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 109 毫秒
1.
Chromatography of bovine kidney aldose reductase using Matrex Orange A affinity gel results in the separation of the unactivated and activated enzyme forms. The former washes through the column, while the latter is eluted with an NADPH step-gradient. The separated enzyme forms display Vmax and Km glycolaldehyde values, and relative sensitivities to inhibition by the aldose reductase inhibitor AL-1576 (spiro[2,7-difluorofluorene-9,4'-imidazolidine]-2',5'- dione), that are similar to those reported previously for the individual forms. However, because Vmax is 17-fold lower for the unactivated enzyme, the purification of aldose reductase via NADP(H) elution from a dye-ligand affinity matrix can result in the selective purification of only the activated enzyme form. These results have direct implications for the study of potential aldose reductase inhibitors, and may explain why linear double-reciprocal plots are commonly observed for enzyme prepared in this manner, while nonlinear plots are seen in other cases.  相似文献   

2.
1. Aldose reductase and aldehyde reductase were purified to homogeneity from human testis. 2. The molecular weight of aldose reductase and aldehyde reductase were estimated to be 36,000 and 38,000 by SDS-PAGE, and the pI values of these enzymes were found to be 5.9 and 5.1 by chromatofocusing, respectively. 3. Aldose reductase had activity for aldo-sugars, whereas aldehyde reductase was virtually inactive for aldo-sugars. The Km values of aldose reductase for D-glucose, D-galactose and D-xylose were 57, 49 and 6.2 mM, respectively. Aldose reductase utilized both NADPH and NADH as coenzymes, whereas aldehyde reductase only NADPH. 4. Sulfate ion caused 3-fold activation of aldose reductase, but little for that of aldehyde reductase. 5. Sodium valproate inhibited significantly aldehyde reductase, but not aldose reductase. Aldose reductase was inhibited strongly by aldose reductase inhibitors being in clinical trials at concentrations of the order of 10(-7)-10(-9) M. Aldehyde reductase was also inhibited by these inhibitors, but its susceptibility was less than aldose reductase. 6. Reaction of aldose reductase with pyridoxal 5'-phosphate (PLP) resulted ca 2.5-fold activation, but aldehyde reductase did not cause the activation. PLP-treated aldose reductase has lost the susceptibility to aldose reductase inhibitor.  相似文献   

3.
Xylose reductase has been purified to apparent homogeneity from cell extracts of the fungus Cryptococcus flavus grown on D-xylose as carbon source. The enzyme, the first of its kind from the phylum Basidiomycota, is a functional dimer composed of identical subunits of 35.3 kDa mass and requires NADP(H) for activity. Steady-state kinetic parameters for the reaction, D-xylose + NADPH + H(+)<--> xylitol + NADP(+), have been obtained at pH 7.0 and 25 degrees C. The catalytic efficiency for reduction of D-xylose is 150 times that for oxidation of xylitol. This and the 3-fold tighter binding of NADPH than NADP(+) indicate that the enzyme is primed for unidirectional metabolic function in microbial physiology. Kinetic analysis of enzymic reduction of aldehyde substrates differing in hydrophobic and hydrogen bonding capabilities with binary enzyme-NADPH complex has been used to characterize the substrate-binding pocket of xylose reductase. Total transition state stabilization energy derived from bonding with non-reacting sugar hydroxyls is approximately 15 kJ/mol, with a major contribution of 5-8 kJ/mol made by interactions with the C-2(R) hydroxy group. The aldehyde binding site is approximately 1.2 times more hydrophobic than n-octanol and can accommodate linear alkyl chains of 相似文献   

4.
A pyrazole-sensitive carbonyl reductase from pig lung was purified to homogeneity by electrophoretic criteria. Chemical cross-linking study suggested that the native enzyme is a tetramer with a Mr of 103,000, consisting of apparent identical subunits of Mr 24,000. The enzyme reduced aliphatic and aromatic carbonyl compounds with NADPH as a preferable cofactor to NADH and catalyzed the oxidation of secondary alcohols and the aldehyde dismutation in the presence of NAD(P)+. Immunohistochemical study with the antibodies against the enzyme revealed that the enzyme was localized in the ciliated cells, nonciliated bronchiolar cells, Type II alveolar pneumocytes, and the epithelial cells of the ducts of the bronchial glands in the pig lung. In addition to the properties and distribution, the pig lung enzyme was immunochemically similar to the pulmonary enzymes in the guinea pig and mouse. However, the pig enzyme showed the following unusual features. (1) The enzyme exhibited an equatorial specificity in the reduction of 3-ketosteroids; the 4-pro-S hydrogen of NADPH was transferred to the carbonyl carbon atom of 5 alpha- and 5 beta-androstanes, and the respective reduced products were identified as 3 beta- and 3 alpha-hydroxysteroids. (2) Although the NADPH-linked reduction of carbonyl compounds apparently obeyed the Michaelis-Menten kinetics at pH 6.0, the double-reciprocal plots of the velocity vs concentrations of the carbonyl substrates were convex at pH higher than 6.5. The Hill coefficients and [S]0.5 values for the substrates decreased as the pH for reaction increased. The results suggest that the pig enzyme exhibits negative cooperativity with respect to the carbonyl substrates and that the hydrogen ion acts as an allosteric effector abolishing the negative interaction.  相似文献   

5.
Human aldose reductase and aldehyde reductase are members of the aldo-keto reductase superfamily that share three domains of homology and a nonhomologous COOH-terminal region. The two enzymes catalyze the NADPH-dependent reduction of a wide variety of carbonyl compounds. To probe the function of the domains and investigate the basis for substrate specificity, we interchanged cDNA fragments encoding the NH2-terminal domains of aldose and aldehyde reductase. A chimeric enzyme (CH1, 317 residues) was constructed in which the first 71 residues of aldose reductase were replaced with first 73 residues of aldehyde reductase. Catalytic effectiveness (kcat/Km) of CH1 for the reduction of various substrates remained virtually identical to wild-type aldose reductase, changing a maximal 4-fold. Deletion of the 13-residue COOH-terminal end of aldose reductase, yielded a mutant enzyme (AR delta 303-315) with markedly decreased catalytic effectiveness for uncharged substrates ranging from 80- to more than 600-fold (average 300-fold). The KmNADPH of CH1 and AR delta 303-315 were nearly identical to that of the wild-type enzyme indicating that cofactor binding is unaffected. The truncated AR delta 303-315 displayed a NADPH/D isotope effect in kcat and an increased D(kcat/Km) value for DL-glyceraldehyde, suggesting that hydride transfer has become partially rate-limiting for the overall reaction. We conclude that the COOH-terminal domain of aldose reductase is crucial to the proper orientation of substrates in the active site.  相似文献   

6.
Aldose reductase (ALR2) has been purified to homogeneity from human psoas muscle. From sodium dodecyl sulfate-polyacrylamide electrophoresis the enzyme is monomeric and has a molecular weight of 37,000. ALR2 catalyzes the primarily NADPH-dependent reduction of a wide variety of aldehydes, although the enzyme can also utilize NADH. The best substrates for ALR2 are aromatic aldehydes (e.g. pyridine-3-aldehyde; Km = 9 microM; kcat/Km = 150,000 s-1 M-1), while among aldoses DL-glyceraldehyde is the preferred substrate (Km = 72 microM; kcat/Km = 17,250). Low (100 microM) concentrations of CaCl2 and CaSO4 cause a marked inhibition (90%) of ALR2 as do higher concentrations (0.2 M) of MgCl2. (NH4)2SO4 caused a 2-fold activation of ALR2. The enzyme is also inhibited by quercetin and the commercially developed aldose reductase inhibitors alrestatin and sorbinil. ALR2 is inhibited only very slightly by sodium valproate and barbiturates. ALR2 cross-reacts immunologically with human brain and human placental aldose reductase and with ALR2 from monkey tissue. There is no precipitin cross-reaction of ALR2 with aldose reductases from other species nor with human aldehyde reductase 1 (ALR1) or with ALR1 from other species. The data show that human muscle is a new and relatively rich source of a monomeric NADPH/NADH reductase which is clearly identifiable as aldose reductase.  相似文献   

7.
Polycyclic aromatic hydrocarbons (PAHs) are metabolized to trans-dihydrodiol proximate carcinogens by human epoxide hydrolase (EH) and CYP1A1. Human dihydrodiol dehydrogenase isoforms (AKR1C1-AKR1C4), members of the aldo-keto reductase (AKR) superfamily, activate trans-dihydrodiols by converting them to reactive and redox-active o-quinones. We now show that the constitutively and widely expressed human AKR, aldehyde reductase (AKR1A1), will oxidize potent proximate carcinogen trans-dihydrodiols to their corresponding o-quinones. cDNA encoding AKR1A1 was isolated from HepG2 cells, overexpressed in Escherichia coli, purified to homogeneity, and characterized. AKR1A1 oxidized the potent proximate carcinogen (+/-)-trans-7,8-dihydroxy-7,8-dihydrobenzo[a]pyrene with a higher utilization ratio (V(max)/K(m)) than any other human AKR. AKR1A1 also displayed a high V(max)/K(m) for the oxidation of 5-methylchrysene-7,8-diol, benz[a]anthracene-3,4-diol, 7-methylbenz[a]anthracene-3,4-diol, and 7,12-dimethylbenz[a]anthracene-3,4-diol. AKR1A1 displayed rigid regioselectivity by preferentially oxidizing non-K-region trans-dihydrodiols. The enzyme was stereoselective and oxidized 50% of each racemic PAH trans-dihydrodiol tested. The absolute stereochemistries of the reactions were assigned by circular dichroism spectrometry. AKR1A1 preferentially oxidized the metabolically relevant (-)-benzo[a]pyrene-7(R),8(R)-dihydrodiol. AKR1A1 also preferred (-)-benz[a]anthracene-3(R),4(R)-dihydrodiol, (+)-7-methylbenz[a]anthracene-3(S),4(S)-dihydrodiol, and (-)-7,12-dimethylbenz[a]anthracene-3(R),4(R)-dihydrodiol. The product of the AKR1A1-catalyzed oxidation of (+/-)-trans-7,8-dihydroxy-7,8-dihydrobenzo[a]pyrene was trapped with 2-mercaptoethanol and characterized as a thioether conjugate of benzo[a]pyrene-7,8-dione by LC/MS. Multiple human tissue expression array analysis showed coexpression of AKR1A1, CYP1A1, and EH, indicating that trans-dihydrodiol substrates are formed in the same tissues in which AKR1A1 is expressed. The ability of this general metabolic enzyme to divert trans-dihydrodiols to o-quinones suggests that this pathway of PAH activation may be widespread in human tissues.  相似文献   

8.
Steady-state kinetics of the action of mung bean phosphoglycerate kinase have been investigated using 3-phosphoglycerate and ATP as substrates in the presence of Mg2+ ions. Keeping a constant and high Mg2+ concentration and varying the concentration of one of the substrates (ATP or 3-phosphoglycerates) at several fixed concentrations of the other substrate (3-phosphoglycerate or ATP), the Km values of Mg.ATP2- and 3-phosphoglycerate were found to be 0.42 and 0.60 mM, respectively. These values are independent of the concentration of the other substrate. A limiting value of Vmax, where the enzyme is saturated with both the substrates, was found to be 39.4 mumoles product formed per min per mg enzyme protein. This corresponds to a turnover number equal to 31.5 sec-1 (for molecular weight of the enzyme equal to 48,000). If [Mg2+] and [ATP4-] are held equal and varied together at several fixed concentrations of 3-phosphoglycerate, deviations from Michaelis-Menten kinetics (non-linear Lineweaver-Burk plots) are observed at lower values of ATP4- and Mg2+ (less than 0.1 mM), giving rise to apparent sigmoidicity in the rate versus [ATP4-] plots. It has been suggested that the real substrate for this enzyme is the Mg.ATP2- complex (and not free ATP4-). The complex dissociates at lower values of [Mg2+] and [ATP4-]. The resulting disproportionate decrease in the concentration of the complex brings about a steeper fall in the rate of reaction than is required by the Michaelis-Menten equation, giving rise to an apparent sigmoidicity.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

9.
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   

10.
NADH:ubiquinone oxidoreductase (NDH-1 or complex I) from Escherichia coli was purified using a combination of anion exchange chromatography and centrifugation in sucrose density gradient. The dependence of enzyme activity on detergent and phospholipids was studied. Artificial hexaammineruthenium reductase activity was not affected by dodecyl maltoside (DDM) and asolectin. Ubiquinone reductase activity had a bell-shape dependence on DDM concentration; 7-10-fold activation could be achieved. Treatment with asolectin subsequently yields additional 2-fold activation with a corresponding increase in the apparent V(max) and without significant changes in apparent K(m). Comparative EPR studies of complex I reduced with NADH, "as prepared" and "activated by asolectin" showed an increase in the signals derived mainly from two [4Fe-4S] clusters in the activated enzyme. One of these signals could be simulated with an axial spectrum with g values of g(xyz)= 1.895, 1.904, 2.05, which corresponds to the parameters reported for the N2 cluster. This data indicates conformational rearrangements of catalytic importance in complex I upon binding of phospholipids.  相似文献   

11.
During the purification of pig kidney aldehyde reductase by an established procedure [Flynn, Cromlish & Davidson (1982) Methods Enzymol. 89, 501-506] a second enzyme with aldehyde reductase activity may be purified. When the procedure was performed in the presence of 5 mM-EDTA, only traces of the second reductase, pig kidney aldehyde reductase (minor form), were present. By the criterion of sodium dodecyl sulphate/polyacrylamide-gel electrophoresis, pig kidney aldehyde reductase (minor form) had Mr 35 000, in comparison with Mr 40 200 found for pig kidney aldehyde reductase. Amino acid analysis of both enzymes and tryptic-peptide-map comparisons indicated differences in primary structure. The N-terminus of pig kidney aldehyde reductase (minor form) had the sequence Lys-Val-Leu, in contrast with the blocked (acetylated) N-terminus of pig kidney aldehyde reductase. The C-terminal sequence of both enzymes was the same. Both reductases were immunologically identical by double immunodiffusion and rocket immunoelectrophoresis. Pig kidney aldehyde reductase (minor form) had 50% of the specific activity of pig kidney aldehyde reductase when tested with a variety of aldehyde substrates. Michaelis constants of both enzymes for these substrates and for NADPH were similar, but values for kcat. and kcat./Km indicated that catalytically pig kidney aldehyde reductase was the more efficient enzyme. Typical aldehyde reductase inhibitors, such as phenobarbital and sodium valproate, had the same effect on both enzymes. It was concluded that pig kidney aldehyde reductase (minor form) is an enzymically active cleavage product of pig kidney aldehyde reductase which is formed when the latter is purified in the absence of the metalloproteinase inhibitor EDTA.  相似文献   

12.
A pathogen elicitor-inducible soluble acyltransferase (tyramine hydroxycinnamoyltransferase [THT], EC 2.3.1), which catalyzes the transfer of hydroxycinnamic acids from hydroxycinnamoyl-coenzyme A (CoA) esters to tyramine in the formation of N-hydroxycinnamoyltyramine, was partially purified with a 380-fold enrichment and a 6% recovery from cell-suspension cultures of potato (Solanum tuberosum L. cv Datura). The enzyme showed specific activities of 33 mkat (kg protein)-1 (formation of feruloyltyramine). The apparent native Mr was found to be approximately 49,000. Highest activity was at pH 6.8 in K-phosphate. The isoelectric point of the enzyme was approximately pH5.2. The apparent energy of activation was calculated to be 96 kJ mol-1. The enzyme activity was stimulated more than 5-fold by 10 mM Ca2+ or Mg2+. The apparent Km values were 36 [mu]M for feruloyl-CoA and 85 and 140 [mu]M for cinnamoyl- and 4-coumaroyl-CoA, respectively. The Km value for tyramine in the presence of feruloyl-CoA was 22 [mu]M. In the presence of 4-coumaroyl-CoA, however, the Km for tyramine increased to about 230 [mu]M. The mode of action was an iso-ordered bi bi mechanism in which A, B, P, and Q equal hydroxycinnamoyl-CoA, tyramine, N-hydroxycinnamoyltyramine, and CoA, respectively. Thus, the reaction occurred in a ternary complex of the enzyme and substrates. The equilibrium constant of the reaction was determined to be 1.3 x 104. This gave a [delta]G[deg][prime] eq value of -23.5 kJ mol-1.  相似文献   

13.
Reaction of glycolaldehyde with the binary E-NADP complex of bovine kidney aldose reductase (ALR2) produces an enzyme-bound chromophore whose absorbance (lambd max 341 nm) and fluorescence (lambda ex max 341 nm; lambda emit max 421 nm) properties are distinct from those of NADPH or E.NADPH yet are consistent with the proposed covalent adduct structure [1,4-dihydro-4-(1-hydroxy-2-oxoethyl)nicotinamide adenine dinucleotide phosphate]. The kinetics of adduct formation, both in solution and at the enzyme active site, support a mechanism involving rate-determining enolization of glycolaldehyde at high [NADP+] or [E.NADP]. At low [NADP+] or [E.NADP] the reaction is second-order overall, but the ALR2-mediated reaction displays saturation by glycolaldehyde due to competition of the aldehyde (plus hydrate) and enol for E.NADP. Measurement of the pre-steady-state burst of E-adduct formation confirms that glycolaldehyde enol is the reactive species and gives a value of 1.3 x 10(-6) for Kenol = [enol]/[( aldehyde] + [hydrate]), similar to that determined by trapping the enol with I3-. At the ALR2 active site, the rate of adduct formation is enhanced 79,000-fold and the adduct is stabilized greater than or equal to 13,000-fold relative to the reaction with NADP+ in solution. A portion of this enhancement is ascribed to specific interaction of NADP+ with the enzyme since the 3-acetylpyridine analogue, (AP)ADP+, gives values that are 15-200-fold lower. Additional evidence for strong interaction of ALR2 with both NADP+ and NADPH is reported. Yet, because dissociation of adduct is slow, catalysis of the overall adduct formation reaction by ALR2 is less than or equal to 67-fold.  相似文献   

14.
Rat kidney was shown to contain two NADPH-linked aldehyde reductases (alcohol:NADP+) oxidoreductase, EC 1.1.1.2) with different substrate affinities. The high-Km aldehyde reductase, which was purified to apparent homogeneity, had a molecular weight of 32 000 as determined by Sephadex G-100 gel filtration, and of 37 000 by polyacrylamide gel electrophoresis in the presence of sodium dodecyl sulfate. The purified enzyme reduced various aliphatic aldehydes of different carbon-chain lengths besides many chemicals containing aldehyde groups. The Km values for n-hexadecanal and n-octadecanal were 8 microM and 4 microM, respectively. Bovine serum albumin (1.8 mM) stimulated the reduction of n-hexadecanal and n-octadecanal, and increased the Vmax values by about 15-fold without changing the Km values. The kidney enzyme was not distinguishable from the brain and liver high-Km aldehyde reductases in mobility on polyacrylamide gel electrophoresis, immunological properties, peptide maps or substrate specificity.  相似文献   

15.
Lucigenin-dependent chemiluminescence together with 2-[4-iodophenyl]-3-[4-nitrophenyl]-5-[2,4-disulfophenyl]-2H tetrazolium monosodium salt (WST-1) reduction can be detected following addition of NADH to many cell types, including human sperm suspensions. Although many reports suggest that such a phenomenon is due to reactive oxygen species production, other oxygen detecting metabolite probes, such as MCLA and luminol, do not produce a chemiluminescent signal in this model system. The enzyme responsible for NADH-dependent lucigenin chemiluminescence was purified and identified as cytochrome-b5 reductase. In support of this concept, COS-7 cells overexpressing cytochrome-b5 reductase displayed at least a 3-fold increase in the previously mentioned activity compared with mock-transfected cells. Fractions containing cytochrome-b5 reductase were capable of inducing both lucigenin-dependent chemiluminescence and WST-1 reduction. Oxygen radicals clearly did not mediate the cytochrome b5-mediated activation of these probes in vitro since neither luminol nor MCLA gave a chemiluminescence response in the presence of the enzyme and the cofactor NADH. These results emphasize the importance of the direct NADH-dependent reduction of these putative superoxide-sensitive probes by cytochrome-b5 reductase even though this enzyme does not, on its own accord, produce reactive oxygen species.  相似文献   

16.
3 beta,20 alpha-Hydroxysteroid oxidoreductase was purified to homogeneity from fetal lamb erythrocytes. The Mr 35,000 enzyme utilizes NADPH and reduces progesterone to 4-pregnen-20 alpha-ol-3-one [Km = 30.8 microM and Vmax = 0.7 nmol min-1 (nmol of enzyme)-1] and 5 alpha-dihydrotestosterone to 5 alpha-androstane-3 beta, 17 beta-diol [Km = 74 microM and Vmax = 1.3 nmol min-1 (nmol of enzyme)-1]. 5 alpha-Dihydrotestosterone competitively inhibits (Ki = 102 microM) 20 alpha-reductase activity, suggesting that both substrates may be reduced at the same active site. 16 alpha-(Bromoacetoxy)progesterone competitively inhibits 3 beta- and 20 alpha-reductase activities and also causes time-dependent and irreversible losses of both 3 beta-reductase and 20 alpha-reductase activities with the same pseudo-first order kinetic t1/2 value of 75 min. Progesterone and 5 alpha-dihydrotestosterone protect the enzyme against loss of the two reductase activities presumably by competing with the affinity alkylating steroid for the active site of 3 beta,20 alpha-hydroxysteroid oxidoreductase. 16 alpha-(Bromo[2'-14C]acetoxy) progesterone radiolabels the active site of 3 beta,20 alpha-hydroxysteroid oxidoreductase wherein 1 mol of steroid completely inactivates 1 mol of enzyme with complete loss of both reductase activities. Hydrolysis of the 14C-labeled enzyme with 6 N HCl at 110 degrees C and analysis of the amino acid hydrolysate identified predominantly N pi-(carboxy[2'-14C]methyl)histidine [His(pi-CM)].(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

17.
Thioltransferase in human red blood cells: kinetics and equilibrium   总被引:2,自引:0,他引:2  
Thioltransferase from human red blood cells (HRBC TTase), coupled to GSSG reductase, catalyzed glutathione (GSH)-dependent reduction of prototype substrates hydroxyethyl disulfide (HEDS) and sodium S-sulfocysteine as well as of other homo- and heterodisulfides, including the protein mixed disulfide albumin-S-S-cysteine. Whereas apparent KM values for the substrates varied over more than a 20-fold range, the Vmax values agreed quite closely, usually within less than a factor of 2, suggesting that initial interaction of oxidized substrate with enzyme is not rate determining. HRBC TTase was inactivated by iodoacetamide (IAA), and this was prevented by pretreatment with disulfides. The pH dependence of IAA inactivation gave a remarkably low apparent pKa of 3.5, which was independent of ionic strength (0.05-2 M). At pH 6, one radiolabeled carboxyamidomethyl moiety was bound to the enzyme after treatment with [14C]IAA. This unusual thiol reactivity suggests that the active-site cysteine moiety of the TTase may be involved in a hydrogen bond with a carboxylate moiety. In contrast, the pH dependence for GSH-dependent TTase catalysis of disulfide reduction displayed an inflection point near pH 8.0, also suggesting that the initial reaction of oxidized substrate with the active-site thiol is not involved in rate determination. Two substrate kinetic studies of HRBC TTase and rat liver TTase (e.g., [GSH] and [HEDS] varied independently) gave patterns of intersecting lines on double-reciprocal plots (1/v vs 1/S), indicating a sequential mechanism for the TTase reactions, rather than a ping-pong mechanism.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

18.
Desulfovibrio gigas NCIMB 9332 cells grown in ethanol-containing medium with 0.1 microM tungstate contained a benzylviologen-linked aldehyde oxidoreductase. The enzyme was purified to electrophoretic homogeneity and found to be a homodimer with a subunit M(r) of 62,000. It contained 0.68 +/- 0.08 W, 4.8 Fe, and 3.2 +/- 0.2 labile S per subunit. After acid iodine oxidation of the purified enzyme, a fluorescence spectrum typical for form A of molybdopterin was obtained. Acetaldehyde, propionaldehyde, and benzaldehyde were excellent substrates, with apparent Km values of 12.5, 10.8, and 20 microM, respectively. The natural electron acceptor is not yet known; benzylviologen was used as an artificial electron acceptor (apparent Km, 0.55 mM). The enzyme was activated by potassium ions and strongly inhibited by cyanide, arsenite, and iodoacetate. In the as-isolated enzyme, electron paramagnetic resonance studies readily detected W(V) as a complex signal with g values in the range of 1.84 to 1.97. The dithionite-reduced enzyme exhibited a broad signal at low temperature with g = 2.04 and 1.92; this is indicative of a [4Fe-4S]1+ cluster interacting with a second paramagnet, possibly the S = 1 system of W(IV). Until now W-containing aldehyde oxidoreductases had only been found in two Clostridium strains and two hyperthermophilic archaea. The D. gigas enzyme is the first example of such an enzyme in a gram-negative bacterium.  相似文献   

19.
M Kurono  I Fujiwara  K Yoshida 《Biochemistry》2001,40(28):8216-8226
Aldose reductase (AR) is an NADPH-dependent enzyme implicated in diabetic complications. AS-3201 [(R)-(-)-2-(4-bromo-2-fluorobenzyl)-1,2,3,4-tetrahydropyrrolo[1,2-a]pyrazine-4-spiro-3'-pyrrolidine-1,2',3,5'-tetrone] is a structurally novel and potent ARI with an inhibitor constant (K(i) = 10(-)(10) M) 2000-fold lower than that of its optical antipode (S-isomer). To elucidate the inhibition modes and the stereochemical differences in their inhibitory potencies, we examined the interaction of these R- and S-isomers with AR under physiological conditions. Enzyme kinetic analysis, which was performed by using physiological substrates at 37 degrees C, showed that both isomers selectively act on the E-NADP(+) complex in both the forward and reverse reactions of AR. However, fluorometric titration analysis demonstrated that the affinities of the isomers for the E-NADP(+) complex are about the same as those for the E-NADPH complex and the apoenzyme. These results suggested that the selective binding to the E-NADP(+) complex arises from the predominance of this enzyme form during steady-state turnover rather than from binding specificity. Both the competition with a known active site-directed ARI and the protective effect on AR inactivation by N-bromosuccinimide showed that the isomers bind to the active site of the enzyme, but the thermodynamic parameters for the binding to AR indicated that additional hydrogen bonds and/or van der Waals interactions contribute to the energetic stabilization in the E-R-isomer complex. Molecular modeling, together with the deductions from spectroscopic studies, suggested that the succinimide ring and the 4-bromo-2-fluorobenzyl group of the R-isomer are optimally located for formation of a hydrogen-bonding network with AR, and that the latter benzyl group is also effective for the differentiation between AR and aldehyde reductase (a closely related enzyme).  相似文献   

20.
Intersubunit transfer of fatty acyl groups during fatty acid reduction   总被引:2,自引:0,他引:2  
Fatty acid reduction in Photobacterium phosphoreum is catalyzed in a coupled reaction by two enzymes: acyl-protein synthetase, which activates fatty acids (+ATP), and a reductase, which reduces activated fatty acids (+NADPH) to aldehyde. Although the synthetase and reductase can be acylated with fatty acid (+ATP) and acyl-CoA, respectively, evidence for acyl transfer between these proteins has not yet been obtained. Experimental conditions have now been developed to increase significantly (5-30-fold) the level of protein acylation so that 0.4-0.8 mol of fatty acyl groups are incorporated per mole of the synthetase or reductase subunit. The acylated reductase polypeptide migrated faster on sodium dodecyl sulfate-polyacrylamide gel electrophoresis than the unlabeled polypeptide, with a direct 1 to 1 correspondence between the moles of acyl group incorporated and the moles of polypeptide migrating at this new position. The presence of 2-mercaptoethanol or NADPH, but not NADP, substantially decreased labeling of the reductase enzyme, and kinetic studies demonstrated that the rate of covalent incorporation of the acyl group was 3-5 times slower than its subsequent reduction with NADPH to aldehyde. When mixtures of the synthetase and reductase polypeptides were incubated with [3H] tetradecanoic acid (+ATP) or [3H]tetradecanoyl-CoA, both polypeptides were acylated to high levels, with the labeling again being decreased by 2-mercaptoethanol or NADPH. These results have demonstrated that acylation of the reductase represents an intermediate and rate-limiting step in fatty acid reduction. Moreover, the activated acyl groups are transferred in a reversible reaction between the synthetase and reductase proteins in the enzyme mechanism.  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