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 250 毫秒
1.
Flavonoid analysis and supplementation experiments with dihydroflavonols and leucocyanidin on two cyanic, two acyanic and one white/red-variegated flowering strain of Dianthus caryophyllus (carnation) showed that in the acyanic strains recessive alleles (aa) of the gene A interrupt the anthocyanin pathway between dihydroflavonols and leucoanthocyanidins. The instability in the variegated strain involves the same step and is obviously caused by the multiple allele a var . In confirmation of these results, dihydroflavonol 4-reductase activity could be demonstrated in enzyme extracts from cyanic flowers and cyanic parts of variegated flowers but not in preparations from acyanic flowers or acyanic parts. The enzyme catalyzes the stereospecific reduction of (+)dihydrokaempferol to (+)-3,4-leucopelargonidin with NADPH as cofactor. A pH optimum around 7.0 and a temperature optimum at 30° C was determined, but the reduction reaction also proceeded at low temperatures. (+)Dihydroquercetin and (+)dihydromyricetin were also reduced to the respective flavan-3,4-cis-diols by the enzyme preparations from carnation flowers, and were even better substrates than dihydrokaempferol.These investigations were supported by grants from Fonds zur Förderung der wissenschaftlichen Forschung and Deutsche Forschungsgemeinschaft. The authors thank the market-gardens Ing. K. Rungaldier (Vienna, Austria), A. Sinner (Tübingen, FRG) and Barbaret & Blanc GMBH (Horhausen, FRG) for generous support with plant material.  相似文献   

2.
Halbwirth H  Stich K 《Phytochemistry》2006,67(11):1080-1087
Yellow flavonols contribute to flower pigmentation in Asteraceae. In contrast to common flavonols, they show additional hydroxyl groups in position 6 and/or 8 of the aromatic A-ring in addition to the basic 5,7-hydroxylation pattern. An enzyme introducing a hydroxyl group in position 8 of flavonols and flavones was demonstrated for the first time with enzyme preparations from petals of Chrysanthemum segetum. Flavanones, dihydroflavonols and glucosylated flavonols and flavones were not accepted as substrates. The enzyme was localized in the microsomal fraction and uses NADPH and FAD as cofactors. Experiments with carbon monoxide/blue light and with antibodies specific for cytochrome P450 reductase did not indicate the involvement of a classical cytochrome P450 dependent monooxygenase in the reaction. Thus, the flavonoid 8-hydroxylase represents a novel type of hydroxylating enzyme in the flavonoid pathway. Apart from flavonoid 8-hydroxylase activity, the presence of all enzymes involved in the formation of flavonoid 7-O-glucosides in C. segetum was demonstrated. The pathway leading to 8-hydroxyflavonoids in C. segetum has been derived from enzyme activities and substrate specificities observed.  相似文献   

3.
Evidence suggesting that Bacillus polymyxa has an active ferredoxin-NADP(+) reductase (EC 1.6.99.4) was obtained when NADPH was found to provide reducing power for the nitrogenase of this organism; direct evidence was provided when it was shown that B. polymyxa extracts could substitute for the native ferredoxin-NADP(+) reductase in the photochemical reduction of NADP(+) by blue-green algal particles. The ferredoxin-NADP(+) reductase was purified about 80-fold by a combination of high-speed centrifugation, ammonium sulfate fractionation, and chromatography on Sephadex G-100 and diethylaminoethyl-cellulose. The molecular weight was estimated by gel filtration to be 60,000. A small amount of the enzyme was further purified by polyacrylamide gel electrophoresis and shown to be a flavoprotein. The reductase was specific for NADPH in the ferredoxin-dependent reduction of cytochrome c and methyl viologen diaphorase reactions; furthermore, NADP(+) was the acceptor of preference when the electron donor was photoreduced ferredoxin. The reductase also has an irreversible NADPH-NAD(+) transhydrogenase (reduced-NADP:NAD oxidoreductase, EC 1.6.1.1) activity, the rate of which was proportional to the concentration of NAD (K(m) = 5.0 x 10(-3)M). The reductase catalyzed electron transfer from NADPH not only to B. polymyxa ferredoxin but also to the ferredoxins of Clostridium pasteurianum, Azotobacter vinelandii, and spinach chloroplasts, although less effectively. Rubredoxin from Clostridium acidi-urici and azotoflavin from A. vinelandii also accept electrons from the B. polymyxa reductase. The pH optima for the various reactions catalyzed by the B. polymyxa ferredoxin-NADP reductase are similar to those of the chloroplast reductase. NAD and acetyl-coenzyme A, which obligatorily activate NADPH- and NADH-ferredoxin reductases, respectively, in Clostridium kluyveri, have no effect on B. polymyxa reductase.  相似文献   

4.
NADPH is an intermediate in the oxidation of organic compounds coupled to Fe(III) reduction in Geobacter species, but Fe(III) reduction with NADPH as the electron donor has not been studied in these organisms. Crude extracts of Geobacter sulfurreducens catalyzed the NADPH-dependent reduction of Fe(III)-nitrilotriacetic acid (NTA). The responsible enzyme, which was recovered in the soluble protein fraction, was purified to apparent homogeneity in a four-step procedure. Its specific activity for Fe(III) reduction was 65 micromol. min(-1). mg(-1). The soluble Fe(III) reductase was specific for NADPH and did not utilize NADH as an electron donor. Although the enzyme reduced several forms of Fe(III), Fe(III)-NTA was the preferred electron acceptor. The protein possessed methyl viologen:NADP(+) oxidoreductase activity and catalyzed the reduction of NADP(+) with reduced methyl viologen as electron donor at a rate of 385 U/mg. The enzyme consisted of two subunits with molecular masses of 87 and 78 kDa and had a native molecular mass of 320 kDa, as determined by gel filtration. The purified enzyme contained 28.9 mol of Fe, 17.4 mol of acid-labile sulfur, and 0.7 mol of flavin adenine dinucleotide per mol of protein. The genes encoding the two subunits were identified in the complete sequence of the G. sulfurreducens genome from the N-terminal amino acid sequences derived from the subunits of the purified protein. The sequences of the two subunits had about 30% amino acid identity to the respective subunits of the formate dehydrogenase from Moorella thermoacetica, but the soluble Fe(III) reductase did not possess formate dehydrogenase activity. This soluble Fe(III) reductase differs significantly from previously characterized dissimilatory and assimilatory Fe(III) reductases in its molecular composition and cofactor content.  相似文献   

5.
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   

6.
A cell-free extract from flowers of Matthiola incana catalyzes a NADPH-dependent stereospecific reduction of (+)-dihydrokaempferol to 3,4-cis-leucopelargonidin (5,7,4-trihydroxyflavan-3,4-cis-diol). The pH-optimum of this reaction is around 6. The rate of reaction with NADH was about 50% of that found with NADPH. (+)-Dihydroquercetin and (+)-dihydromyricetin were also reduced by the enzyme preparation to the corresponding flavan-3,4-cis-diols. Correlation between the genotype of M. incana and the presence of dihydroflavonol 4-reductase is strong evidence that this enzyme is involved in anthocyanin biosynthesis.  相似文献   

7.
Chromosomal dihydrofolate reductase from Escherichia coli catalyzes the reduction of dihydrofolate to tetrahydrofolate using NADPH as a cofactor. The thermodynamics of ligand binding were examined using an isothermal titration calorimetry approach. Using buffers with different heats of ionization, zero to a small, fractional proton release was observed for dihydrofolate binding, while a proton was released upon NADP(+) binding. The role of water in binding was additionally monitored using a number of different osmolytes. Binding of NADP(+) is accompanied by the net release of ~5-24 water molecules, with a dependence on the identity of the osmolyte. In contrast, binding of dihydrofolate is weakened in the presence of osmolytes, consistent with "water uptake". Different effects are observed depending on the identity of the osmolyte. The net uptake of water upon dihydrofolate binding was previously observed in the nonhomologous R67-encoded dihydrofolate reductase (dfrB or type II enzyme) [Chopra, S., et al. (2008) J. Biol. Chem. 283, 4690-4698]. As R67 dihydrofolate reductase possesses a nonhomologous sequence and forms a tetrameric structure with a single active site pore, the observation of weaker DHF binding in the presence of osmolytes in both enzymes implicates cosolvent effects on free dihydrofolate. Consistent with this analysis, stopped flow experiments find betaine mostly affects DHF binding via changes in k(on), while betaine mostly affects NADPH binding via changes in k(off). Finally, nonadditive enthalpy terms when binary and ternary cofactor binding events are compared suggest the presence of long-lived conformational transitions that are not included in a simple thermodynamic cycle.  相似文献   

8.
Slade D  Ferreira D  Marais JP 《Phytochemistry》2005,66(18):2177-2215
Circular dichroism is a powerful tool for establishing the absolute configuration of flavonoids and proanthocyanidin analogues. It has been utilized to study the configuration of flavanones, dihydroflavonols (3-hydroxyflavanones), flavan-3-ols, flavan-4-ols, flavan-3,4-diols, flavans, isoflavans, isoflavanones, pterocarpans, 6a-hydroxypterocarpans, rotenoids, 12a-hydroxyrotenoids, neoflavonoids, 3,4-dihydro-4-arylcoumarins, 4-arylflavan-3-ols, auronols, homoisoflavanones, proanthocyanidins, and various classes of biflavonoids. Results relevant to the correlation of circular dichroic data and the absolute configuration of the diastereoisomers of some of the above classes of compounds will be discussed.  相似文献   

9.
3-Hydroxy-3-methylglutaryl coenzyme A reductase has been purified from rat liver microsomes with a recovery of approx. 25%. The enzyme was homogeneous on gel electrophoresis and enzyme activity comigrated with the single protein band. The molecular weight of the reductase determined by gel filtration on Sephadex G-200 was 200,000. SDS-polyacrylamide gel electrophoresis gave a subunit molecular weight of 52,000 +/- 2000, suggesting that the enzyme was a tetramer. The specific activities of the purified enzyme obtained from rats fed diets containing 0% or 5% cholestyramine were 11,303 and 19,584 nmol NADPH oxidized/min per mg protein, respectively. The reductase showed unique binding properties to Cibacron Blue Sepharose; the enzyme was bound to the Cibacron Blue via the binding sites for both substrates, NADPH and (S)-3-hydroxy-3-methylglutaryl coenzyme A. Antibodies prepared against purified reductase inactivated 100% of the soluble and at least 91% of the microsomal enzyme activity. Immunotitrations of solubilized enzyme obtained from normal and cholestyramine-fed rats indicated that cholestyramine feeding both increased the amount of enzyme protein and resulted in enzyme activation. Administration of increasing amounts of mevalonolactone to rats decreased the equivalence point obtained from immunotitration studies with solubilized enzyme. These data indicate that the antibody cross-reacts with the inactive enzyme formed after mevalonolactone treatment.  相似文献   

10.
Anthocyanins are colorful plant pigments with promising applications as pharmaceuticals and colorants. In order to engineer efficient pigment biosynthesis in Escherichia coli, the activities of various dihydroflavonol 4-reductases (DFRs) were characterized for the three primary dihydroflavonol substrates. The biochemical assays demonstrated variable DFR activities for dihydroflavonol with one B-ring hydroxyl group, the precursor of pelargonidin derivatives. In contrast, dihydroflavonols with two and three B-ring hydroxylation were metabolized with comparable efficiency. Furthermore, the catalysis of DFR for the secondary substrates, flavanones, also depended on the number of B-ring hydroxyl groups. Engineering the expression of the DFR clones together with plant-specific 4-coumaroyl:CoA ligase, chalcone synthase, chalcone isomerase, and flavanone 3-hydroxylase in E. coli resulted in the synthesis of pelargonidin at various levels, from p-coumaric acids. The identification of a robust DFR from this study can also be used for engineering recombinant synthesis of other bioactive flavonoids, such as flavan-3-ols.  相似文献   

11.
The reduced nicotinamide adenine dinucleotide phosphate (NADPH):nitrate oxidoreductase (EC 1.6.6.2) from Aspergillus nidulans wild-type bi-1 was purified by means of salt fractionation, gel filtration, affinity chromatography, and polyacrylamide gel electrophoresis. Enzyme which was adsorbed on Cibacron blue agarose could be eluted with 2 mM NADPH or 2 mM oxidized NADP (NADP(+)), the former being about three times more effective than the latter. About half the total NADPH:nitrate reductase activity adsorbed on agarose required elution with 1 M NaCl. This salt-elutable form remained active with NADPH and was not converted to the NADPH-elutable form after readsorption on Cibacron blue agarose. The NADPH-eluted enzyme exhibited a markedly different electrophoretic mobility than the enzyme eluted with NADP(+) or NaCl. After electrophoresis on polyacrylamide gels, the NADPH-eluted NADPH:nitrate reductase was separated into four proteins, two of which contained nonheme iron and exhibited reduced methyl viologen-nitrate reductase activity. None of these proteins, singly or in combination, reduced nitrate with NADPH as substrate. Difference spectra analyses and specific heme iron stains revealed the presence of cytochrome b(557) in the largest of the proteins. The molecular weights of the four proteins, which were determined from the relationship of their mobilities on varied concentrations of acrylamide gel, were 360,000, 300,000, 240,000, and 118,000. The subunit molecular weights of these, which are determined via sodium dodecyl sulfate slab gel electrophoresis, were 49,000, 50,000, and 75,000. The key role of NADPH in maintenance of the active form of the heteromultimer is further substantiated.  相似文献   

12.
The kinetic mechanism of NADPH-dependent aldehyde reductase II and aldose reductase, purified from human placenta, has been studied using L-glucuronate and DL-glyceraldehyde as their respective substrates. For aldehyde reductase II, the initial velocity and product inhibition studies (using NADP and gulonate) indicate that the enzyme reaction sequence is ordered with NADPH binding to the free enzyme and NADP being the last product to be released. Inhibition patterns using menadione (an analog of the aldehydic substrate) and ATP-ribose (an analog of NADPH) are also consistent with a compulsory ordered reaction sequence. Isotope effects of deuterium-substituted NADPH (NADPD) also corroborate the above reaction scheme and indicate that hydride transfer is not the sole rate-limiting step in the reaction sequence. For aldose reductase, initial velocity patterns, product, and dead-end inhibition studies indicate a random binding pattern of the substrates and an ordered release of product; the coenzyme is released last. A steady-state random mechanism is also consistent with deuterium isotope effects of NADPD on the reaction sequence catalyzed by this enzyme. However, the hydride transfer step seems to be more rate determining for aldose reductase than for aldehyde reductase II.  相似文献   

13.
Anthocyanins are colorful plant pigments with promising applications as pharmaceuticals and colorants. In order to engineer efficient pigment biosynthesis in Escherichia coli, the activities of various dihydroflavonol 4-reductases (DFRs) were characterized for the three primary dihydroflavonol substrates. The biochemical assays demonstrated variable DFR activities for dihydroflavonol with one B-ring hydroxyl group, the precursor of pelargonidin derivatives. In contrast, dihydroflavonols with two and three B-ring hydroxylation were metabolized with comparable efficiency. Furthermore, the catalysis of DFR for the secondary substrates, flavanones, also depended on the number of B-ring hydroxyl groups. Engineering the expression of the DFR clones together with plant-specific 4-coumaroyl:CoA ligase, chalcone synthase, chalcone isomerase, and flavanone 3-hydroxylase in E. coli resulted in the synthesis of pelargonidin at various levels, from p-coumaric acids. The identification of a robust DFR from this study can also be used for engineering recombinant synthesis of other bioactive flavonoids, such as flavan-3-ols.  相似文献   

14.
The methylenetetrahydrofolate reductase from the carbon-monoxide-utilizing homoacetogen Peptostreptococcus productus (strain Marburg) has been purified to apparent homogeneity. The purified enzyme catalyzed the oxidation of NADH with methylenetetrahydrofolate as the electron acceptor at a specific activity of 380 mumols.min-1 mg protein-1 (37 degrees C; pH 5.5). The apparent Km for NADH was near 10 microM. The apparent molecular mass of the enzyme was determined by gel filtration to be approximately 250.0 kDa. The enzyme consists of eight identical subunits with a molecular mass of 32 kDa. It contains 4 FAD/mol octamer which were reduced by the enzyme with NADH as the electron donor; iron could not be detected. Oxygen had no effect on the enzyme. Ultracentrifugation of cell extracts revealed that about 40% of the enzyme activity was recovered in the particulate fraction, suggesting that the enzyme is associated with the membrane. The enzyme also catalyzed the methylenetetrahydrofolate reduction with methylene blue as an artificial electron donor. The oxidation of methyltetrahydrofolate was mediated with methylene blue as the electron acceptor; neither NAD+ nor viologen dyes could replace methylene blue in this reaction. NADP(H) or FAD(H2) were not used to substrates for the reaction in either direction. The activity of the purified enzyme, which was proposed to be involved in sodium translocation across the cytoplasmic membrane, was not affected by the absence or presence of added sodium. The properties of the enzyme differ from those of the ferredoxin-dependent methylenetetrahydrofolate reductase of the homoacetogen Clostridium formicoaceticum and of the NADP(+)-dependent reductase of eucaryotes investigated so far.  相似文献   

15.
An NAD(+)-dependent alcohol dehydrogenase was purified to homogeneity from Nocardia fusca AKU 2123. The enzyme catalyzed (S)-specific oxidation of 3-pentyn-2-ol (PYOH), i.e., part of the stereoinversion reaction for the production of (R)-PYOH, which is a valuable chiral building block for pharmaceuticals, from the racemate. The enzyme used a broad variety of secondary alcohols including alkyl alcohols, alkenyl alcohols, acetylenic alcohols, and aromatic alcohols as substrates. The oxidation was (S)-isomer specific in every case. The K(m) and Vmax for (S)-PYOH and (S)-2-hexanol oxidation were 1.6 mM and 53 mumol/min/mg, and 0.33 mM and 130 mumol/min/mg, respectively. The enzyme also catalyzed stereoselective reduction of carbonyl compounds. (S)-2-Hexanol and ethyl (R)-4-chloro-3-hydroxybutanoate in high optical purity were produced from 2-hexanone and ethyl 4-chloro-3-oxobutanoate by the purified enzyme, respectively. The K(m) and Vmax for 2-hexanone reduction were 2.5 mM and 260 mumol/min/mg. The enzyme has a relative molecular mass of 150,000 and consists of four identical subunits. The NH2-terminal amino acid sequence of the enzyme shows similarity with those of the carbonyl reductase from Rhodococcus erythropolis and phenylacetaldehyde reductase from Corynebacterium sp.  相似文献   

16.
A purified NADPH-cytochrome c reductase (NADPH: ferricytochrome oxidoreductase, EC 1.6.2.4) was prepared from swine testis microsomes by detergent solubilization followed by a procedure including chromatofocusing. The reductase was eluted at an isoelectric point of 4.8 from the chromatofocusing column. 730-fold purification was achieved with an overall yield of 1.2%. The preparation was found to be homogeneous upon polyacrylamide gel electrophoresis in the absence of sodium dodecyl sulfate (SDS). Upon SDS-polyacrylamide gel electrophoresis, however, the purified preparation resolved into one major band (Mr 78 000) and two minor bands (Mr 60 000 and 15 000). The enzyme contained about 1 mol each of FMN and FAD, which were both extractable with trichloroacetic acid and also boiling water. The oxidized form of the enzyme showed the absorption spectrum of a typical flavoprotein. Aerobic reduction with NADPH resulted in conversion of the spectrum into one of an air-stable semiquinone form. The activity of the purified preparation was 26 mumol cytochrome c reduced/min per mg protein under the standard assay conditions at 22 degrees C. The enzyme catalyzed the reaction through a ping-pong mechanism.  相似文献   

17.
The formation of delta-aminolevulinic acid, the first committed precursor in porphyrin biosynthesis, occurs in certain bacteria and in the chloroplasts of plants and algae in a three-step, tRNA-dependent transformation of glutamate. Glutamyl-tRNA reductase, the second enzyme of this pathway, reduces the activated carboxyl group of glutamyl-tRNA (Glu-tRNA) in the presence of NADPH and releases glutamate 1-semialdehyde (GSA). We have purified Glu-tRNA reductase from Chlamydomonas reinhardtii by employing six different chromatographic separations. The apparent molecular mass of the protein when analyzed under both denaturing (sodium dodecyl sulfate-polyacrylamide gel electrophoresis) and nondenaturing conditions (rate zonal sedimentation on glycerol gradients) was 130,000 Da; this indicates that the active enzyme is a monomer. In the presence of NADPH Glu-tRNA reductase catalyzed the reduction to GSA of glutamate acylated to the homologous tRNA. Thus, the reductase alone is sufficient for conversion of Glu-tRNA to GSA. In the absence of NADPH, a stable complex of Glu-tRNA reductase with Glu-tRNA can be isolated.  相似文献   

18.
Previous studies have shown that the interaction of P450 reductase with bound NADP(H) is essential to ensure fast electron transfer through the two flavin cofactors. In this study we investigated in detail the interaction of the house fly flavoprotein with NADP(H) and a number of nucleotide analogues. 1,4,5,6-Tetrahydro-NADP, an analogue of NADPH, was used to characterize the interaction of P450 reductase with the reduced nucleotide. This analogue is inactive as electron donor, but its binding affinity and rate constant of release are very close to those for NADPH. The 2'-phosphate contributes about 5 kcal/mol of the binding energy of NADP(H). Oxidized nicotinamide does not interact with the oxidized flavoprotein, while reduced nicotinamide contributes 1.3 kcal/mol of the binding energy. Oxidized P450 reductase binds NADPH with a K(d) of 0.3 microM, while the affinity of the reduced enzyme is considerably lower, K(d) = 1.9 microM. P450 reductase catalyzes a transhydrogenase reaction between NADPH and oxidized nucleotides, such as thionicotinamide-NADP(+), acetylpyridine-NADP(+), or [(3)H]NADP(+). The reverse reaction, reduction of [(3)H]NADP(+) by the reduced analogues, is also catalyzed by P450 reductase. We define the mechanism of the transhydrogenase reaction as follows: NADPH binding, hydride ion transfer, and release of the NADP(+) formed. An NADP(+) or its analogue binds to the two-electron-reduced flavoprotein, and the electron-transfer steps reverse to transfer hydride ion to the oxidized nucleotide, which is released. Measurements of the flavin semiquinone content, rate constant for NADPH release, and transhydrogenase turnover rates allowed us to estimate the steady-state distribution of P450 reductase species during catalysis, and to calculate equilibrium constants for the interconversion of catalytic intermediates. Our results demonstrate that equilibrium redox potentials of the flavin cofactors are not the sole factor governing rapid electron transfer during catalysis, but conformational changes must be considered to understand P450 reductase catalysis.  相似文献   

19.
Chloroplast ferredoxin-NADP(+) reductase has a 32,000-fold preference for NADPH over NADH, consistent with its main physiological role of NADP(+) photoreduction for de novo carbohydrate biosynthesis. Although it is distant from the 2'-phosphoryl group of NADP(+), replacement of the C-terminal tyrosine (Tyr(308) in the pea enzyme) by Trp, Phe, Gly, and Ser produced enzyme forms in which the preference for NADPH over NADH was decreased about 2-, 10-, 300-, and 400-fold, respectively. Remarkably, in the case of the Y308S mutant, the k(cat) value for the NADH-dependent activity approached that of the NADPH-dependent activity of the wild-type enzyme. Furthermore, difference spectra of the NAD(+) complexes revealed that the nicotinamide ring of NAD(+) binds at nearly full occupancy in the active site of both the Y308G and Y308S mutants. These results correlate well with the k(cat) values obtained with these mutants in the NADH-ferricyanide reaction. The data presented support the hypothesis that specific recognition of the 2'-phosphate group of NADP(H) is required but not sufficient to ensure a high degree of discrimination against NAD(H) in ferredoxin-NADP(+) reductase. Thus, the C-terminal tyrosine enhances the specificity of the reductase for NADP(H) by destabilizing the interaction of a moiety common to both coenzymes, i.e. the nicotinamide.  相似文献   

20.
The present study describes the solubilization and purification of a NADPH-specific trans-2-enoyl-CoA reductase from rat liver microsomes. The final preparation was purified to near homogeneity and had a minimal molecular weight of 51,000 +/- 2,000, as judged by sodium dodecylsulfate (SDS)-polyacrylamide gel electrophoresis. This enzyme specifically used NADPH, as cofactor, and was chromatographically (2',5'-ADP-agarose) separated from another trans-2-enoyl-CoA reductase which utilized either NADH or NADPH as cofactor. The NADPH-specific trans-2-enoyl-CoA reductase catalyzed the reduction of trans-2-enoyl-CoAs from 4 to 16 carbon units. The Km values for crotonyl-CoA, trans-2-hexenoyl-CoA, and trans-2-hexadecenoyl-CoA were 20, 0.5, and 1.0 microM, while the Km value for NADPH was 10 microM. Although N-ethylmaleimide, heat treatment, and limited proteolysis with trypsin affected the reduction of short-chain (C4) and long-chain (C16) substrates equally, and in spite of the fact that a single protein band was observed on SDS-gels, at the present time one cannot state unequivocally that the purified preparation contained only one reductase. trans-2-Hexenoyl-CoA, for example, did not inhibit the reduction of trans-2-hexadecenoyl-CoA to palmitoyl-CoA and trans-2-decenoyl-CoA to decanoyl-CoA whereas it strongly inhibited the conversion of crotonyl-CoA to butyryl-CoA. The potential implications of this finding are discussed. Finally, the reductase preparation was shown not to contain either heme, nonheme iron, or a flavin prosthetic group.  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