gel electrophoresis showed that the purified enzyme was homogeneous and had a molecular mass of 45 kDa. The purified enzyme contained a heme group and had a characteristic absorption peak at 448 nm in the reduced carbon monoxide difference spectrum. This enzyme was a monomeric protein and catalyzed the conversion of salicylic acid to catechol in the presence of NADH or NADPH. The N-terminal amino acid sequence indicated that the Trichosporon cutaneum cytochrome P450 did not show homology to most eukaryotic cytochromes P450, but had a high degree of homology to one cytochrome P450, the nitric oxide reductase, of Fusarium oxysporum.  相似文献   

16.
Characterization of aldose reductase and aldehyde reductase from rat testis   总被引:4,自引:0,他引:4  
Aldose reductase (alditol:NAD(P)+ 1-oxidoreductase, EC 1.1.1.21) and aldehyde reductase (alcohol:NADP+ oxidoreductase, EC 1.1.1.2) were purified to a homogeneity from rat testis. The molecular weights of aldose reductase and aldehyde reductase were estimated to be 38,000 and 41,000 by SDS-polyacrylamide gel electrophoresis, and the pI values of these enzymes were found to be 5.3 and 6.1 by chromatofocusing, respectively. Aldose reductase had activity for aldo-sugars such as xylose, glucose and galactose, whereas aldehyde reductase was virtually inactive for these aldo-sugars. The Km values of aldose reductase for aldo-sugars were relatively high. When a correction was made for the fraction of aldo-sugar present as the aldehyde form, which is the real substrate of the enzyme, the Km values were much lower. Aldose reductase utilized both NADPH and NADH as coenzyme, whereas aldehyde reductase utilized only NADPH. Aldose reductase was activated significantly by sulfate ion, while aldehyde reductase was little affected. Both enzymes were inhibited strongly by the known aldose reductase inhibitors. However, aldehyde reductase was in general less susceptible to these inhibitors when compared to aldose reductase. Both aldose reductase and aldehyde reductase treated with pyridoxal 5-phosphate have lost the susceptibility to aldose reductase inhibitor, suggesting that in these two enzymes aldose reductase inhibitor interacts with a lysine residue.  相似文献   

17.
Aldehyde reductase (alcohol:NADP+ oxidoreductase, EC 1.1.1.2), aldose reductase (alditol:NAD(P)+ 1-oxidoreductase, EC 1.1.1.21) and carbonyl reductase (secondary-alcohol:NADP+ oxidoreductase, EC 1.1.1.184) constitute the enzyme family of the aldo-keto reductases, a classification based on similar physicochemical properties and substrate specificities. The present study was undertaken in order to obtain information about the structural relationships between the three enzymes. Treatment of human aldehyde and carbonyl reductase with phenylglyoxal and 2,3-butanedione caused a complete and irreversible loss of enzyme activity, the rate of loss being proportional to the concentration of the dicarbonyl reagents. The inactivation of aldehyde reductase followed pseudo-first-order kinetics, whereas carbonyl reductase showed a more complex behavior, consistent with protein modification cooperativity. NADP+ partially prevented the loss of activity of both enzymes, and an even better protection of aldehyde reductase was afforded by the combination of coenzyme and substrate. Aldose reductase was partially inactivated by phenylglyoxal, but insensitive to 2,3-butanedione. The degree of inactivation with respect to the phenylglyoxal concentration showed saturation behavior. NADP+ partially protected the enzyme at low phenylglyoxal concentrations (0.5 mM), but showed no effect at high concentrations (5 mM). These findings suggest the presence of an essential arginine residue in the substrate-binding domain of aldehyde reductase and the coenzyme-binding site of carbonyl reductase. The effect of phenylglyoxal on aldose reductase may be explained by the modification of a reactive thiol or lysine rather than an arginine residue.  相似文献   

18.
Three kinds of NADPH-dependent aldehyde reducing enzymes were present in the dog kidney. Aldose reductase was located in the inner medulla region and aldehyde reductase in all regions of the renal cortex, outer medulla and inner medulla. In addition, a new reductase designated tentatively as high-Km aldose reductase, which was converted into an aldose reductase-like enzyme, was present in the inner medulla region of the kidney. Aldose reductase, aldehyde reductase and high-Km aldose reductase were purified to homogeneity from each region of the dog kidney. The molecular weight of aldose reductase was estimated to be 38,500 by SDS-polyacrylamide gel electrophoresis and the isoelectric point was found to be 5.7 by chromatofocusing. Aldose reductase had activity for aldo-sugars such as D-xylose, D-glucose and D-galactose as substrates and utilized both NADPH and NADH as coenzymes. Sulfate ions resulted in over 2-fold activation of aldose reductase. All aldehyde reductases from the three regions had the same properties. The molecular weights and isoelectric points of aldehyde reductases were 40,000 and 6.1, respectively. The aldehyde reductases were inactive for D-hexose, utilized only NADPH as coenzyme and were not affected by sulfate ions. High-Km aldose reductase had a molecular weight of 38,500 and an isoelectric point of 5.4. It had activity for aldo-sugars, but showed much higher Km and lower kcat/Km values than aldose reductase. Sulfate ions inhibited high-Km aldose reductase. It was converted into an aldose reductase-like enzyme by incubation in phosphate buffer at pH 7.0. The three kinds of enzymes were strongly inhibited by the known aldose reductase inhibitors. However, aldehyde reductase and high-Km aldose reductase were, in general, less susceptible than aldose reductase.  相似文献   

19.
Monodehydroascorbate reductase (EC 1.6.5.4) was purified from cucumber fruit to a homogeneous state as judged by polyacrylamide gel electrophoresis. The cucumber monodehydroascorbate reductase was a monomer with a molecular weight of 47,000. It contained 1 mol of FAD/mol of enzyme which was reduced by NAD(P)H and reoxidized by monodehydroascorbate. The enzyme had an exposed thiol group whose blockage with thiol reagents inhibited the electron transfer from NAD(P)H to the enzyme FAD. Both NADH and NADPH served as electron donors with Km values of 4.6 and 23 microM, respectively, and Vmax of 200 mol of NADH and 150 mol of NADPH oxidized mol of enzyme-1 s-1. The Km for monodehydroascorbate was 1.4 microM. The amino acid composition of the enzyme is presented. In addition to monodehydroascorbate, the enzyme catalyzed the reduction of ferricyanide and 2,6-dichloroindophenol but showed little reactivity with calf liver cytochrome b5 and horse heart cytochrome c. The kinetic data suggested a ping-pong mechanism for the monodehydroascorbate reductase-catalyzed reaction. Cucumber monodehydroascorbate reductase occurs in soluble form and can be distinguished from NADPH dehydrogenase, NADH dehydrogenase, DT diaphorase, microsome-bound NADH-cytochrome b5 reductase, and NADPH-cytochrome c reductase by its molecular weight, amino acid composition, and specificity of electron acceptors and donors.  相似文献   

20.
The kinetic theory of the substrate reaction during modification of enzyme activity previously described by Tsou [Tsou (1988),Adv. Enzymol. Relat. Areas Mol. Biol. 61, 381–436] has been applied to a study of the kinetics of the course of inactivation of the mitochondrial succinate-ubiquinone reductase by 5,5′-dithiobis-(2-nitro-benzoic acid) (DTNB). The results show that the inactivation of this enzyme by DTNB is a conformation-change-type inhibition which involves a conformational change of the enzyme before inactivation. The microscopic rate constants were determined for the reaction of the inactivator with the enzyme. The presence of the substrate provides marked protection of this enzyme against inactivation by DTNB. The modification reaction of the enzyme using DTNB was shown to follow a triphasic course by following the absorption at 412 nm. Among these reactive thiol groups, the fast-reaction thiol group is essential for the enzyme activity. The results suggest that the essential thiol group is situated at the succinate-binding site of the mitochondrial succinate-ubiquinone reductase.  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