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.
Microsomal preparations from carrot (Daucus carota L.) cell suspension cultures catalyze the formation of trans-5-O-caffeoyl-D-quinate (chlorogenate) from trans-5-O-(4-coumaroyl)-D-quinate. trans-5-O-(4-Coumaroyl)shikimate is converted to about the same extent to trans-5-O-caffeoylshikimate. trans-4-O-(4-Coumaroyl)-D-quinate, trans-3-O-(4-coumaroyl)-D-quinate, trans-4-coumarate, and cis-5-O-(4-coumaroyl)-D-quinate do not act as substrates. The reaction is strictly dependent on molecular oxygen and on NADPH as reducing cofactor. NADH and ascorbic acid cannot substitute for NADPH. Cytochrome c, Tetcyclacis, and carbon monoxide inhibit the reaction suggesting a cytochrome P-450-dependent mixed-function monooxygenase. Competition experiments as well as induction and inhibition phenomena indicate that there is only one enzyme species which is responsibl for the hydroxylation of the 5-O-(4-coumaric) esters of both D-quinate and shikimate. The activity of this enzyme is greatly increased by in vivo irradiation of the cells with blue/uv light. We conclude that the biosynthesis of the predominant caffeic acid conjugates in carrot cells occurs via the corresponding 4-coumaric acid esters. Thus, in this system, 5-O-(4-coumaroyl)-D-quinate can be seen as the final intermediate in the chlorogenic acid pathway.  相似文献   

2.
Under anaerobiosis, Euglena gracilis mitochondria perform a malonyl-CoA independent synthesis of fatty acids leading to accumulation of wax esters, which serve as the sink for electrons stemming from glycolytic ATP synthesis and pyruvate oxidation. An important enzyme of this unusual pathway is trans-2-enoyl-CoA reductase (EC 1.3.1.44), which catalyzes reduction of enoyl-CoA to acyl-CoA. Trans-2-enoyl-CoA reductase from Euglena was purified 1700-fold to electrophoretic homogeneity and was active with NADH and NADPH as the electron donor. The active enzyme is a monomer with molecular mass of 44 kDa. The amino acid sequence of tryptic peptides determined by electrospray ionization mass spectrometry were used to clone the corresponding cDNA, which encoded a polypeptide that, when expressed in Escherichia coli and purified by affinity chromatography, possessed trans-2-enoyl-CoA reductase activity close to that of the enzyme purified from Euglena. Trans-2-enoyl-CoA reductase activity is present in mitochondria and the mRNA is expressed under aerobic and anaerobic conditions. Using NADH, the recombinant enzyme accepted crotonyl-CoA (km=68 microm) and trans-2-hexenoyl-CoA (km=91 microm). In the crotonyl-CoA-dependent reaction, both NADH (km=109 microm) or NADPH (km=119 microm) were accepted, with 2-3-fold higher specific activities for NADH relative to NADPH. Trans-2-enoyl-CoA reductase homologues were not found among other eukaryotes, but are present as hypothetical reading frames of unknown function in sequenced genomes of many proteobacteria and a few Gram-positive eubacteria, where they occasionally occur next to genes involved in fatty acid and polyketide biosynthesis. Trans-2-enoyl-CoA reductase assigns a biochemical activity, NAD(P)H-dependent acyl-CoA synthesis from enoyl-CoA, to one member of this gene family of previously unknown function.  相似文献   

3.
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   

4.
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   

5.
Evidence for the existence of a previously unknown rat hepatic microsomal reductase, short chain trans-2-enoyl-CoA reductase (SC reductase) is presented. This reductase has a specific requirement for NADPH, is unable to utilize NADH, and catalyzes the conversion of crotonyl-CoA and trans-2-hexenoyl-CoA to butyric acid and hexenoic acid at a rate of 5 and 65 nmol per min per mg of microsomal protein, respectively. Highly purified NADPH cytochrome P-450 reductase incorporated into liposomes prepared from dilauroyl phosphatidylcholine in the presence or absence of cytochrome P-450 possesses no SC reductase activity. These liposomal preparations did, however, catalyze mixed function oxidations of benzphetamine and testosterone. Rabbit antibody to rat liver NADPH cytochrome P-450 reductase had little to no effect on the conversion of crotonyl-CoA and trans-2-hexenoyl-CoA, suggesting that the SC reductase accepts reducing equivalents directly from NADPH. When acetoacetyl-CoA was incubated with hepatic microsomes and either NADH or NADPH, no formation of butyrate was detected; however, when both cofactors were present, a rate of formation of 3 nmol of butyrate was determined per min per mg of microsomal protein. These results suggest the presence of a previously unknown short chain beta-ketoreductase which catalyzes the reduction of short chain beta-keto acids, only in the presence of NADH. Our results also indicate that the electrons from NADH to the beta-ketoreductase bypass cytochrome b5. The physiological significance is discussed in terms of lipogenesis and ketone body utilization by the liver.  相似文献   

6.
Fillgrove KL  Anderson VE 《Biochemistry》2001,40(41):12412-12421
The chemical mechanism of the 2,4-dienoyl-CoA reductase (EC 1.3.1.34) from rat liver mitochondria has been investigated. This enzyme catalyzes the NADPH-dependent reduction of 2,4-dienoyl-coenzyme A (CoA) thiolesters to the resulting trans-3-enoyl-CoA. Steady-state kinetic parameters for trans-2,trans-4-hexadienoyl-CoA and 5-phenyl-trans-2,trans-4-pentadienoyl-CoA were determined and demonstrated that the dienoyl-CoA and NADPH bind to the 2,4-dienoyl-CoA reductase via a sequential kinetic mechanism. Kinetic isotope effect studies and the transient kinetics of substrate binding support a random order of nucleotide and dienoyl-CoA addition. The large normal solvent isotope effects on V/K ((D)(2)(O)V/K) and V ((D)(2)(O)V) for trans-2,trans-4-hexadienoyl-CoA reduction indicate that a proton transfer step is rate limiting for this substrate. The stability gained by conjugating the phenyl ring to the diene in PPD-CoA results in the reversal of the rate-determining step, as evidenced by the normal isotope effects on V/K(CoA) ((D)V/K(CoA)) and V/K(NADPH) ((D)V/K(NADPH)). The reversal of the rate-determining step was supported by transient kinetics where a burst was observed for the reduction of trans-2,trans-4-hexadienoyl-CoA but not for 5-phenyl-trans-2,trans-4-pentadienoyl-CoA reduction. The chemical mechanism is stepwise where hydride transfer from NADPH occurs followed by protonation of the observable dienolate intermediate, which has an absorbance maximum at 286 nm. The exchange of the C alpha protons of trans-3-decenoyl-CoA, catalyzed by the 2,4-dienoyl-CoA reductase, in the presence of NADP(+) suggests that formation of the dienolate is catalyzed by the enzyme active site.  相似文献   

7.
Cell-free extracts of a streptomycin-bleached strain of Euglena gracilis var. bacillaris have been examined for enzyme systems primarily responsible for the oxidation of reduced pyridine nucelotides. NADH lipoyl dehydrogenase, NADH and NADPH oxidase, NADH and NADPH diaphorase, and NADH and NADPH cytochrome c reductase have been demonstrated. The NADPH-linked enzymes had lower activity rates and were less sensitive to N-ethyl maleimide and p-hydroxymercuribenzoate than their NADH-linked counterparts. NADH cytochrome c reductase was the most sensitive to antimycin A. Michaelis-Menten constants (Km) determined were as follows: NADH diaphorase, 350 muM; NADPH oxidase 150 muM ; NADH lipoyl dehydrogenase, 0.35 muM. Enzyme activities after storage at -5 C indicate that the diaphorases are less labile than the other tested enzymes, and the differential activities of the NADH and NADPH linked enzymes suggest that functionally they may have different roles.  相似文献   

8.
Glutathione reductase from Escherichia coli is inactivated when incubated with either NADPH or NADH. The process is inversely dependent on the enzyme concentration. Inactivation is rapid and monophasic with 1 microM NADPH and 1 nM enzyme FAD giving a t1/2 of 1 min. Complex formation between NADPH and the two-electron reduced enzyme (EH2) at higher levels of NADPH protects against rapid inactivation. NADP+, produced in a side reaction with oxygen, also protects by forming a complex with EH2. These complexes make analysis of the concentration dependence of the inactivation process difficult. Inactivation with NADH, where complexes do not interfere, is slower but can be analyzed more readily. With 152 microM NADH and 5.4 nM enzyme FAD, the time required for 50% inactivation is 17 min. The process is markedly biphasic, reaching the final inactivation level after 5-7 h. Analysis of the relationship between the final level of inactivation with NADH and the enzyme concentration indicates that inactivation is due to dissociation of the normally dimeric enzyme. Thus, the position of the dimer-monomer equilibrium between an active dimeric two-electron reduced species and an inactive monomeric two-electron reduced form determines the enzyme activity. An apparent equilibrium constant (Kd) for dissociation of dimer obtained from the anaerobic concentration dependent inactivation curves is 220 nM. Enzyme inactivated with NADH can be reactivated with glutathione, and the reactivation kinetics are second order, monomer-monomer over 75% of the reaction with an average apparent association rate constant (ka) of 13.1 (+/- 5.5) X 10(6) M-1 min-1.  相似文献   

9.
Biliverdin reductase was purified from pig spleen soluble fraction to a purity of more than 90% as judged by sodium dodecyl sulfate-polyacrylamide gel electrophoresis. The enzyme was a monomer protein with a molecular weight of about 34,000. Its isoelectric point was at 6.1-6.2. The enzyme was strictly specific to biliverdin and no other oxiodoreductase activities could be detected in the purified enzyme preparation. The purified enzyme could utilize both NADPH and NADH as electron donors for the reduction of biliverdin. However, there were considerable differences in the kinetic properties of the NADPH-dependent and the NADH-dependent biliverdin reductase activities: Km for NADPH was below 5 microM while that for NADH was 1.5-2 mM; the pH optimum of the reaction with NADPH was 8.5 whereas that of the reaction with NADH was 6.9; Km for biliverdin in the NADPH system was 0.3 microM whereas that in the NADH system was 1-2 microM. In addition, both the NADPH-dependent and NADH-dependent activities were inhibited by excess biliverdin, but this inhibition was far more pronounced in the NADPH system than in the NADH system. IX alpha-biliverdin was the most effective substrate among the four biliverdin isomers, and the dimethylester of IX alpha-biliverdin could not serve as a substrate. Biliverdin reductase was also purified about 300-fold from rat liver soluble fraction. The hepatic enzyme was also a monomer protein with a molecular weight of 34,000 and showed properties quite similar to those of the splenic enzyme as regards the biliverdin reductase reaction. The isoelectric point of the hepatic enzyme, however, was about 5.4. It was assumed that NADPH rather than NADH is the physiological electron donor in the intracellular reduction of IX alpha-biliverdin. The stimulatory effects of bovine and human serum albumins on the biliverdin reductase reactions were also examined.  相似文献   

10.
1. After immunization of BALB/c mouse, four monoclonal antibodies against soluble NADH diaphorase from ejaculated boar spermatozoa were produced and characterized. The monoclonal antibodies were designated as follows Mab 1F2, Mab 4E2, Mab 5B8, Mab 5D8. 2. These monoclonal antibodies react with other enzyme forms-sedimentary NADH and NADPH and soluble NADPH and inhibit (although not completely) their activity. It is supposed that different forms of the enzyme share some common epitopes. 3. Treatment of ejaculated boar semen with 2O-methylcholanthrene causes an increase of the activity of the soluble diaphorase form only. 4. These results lead to the assumption that the sperm diaphorase is a dynamic enzyme system consisting of four immunologically similar isoenzymes although their functions are different.  相似文献   

11.
Four acylated pelargonidin glycosides and pelargonidin 3-sophoroside-5-glucoside were isolated from 23 red-purple cultivars of Pharbitis nil. The acylated anthocyanins were all based on pelargonidin 3-sophoroside-5-glucoside and were identified as the 3-O-[2-O-(beta-D-glucopyranosyl)-6-O-(trans-caffeyl)-beta-D- glucopyranoside]-5-O-(beta-D-glucopyranoside), the 3-O-[2-O-(6-O-(trans-3-O-(beta-D-glucopyranosyl)caffeyl)-beta- D-glucopyranosyl)-beta-D-glucopyranoside]-5-O-(beta-D-glucopyranoside), the 3-O-[2-O-(6-O-(trans-3-O-(beta-D-glucopyranosyl)caffeyl)-beta- D-glucopyranosyl)-6-O-(trans-caffeyl)-beta-D-glucopyranoside]-5-O-(beta- D-glucopyranoside); and the 3-O-[2-O-(6-O-(trans-3-O-(beta-D-glucopyranosyl)caffeyl)-beta-D- glucopyranosyl)-6-O-(trans-4-O-(6-O-(trans-3-O-(beta-D- glucopyranosyl)caffeyl)- beta-D-glucopyranosyl)caffeyl)-beta-D-glucopyranoside]-5-O-(beta-D- glucopyranoside). By the analysis of these anthocyanin constituents variously in 23 cultivars, it was found that the red flower colour gradually changed into more bluish colour with increasing numbers of caffeic acid residues in the acylated pelargonidin glycosides. The stabilities of these anthocyanins increased in the order of increasing caffeyl substitution.  相似文献   

12.
NADPH-cytochrome P-450 (cytochrome c) reductase (EC 1.6.2.4) was solubilized by detergent from microsomal fraction of wounded Jerusalem-artichoke (Helianthus tuberosus L.) tubers and purified to electrophoretic homogeneity. The purification was achieved by two anion-exchange columns and by affinity chromatography on 2',5'-bisphosphoadenosine-Sepharose 4B. An Mr value of 82,000 was obtained by SDS/polyacrylamide-gel electrophoresis. The purified enzyme exhibited typical flavoprotein redox spectra and contained equimolar quantities of FAD and FMN. The purified enzyme followed Michaelis-Menten kinetics with Km values of 20 microM for NADPH and 6.3 microM for cytochrome c. In contrast, with NADH as substrate this enzyme exhibited biphasic kinetics with Km values ranging from 46 microM to 54 mM. Substrate saturation curves as a function of NADPH at fixed concentration of cytochrome c are compatible with a sequential type of substrate-addition mechanism. The enzyme was able to reconstitute cinnamate 4-hydroxylase activity when associated with partially purified tuber cytochrome P-450 and dilauroyl phosphatidylcholine in the presence of NADPH. Rabbit antibodies directed against plant NADPH-cytochrome c reductase affected only weakly NADH-sustained reduction of cytochrome c, but inhibited strongly NADPH-cytochrome c reductase and NADPH- or NADH-dependent cinnamate hydroxylase activities from Jerusalem-artichoke microsomal fraction.  相似文献   

13.
The malic enzyme enriched from Acinetobacter calcoaceticus is inhibited by NADPH and NADH. The inhibition afforded by the reduced coenzymes is not affected by NAD+, AMP and 3'.5'-AMP. Against L-malate, NADPH inhibits the enzyme in a noncompetitive linear fashion (Ki = 1.5 x 10(-4) M), against NADP+, competitively linearly (Ki = 5.0 x 10(-5) M). While NADPH acted as a product inhibitor, NADH seems to be an allosteric effector of the malic enzyme, because with L-malate as the variable substrate in the double reciprocal plot, a nonlinear curve is obtained.  相似文献   

14.
The reaction process of adrenodoxin reductase with NADPH and NADH were investigated. The appearance of new intermediate with a broad absorption band at around 520 nm has been detected by rapid-scan stopped-flow spectrophotometry. Although the formation of this intermediate is more rapid with NADPH than with NADH, the rates of the subsequent decay to the fully reduced state are almost identical (Kobs values were 20.5 and 16.0s-1). These results indicate that the new intermediate is the complex formed between the oxidized enzyme and reduced pyridine nucleotide (enzyme-substrate complex), and that subsequent decay of the intermidiate is caused by a two-electron transfer process from the reduced pyridine nucleotide to the enzyme flavin. On the other hand, spectral and kinetic properties in the steady state of the reoxidation reaction of the enzyme reduced with NADPH and NADH were somewhat different. The rate of reoxidation of the enzyme under aerobic conditions from the reduced state to the oxidized state was 6.5 times faster when a 10-fold molar excess of NADH was used than when NADPH of the same concentration was used. This result is consistent with the fact that the NADH-dependent oxidase activity was 6.4 times greater than that dependent on NADPH. During reoxidation of the reduced enzyme under aerobic conditions in the presence of an excess of NADPH or NADH, the EPR spectra indicated the formation of the flavin semiquinone radical species. Similarly, the formation of semiquinone was observed in the absorption spectrum with either NADPH or NADH under the same conditions as in the EPR measurement. The intensity of the semiquinone signal on EPR was considerably smaller with NADH than with NADPH. These results suggest that NADP+ complex with the enzyme semiquinone protects the radical from oxidation by oxygen to a greater extent than NAD+, and consequently the semiquinone is easier to detect with NADPH than with NADH.  相似文献   

15.
The hepatic microsomal haem oxygenase activity of rats treated with CoCl2 was studied kinetically by measuring biliverdin, the immediate product of the reaction. Biliverdin was extracted with diethyl ether/ethanol mixture, and was determined by the difference between A690 and A800. The apparent Km value for NADPH (at 50 microM-haematin) was about 0.2 microM when an NADPH-generating system was used, whereas that for NADH was about 630 microM. Essentially the same Vmax. values were obtained for both the NADH- and NADPH-dependent haem oxygenase reactions. No synergism was observed with NADH and NADPH. The NADH-dependent reaction was competitively inhibited by NADP+, with a Ki of about 10 microM. The inhibitoin of the NADH-dependent reaction by the antibody against rat liver microsomal NADPH-cytochrome c reductase was essentially complete, with a pattern similar to that of the NADPH-dependent reaction. The immunochemical experiment and the comparison of the kinetic values with the reported data on isolated NADH-cytochrome b5 reductase and NADPH--cytochrome c reductase indicated the involvement of the latter enzyme in NADH-dependent haem oxygenation by microsomal fraction in situ.  相似文献   

16.
In cell-free extracts from Pseudomonas sp. strain CBS3 the conversion of 4-chlorophenylacetate to 3,4-dihydroxyphenylacetate was demonstrated. By Sephacryl S-200 chromatography two protein fractions, A and B, were obtained which both were essential for enzyme activity. Fe2+ and NADH were cofactors of the reaction. NADPH also activated the enzyme, but less effectively than NADH. FAD had no influence on enzyme activity. 4-Hydroxyphenylacetate, 4-chloro-3-hydroxyphenylacetate, and 3-chloro-4-hydroxyphenylacetate were poor substrates for the enzyme, suggesting that these substances are not intermediates of the reaction. We therefore suggest that the reaction proceeds via a dioxygenated intermediate.  相似文献   

17.
4-Hydroxycyclohexanecarboxylate dehydrogenase, which requires NAD as a cofactor, was detected in crude soluble extracts of Corynebacterium cyclohexanicum grown on cyclohexanecarboxylic acid as the sole carbon source. The dehydrogenase was purified from extracts to an electrophoretically homogenous state by ammonium sulfate precipitation and chromatography on DEAE-650s, agarose-NAD and hydroxyapatite. The enzyme consisted of two identical subunits and had a native relative molecular mass of 53,600. There were two residues each of cysteine and tryptophan in the enzyme molecule. Oxo acid rather than hydroxy acid was routinely used as substrate for assay of the enzyme. The enzyme is highly specific for 4-oxocyclohexanecarboxylic acid: the carboxyl group is essential and the position of carbonyl group is important; neither the 2-oxo nor the 3-oxo homologue was used as substrate. A methyl substitution on the ring of 4-oxocyclohexanecarboxylate resulted in an almost complete loss of its activity. The reduction product was identified as trans-4-hydroxycyclohexanecarboxylic acid by gas-liquid chromatography and mass spectrometry. It was used as a substrate for the reverse reaction in the presence of NAD but not its cis-isomer. The enzyme was specific for the B-side (pro-S) hydrogen of NADH in the hydrogen transfer from NADH to 4-oxocyclohexanecarboxylate. The Km values for 4-oxocyclohexanecarboxylate and NADH in the reduction reaction at pH 6.8 were 0.50 mM and 0.28 mM, respectively, whereas those for trans-4-hydroxycyclohexanecarboxylate and NAD in the oxidation reaction at pH 8.8 were 0.51 mM and 0.23 mM, respectively. The equilibrium constant of the reaction was 1.79 x 10(-10) M. The enzyme was strongly inhibited by N-bromosuccinimide.  相似文献   

18.
trans-4-Phenyl-3-buten-2-one (PBO), a flavoring additive, was transformed to the carbonyl-reduced product, trans-4-phenyl-3-buten-2-ol (PBOL) by rat liver microsomes, but not by liver cytosol, in the presence of NADH or NADPH. PBOL formed was identified by comparison with an authentic sample. The reductase activity was not inhibited by quercitrin, an inhibitor of cytosolic carbonyl reductase. The carbonyl reduction product of PBO by liver microsomes was identified as the R-enantiomer of PBOL by HPLC analysis. Rat blood also exhibited the carbonyl reductase activity in the presence of NADH or NADPH, but to a lesser extent.  相似文献   

19.
1. Oxidation of NADPH by various acceptors catalyzed by submitochondrial particles and a partially purified NADH dehydrogenase from beef heart was investigated. Submitochondrial particles devoid of nicotinamide nucleotide transhydrogenase activity catalyze an oxidation of NADPH by oxygen. The partially purified NADH dehydrogenase prepared from these particles catalyzes an oxidation of NADPH by acetylpyridine-NAD. In both cases the rates of oxidation are about two orders of magnitude lower than those obtained with NADH as electron donor. 2. The kinetic characteristics of the NADPH oxidase reaction and reduction of acetylpyridine-NAD by NADPH are similar with regard to pH dependences and affinities for NADPH, indicating that both reactions involve the same binding site for NADPH. The binding of NADPH to this site appears to be rate limiting for the overall reactions. 3. At redox equilibrium NADPH and NADH reduce FMN and iron-sulphur center 1 of NADH dehydrogenase to the same extents. The rate of reduction of FMN by NADPH is at least two orders of magnitude lower than with NADH. 4. It is concluded that NADPH is a substrate of NADH dehydrogenase and that the nicotinamide nucleotide is oxidized by submitochondrial particles via the NADH--binding site of the enzyme.  相似文献   

20.
Crude extracts of Methanospirillum hungatei strain GP1 contained NADH and NADPH diaphorase activities. After a 483-fold purification of the NADH diaphorase the enzyme was further separated from contaminating proteins by polyacrylamide disc gel electrophoresis. Two distinct activity bands were extracted from the acrylamide, each one having oxygen, 2,6-dichlorophenolindophenol, and cytochrome c linked activities. In these preparations NADPH could not replace NADH as electron donor. During the initial purification steps all activity was lost due to the removal of a readily released cofactor. Enzyme activity was restored by either FAD or a FAD fraction isolated from M. hungatei. Oxidase activity exhibited a broad pH optimum from 7.0 to 8.5 and apparent Km values of 26 microM for NADH and 0.2 microM for FAD. Superoxide anion, formed in the presence of oxygen, accounted for all of the NADH consumed in the reaction. The molecular weight of the diaphorase was about 117 500 by sodium dodecyl sulfate gel electrophoresis. Sulfhydryl reagents and chelating agents were inhibitory. Inactivation, which occurred during storage in phosphate buffer at 4 degrees C, was delayed by dithiothreitol. The isolated NADH diaphorase lacked NADPH:NAD transhydrogenase and NAD reductase activities.  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