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.
Uridine 5′-diphosphate glucose 4-epimerase (EC 5.1.3.2) from Ehrlich ascites carcinoma cells was purified to apparent homogeneity using conventional procedures and NAD-hexane-agarose affinity chromatography. The protein had a molecular weight of 96,000. The ascites enzyme had an absolute requirement for exogenously added NAD (10 ΜM) for stability. This appears to be a unique feature of ascites epimerase since epimerase from other mammalian sources did not exhibit such a dependence. Exogenously added NAD was also needed for catalysis with an apparentK m value of 2.5 ΜM. NADH was a very potent competitive inhibitor (K i = 0.11 ΜM with respect to NAD) of the enzyme activity at pH values close to intracellular pH. The dependence of the enzyme on NAD for stability and its inhibition by NADH may have some potential significance in tumor metabolism  相似文献   

2.
Rat ovarian 20α-hydroxysteroid dehydrogenase was purified 230-fold with a 48% recovery through a 3-step process involving hydrophobic, gel filtration and gree dye affinity chromatography. The purified enzyme was demonstrated to be a single polypeptide chain of Mr 36 000. Initial velocity studies of all four substrates in the forward and reverse reactions indicated a sequential mechanism for the enzyme. Product inhibition and dead-end inhibition studies with substrate analogs were consistent with an ordered bi-bi mechanism in which NADP is the first substrate bound to the enzyme and NADPH, the second product released, Several NADP analogs were demonstrated to function as coenzymes in the reaction catalyzed. The purified enzyme was denatured at moderate temperatures and the binding of NADP protected the enzyme against thermal denaturation.  相似文献   

3.
Kinetics of carbonyl reductase from human brain.   总被引:3,自引:1,他引:2       下载免费PDF全文
Initial-rate analysis of the carbonyl reductase-catalysed reduction of menadione by NADPH gave families of straight lines in double-reciprocal plots consistent with a sequential mechanism being obeyed. The fluorescence of NADPH was increased up to 7-fold with a concomitant shift of the emission maximum towards lower wavelength in the presence of carbonyl reductase, and both NADPH and NADP+ caused quenching of the enzyme fluorescence, indicating formation of a binary enzyme-coenzyme complex. Deuterium isotope effects on the apparent V/Km values decreased with increasing concentrations of menadione but were independent of the NADPH concentration. The results, together with data from product inhibition studies, are consistent with carbonyl reductase obeying a compulsory-order mechanism, NADPH binding first and NADP+ leaving last. No significant differences in the kinetic properties of three molecular forms of carbonyl reductase were detectable.  相似文献   

4.
M Yamaguchi  Y Hatefi 《Biochemistry》1989,28(14):6050-6056
The mitochondrial nicotinamide nucleotide transhydrogenase is a dimeric enzyme of monomer Mr 110,000. It is located in the inner mitochondrial membrane and catalyzes hydride ion transfer between NAD(H) and NADP(H) in a reaction that is coupled to proton translocation across the inner membrane. The amino acid sequence and the nucleotide binding sites of the enzyme have been determined [Yamaguchi, M., Hatefi, Y., Trach, K., & Hoch, J.A. (1988) J. Biol. Chem. 263, 2761-2767; Wakabayashi, S., & Hatefi, Y. (1987) Biochem. Int. 15, 915-924]. N-Ethylmaleimide, as well as other sulfhydryl group modifiers, inhibits the transhydrogenase. The presence of NADP in the incubation mixture suppressed the inhibition rate by N-ethylmaleimide, and the presence of NADPH greatly increased it. NAD and NADH had little or no effect. The NADPH effect was concentration dependent and saturable, with a half-maximal NADPH concentration effect close to the Km of the enzyme for NADPH. Study of the effect of pH on the N-ethylmaleimide inhibition rate showed that NADPH binding by the enzyme lowers the apparent pKa of the N-ethylmaleimide-sensitive group by 0.4 of a pH unit and NADP binding raises this pKa by 0.4 of a pH unit, thus providing a rationale for the effects of NADP and NADPH on the N-ethylmaleimide inhibition rate. With the use of N-[3H]ethylmaleimide, the modified sulfhydryl group involved in the NADP(H)-modulated inhibition of the transhydrogenase was identified as that belonging to Cys-893, which is located 113 residues upstream of the tyrosyl residue modified by [p-(fluorosulfonyl)benzoyl]-5'-adenosine at the putative NADP(H) binding site of the enzyme (see above references).(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

5.
D.K. Srivastava  L.E. Anderson 《BBA》1983,724(3):359-369
Glucose-6-phosphate dehydrogenase (D-glucose-6-phosphate: NADP oxidoreductase, EC 1.1.1.49) has been purified to electrophoretic homogeneity from pea chloroplasts. The enzyme, which has a Stokes radius of 52 Å, is a tetramer made up of four 56000 Da monomers. The pH optimum is around 8.2. The enzyme is absolutely specific for NADP. The apparent Km(NADP) is 2.4 ± 0.1 μM. NADPH inhibition of the enzyme is competitive with respect to NADP (mean Ki, 18 ± 5 μM) and is mixed (Kp >Km, Vmax >Vp) with respect to glucose 6-phosphate (mean crossover point, 0.5 ± 0.1 mM). The apparent Km(glucose 6-phosphate) is 0.37 ± 0.01 mM. The purified enzyme is inactivated in the light in the presence of dilute stroma and washed thylakoids, and by dithiothreitol. Enzyme which has been partially inactivated by treatment with dithiothreitol can be further inactivated in the light in the presence of dilute stroma and washed thylakoids and reactivated in the dark, but only to the extent of the reverse of light inactivation. Dithiothreitol-inactivated enzyme is not reactivated further by addition of crude stroma or oxidized thioredoxin. Dithiothreitol-dependent inactivation of the enzyme follows pseudo-first-order kinetics and shows rate saturation. The enzyme which has been partially inactivated by treatment with dithiothreitol does not differ from the untreated control with respect to thermal and tryptic inactivation. However, enzyme which has been partially light inactivated shows different thermal and tryptic inactivation patterns as compared to the dark control. These observations suggest that the changes in the enzyme brought about by light modulation are not necessarily identical with those brought about by dithiothreitol inactivation.  相似文献   

6.
NADP-isocitrate dehydrogenase from nodules of pigeonpea (Cajanus cajan L. cv UPAS-120) was partially purified to about 57 folds and its properties were studied. The enzyme showed an absolute requirement for a divalent cation which was fulfilled either by Mn+2 or Mg+2 and to a smaller extent by Co+2. The enzyme exhibited a sigmoidal response to increasing concentrations of Mn2+ (S0.5=0.3mM). The apparent Km values for isocitrate, NADP and Mg2+ were 21, 23 and 280 μM, respectively. It had an optimum pH of 8.0–8.2. The enzyme activity was not affected by various organic acids, amino acids and amides. NADH inhibited the activity non-competitively with respect to NADP. An apparent inhibition by ATP and ADP was due to chelation of divalent cation. NADPH acted competitively against NADP and non-competitively against isocitrate. Glutamate caused uncompetitive inhibition with respect to NADP and competitive against isocitrate. Kinetic studies suggested the reaction mechanism to be probably random sequential. Possible regulation of the enzyme activity in the nodules via cellular redox state and the levels of reaction products is discussed.  相似文献   

7.
6-Phosphogluconate dehydrogenase (6PG) was purified from rat small intestine with 36% yield and a specific activity of 15 U/mg. On SDS/PAGE, one band with a mass of 52 kDa was found. On native PAGE three protein and two activity bands were observed. The pH optimum was 7.35. Using Arrhenius plots, Ea, ΔH, Q10 and Tm for 6PGD were found to be 7.52 kcal/mol, 6.90 kcal/mol, 1.49 and 49.4°C, respectively. The enzyme obeyed “Rapid Equilibrium Random Bi Bi” kinetic model with Km values of 595 ± 213 μM for 6PG and 53.03±1.99 μM for NADP. 1/Vm versus 1/6PG and 1/NADP plots gave a Vm value of 8.91±1.92 U/mg protein. NADPH is the competitive inhibitor with a Ki of 31.91±1.31 μM. The relatively small Ki for the 6PGD:NADPH complex indicates the importance of NADPH in the regulation of the pentose phosphate pathway through G6PD and 6PGD.  相似文献   

8.
Bacteroids formed by Mesorhizobium ciceri CC 1192 in symbiosis with chickpea plants (Cicer arietinum L.) contained a single form of citrate synthase [citrate oxaloacetate-lyase (CoA-acetylating) enzyme; EC 4.1.3.7], which had the same electrophoretic mobility as the enzyme from the free-living cells. The citrate synthase from CC 1192 bacteroids had a native molecular mass of 228 ± 32 kDa and was activated by KCl, which also enhanced stability. Double reciprocal plots of initial velocity against acetyl-CoA concentration were linear, whereas the corresponding plots with oxaloacetate were nonlinear. The K m value for acetyl-CoA was 174 μM in the absence of added KCl, and 88 μM when the concentration of KCl in reaction mixtures was 100 mM. The concentrations of oxaloacetate for 50% of maximal activity were 27 μM without added KCl and 14 μM in the presence of 100 mM KCl. Activity of citrate synthase was inhibited 50% by 80 μM NADH and more than 90% by 200 μM NADH. Inhibition by NADH was linear competitive with respect to acetyl-CoA (K is = 23.1 ± 3 μM) and linear noncompetitive with respect to oxaloacetate (K is = 56 ± 3.8 μM and K ii = 115 ± 15.4 μM). NADH inhibition was relieved by NAD+ and by micromolar concentrations of 5′-AMP. In the presence of 50 or 100 mM KCl, inhibition by NADH was apparent only when the proportion of NADH in the nicotinamide adenine dinucleotide pool was greater than 0.6. In the microaerobic environment of bacteroids, NADH may be at concentrations that are inhibitory for citrate synthase. However, this inhibition is likely to be relieved by NAD+ and 5′-AMP, allowing carbon to enter the tricarboxylic acid cycle. Received: 14 July 1999 / Accepted: 20 September 1999  相似文献   

9.
An electrophoretically homogeneous preparation of mitochondrial NADP-dependent malate dehydrogenase with a specific activity of 155 u./mg and a 67% yield has been obtained, using ammonium sulfate fractionation, gel filtration through Toyopearl HW-55 F, ion-exchange chromatography on DEAE-Toyopearl 650 M and affinity chromatography on 2',5'-ADP-Sepharose 4B. The molecular mass of native malate dehydrogenase is 260 kD; Mr of the SDS-treated enzyme is 61 kD, which is suggestive of a tetrameric structure of the protein. Malate dehydrogenase is active only in the presence of Mg2+ or Mn2+, but not Ca2+ or Ba2+. The Km' values for Mn2+ and Mg2+ are 50 and 66 microM, respectively. At low malate concentrations and NADP saturation, the enzyme is characterized by a sigmoidal kinetics which changes to hyperbolic at low concentrations of NADP. The Lineweaver--Burk plots for the dependence of the initial reaction rate on the concentration of one substrate at several fixed concentrations of the other substrate intersect to the left of the B-axis. NADPH competes with NADP:pyruvate inhibits malate dehydrogenase ++noncompetitively with respect to the coenzyme. NADPH and pyruvate inhibit the malate dehydrogenase-catalyzed reaction via a mixed type mechanism with respect to malate. The data obtained are consistent with a consecutive mechanism of reaction, whose first substrate is NADP and the last product is NADPH.  相似文献   

10.
Cibacron Blue F3G-A, a probe used to monitor nucleotide binding domains in enzymes, inhibited sheep liver 5,10-methylenetetrahydrofolate reductase competitively with respect to 5-methyltetrahydrofolate and NADPH. TheK i values obtained by kinetic methods and theK d value for the binding of the dye to the enzyme estimated by protein fluorescence quenching were in the range 0.9–1.2 μM. Another triazine dye, Procion Red HE-3B interacted with the enzyme in an essentially similar manner to that observed with Cibacron Blue F3G-A. These results as well as the interaction of the dye with the enzyme monitored by difference spectroscopy and intrinsic protein fluorescence quenching methods indicated that the dye was probably interacting at the active site of the enzyme by binding at a hydrophobic region.  相似文献   

11.
Reimers, J. M., Huang, Q., Albe, K. R., and Wright, B. E. 1993. Purification and kinetic characterization of glucose-6-phosphate dehydrogenase from Dictyostelium discoideum. Experimental Mycology 17, 1-6. Glucose-6-phosphate dehydrogenase from Dictyostelium discoideum was purified 650-fold and kinetically characterized. The enzyme catalyzed the conversion of G6P + NADP to 6PG + NADPH stoichiometrically and irreversibly in vitro . The purified enzyme is specific for NADP. Michaelis constants for G6P and NADP were 0.040 and 0.011 mM, respectively. NADPH was found to be a competitive inhibitor with respect to NADP with a Ki of 0.006 mM and a noncompetitive inhibitor with respect to G6P. The data from initial velocity and product inhibition studies were consistent with a sequential mechanism.  相似文献   

12.
The F420-dependent NADP reductase of Methanobacterium thermoautotrophicum has been purified employing a combination of DEAE-cellulose ion-exchange chromatography, affinity chromatography with Blue Sepharose, Sephadex G-200 column chromatography and Red Sepharose affinity chromatography. The enzyme, which requires reduced F420 as an electron donor, has been purified over 3000-fold with a recovery of 65%. A molecular weight of 112000 was determined by Sephadex G-200 chromatography. A subunit molecular weight of 28 500 was determined by Sephadex G-200 chromatography. A subunit native enzyme is a tetramer. The optimal temperature for enzymatic activity was found to be 60°C with a pH optimum of 8.0. The NADP reductase had an apparent Km of 128 μMJ for reduced F420 and 40 μM for NADP. The enzyme was stable for at least 4 h at 65°C and pH 7.5. No loss of enzyme activity was detected when purified enzyme was stored aerobically in buffer containing 2-mercaptoethanol for 10 days at 4°C. Neither FMNH2 nor FADH2 could serve as electron donors; NAD was not utilized as electron acceptor.  相似文献   

13.
ω-Hydroxyfatty acid:NADP oxidoreductase, an enzyme involved in suberin biosynthesis, is induced by wounding potato tubers. Initial velocity and product inhibition studies with the purified enzyme suggested an ordered sequential mechanism, where NADPH is added first, followed by 16-oxohexadecanoate, and NADP is released after 16-hydroxyhexadecanoate. Substrate inhibition by NADPH was observed at concentrations higher than 0.2 mm. The inhibitory NADPH molecule competes with 16-oxohexadecanoate, indicating that it forms a dead-end complex with the E-NADPH form of the enzyme. The kinetics for the NADPH inhibition suggested that n > 1 in the rate equation v = V[NADPH](Km + [NADPH]+ [NADPH]n+1Ki); i.e., more than two NADPH molecules bind to enzyme. The Km for 16-oxohexadecanoate did not change from pH 7.5 to 9.0 but increased about 10-fold from pH 9.0 to 10.0, whereas the Km for NADPH and hexadecanal did not vary significantly in this pH range. Phenylglyoxal inactivated the enzyme; NADPH and AMP (which competes with NADPH; Ki = 1.1 mM) provided protection against such inactivation. Diethylpyrocarbonate also caused inactivation which was reversed by hydroxylamine; NADPH but not AMP protected the enzyme from this inhibition. Pyridoxal-5′-phosphate reversibly inactivated the enzyme and NaBH4 reduction of the pyridoxal phosphate-treated enzyme resulted in irreversible inhibition; a combination of NADPH and ω-oxo C16 acid provided protection against such inactivation. As the chain length of alkanals increased from C3 to C8, the Km for the substrate decreased drastically from 7000 to 90μm and a further increase in chain length from C8 to C20 resulted in only a small decrease in Km. The Km and V for 8-oxooctanoate and 10-oxodecanoate are compared with the values obtained for 16-oxohexadecanoate. Based on these results, it is proposed that arginine acts as the binding site for NADPH, a hydrophobic crevice with lysine at the bottom forms the binding site for 16-oxohexadecanoate and histidine participates in the reaction as the proton donor.  相似文献   

14.
The β-subunit of the voltage-sensitive K+ channels shares 15–30% amino acid identity with the sequences of aldo–keto reductases (AKR) genes. However, the AKR properties of the protein remain unknown. To begin to understand its oxidoreductase properties, we examine the pyridine coenzyme binding activity of the protein in vitro. The cDNA of Kvβ2.1 from rat brain was subcloned into a prokaryotic expression vector and overexpressed in Escherichia coli. The purified protein was tetrameric in solution as determined by size exclusion chromatography. The protein displayed high affinity binding to NADPH as determined by fluorometric titration. The KD values for NADPH of the full-length wild-type protein and the N-terminus deleted protein were 0.1±0.007 and 0.05±0.006 M, respectively — indicating that the cofactor binding domain is restricted to the C-terminus, and is not drastically affected by the absence of the N-terminus amino acids, which form the ball and chain regulating voltage-dependent inactivation of the α-subunit. The protein displayed poor affinity for other coenzymes and the corresponding values of the KD for NADH and NAD were between 1–3 μM whereas the KD for FAD was >10 μM. However, relatively high affinity binding was observed with 3-acetyl pyridine NADP, indicating selective recognition of the 2′ phosphate at the binding site. The selectivity of Kvβ2.1 for NADPH over NADP may be significant in regulating the K+ channels as a function of the cellular redox state.  相似文献   

15.
Inhibition of electron transport activities in the spheroplasts ofSynechococcus 6301 by HgCl2 is dependent on the concentration of mercury ions. The inhibition of whole chain electron transport activity occurs at low concentration of Hg2+ (6 ΜM@#@). This inhibition occurs mostly due to interaction of Hg2+ on plastocyanin. At an elevated concentration (24 ΜM@#@), mercury induces inhibition chiefly in photosystem II catalyzed electron transport. At this concentration it also alters both the absorption and emission characteristics of the phycocyanin. The photosystem I catalyzed electron transport was inhibited by 50% only at high concentrations (36 ΜM@#@) of HgCl2. However, electron transport catalyzed by photosystems I and II from reduced duroquinone to methylviologen which involves intersystem electron transport is extremely sensitive to mercury (low concentration 6–9 ΜM) like that of whole chain assay indicating that the observed inhibition in whole chain electron transport at low concentrations is mostly contributed by the damage involving other intersystem electron transport carrier(s) like plastocyanin. Thus mercury ions depending on the concentration affects the electron transport at multiple sites in the spheroplasts ofSynechococcus.  相似文献   

16.
An inducible pyridine nucleotide-linked cyclohexanol dehydrogenase activity was present in crude extracts from aNocardia species following growth on cyclohexane. The enzyme was purified 126-fold by affinity chromatography and has an oligomeric molecular weight of 145,000 ±5,000. There was an absolute requirement for NAD for activity and the products of the dehydrogenase reaction were stoichiometric amounts of NADH and cyclohexanone. The enzyme had a broad specificity for secondary alcohols including straight-chain secondary alcohols, cyclic and substituted cyclic alcohols, and cyclohexane diols. The apparentK m values for cyclohexanol and NAD were 3.7×10−5 M and 2.4×10−5 M, respectively, and the optimal pH for cyclohexanol oxidation was 10.5. The enzyme was heat sensitive, losing about 50% activity after a 1-min incubation at 45°C. Enzyme activity was completely inhibited by the thiol agent,p-chloromercuribenzoate but not by metal chelating agents.  相似文献   

17.
Lactate and malate dehydrogenases (EC 1.1.1.27 and EC 1.1.1.37, respectively) were precipitated with ammonium sulfate, redissolved in 100 mM phosphate buffer, and the kinetic parameters of each enzyme determined. Lactate dehydrogenase: The enzyme preparation had a specific activity of 0.35 μmole NADH oxidized/min/mg protein for pyruvate reduction, and 0.10 μmole NAD reduced/min/mg protein for lactate oxidation. Km values for the substrates and cofactors were as follows: pyruvate = 0.51, mM; lactate = 3.8 mM; NADH = 0.011 mM; and NAD = 0.17 mM. NADPH, NADP, or d(?)-lactate would not replace NADH, NAD, or l(+)-lactate, respectively. The enzyme was relatively stable at 50 C for 45 min, but much less stable at 60 C; repeated freezing and thawing of the enzyme preparation had little effect on LDH activity. Both p-chloromercuribenzoate (p-CMB) and N-ethylmaleimide (NEM) significantly inhibited LDH activity. Polyacrylamide gel electrophoresis demonstrated the presence of at least two LDH isoenzymes in the unpurified enzyme preparation. The molecular weight was estimated at 160,000 by gel chromatography. Malate dehydrogenase: The enzyme preparation had a specific activity of 6.70 μmole NADH oxidized/min/mg protein for oxaloacetate reduction, and 0.52 μmole NAD reduced/ min/mg protein for malate oxidation. Km values for substrates and cofactors were as follows: l-malate = 1.09 mM; oxaloacetate = 0.0059 mM; NADH = 0.017 mM; and NAD = 0.180 mM. NADP and NADPH would not replace NAD and NADH, respectively, d-malate was oxidized slowly when present in high concentrations (>100 mM). Significant substrate inhibition occurred with concentrations of l-malate and oxaloacetate above 40 mM and 0.5 mM, respectively. The enzyme was unstable at temperatures above 40 C, but repeated freezing and thawing of the enzyme preparation had little effect on MDH activity. Only p-CMB inhibited MDH activity. Polyacrylamide gel electrophoresis demonstrated the presence of at least three MDH isoenzymes in the unpurified enzyme preparation, and the molecular weight was estimated at 49,000 by gel chromatography.  相似文献   

18.
The present results show that the NADP specific isocitrate dehydrogenase from pig heart exhibits a time lag before the reaction rate approaches a constant value at low metal ion concentrations. Addition of NADPH or EDTA to the assay mixture abolished the lag, and will under certain conditions activate the enzyme.The lag time increased with increasing concentrations of isocitrate and decreased with increasing enzyme concentration. The NADP and metal ion concentration affected the lag in a complex manner. At low NADP and isocitrate concentration, the lag was reduced 50% by an NADPH concentration of less than 2 μm. Stopped flow experiments showed that premixing of NADP or NADPH with the enzyme abolished the effect of NADPH on the lag time. NADPH activated the enzyme at high NADP concentrations. This activating effect could be accounted for by removal of substrate inhibition by NADP.Evidence was obtained to show that the effect of NADPH on the activity was caused by binding of the reduced coenzyme to a site separate from the normal coenzyme binding site. Binding of metal ions by the reduced coenzyme is probably of importance as EDTA affects the lag time and activity in a manner similar to NADPH. The NADPH effect seems to be a general property of NADP-linked isocitrate dehydrogenases.  相似文献   

19.
Properties of partially purified NADP-malic enzyme (EC 1.1.1.40) from glumes of developing wheat grains were examined. The pH optimum for enzyme activity was influenced by malate and shifted from 7.3 to 7.6 when the concentration of malate was increased from 2 to 10 mM. The Km values, at pH 7.3, for various substrates were: malate, 0.76 mM; NADP, 20 μM and Mn2+, 0.06 mM. The requirement of Mn2+ cation for enzyme activity could be partially replaced by Mg2+ or Co2+. Mn2+ dependent enzyme activity was inhibited by Pb2+, Ni2+, Hg2+, Zn2+, Cd2+, Al3+ and Fe3+. During the reaction, substrate molecules (malate and NADP) reacted with enzyme sequentially. Activity of malic enzyme was inhibited by products of the reaction viz pyruvate, HCO3? and NADPH2. At a limiting fixed concentration of NADP, these products induced a positive cooperative response to increasing concentrations of malate.  相似文献   

20.
《Phytochemistry》1987,26(10):2665-2670
The bifunctional enzyme dehydroquinase (DHQase, EC 4.2.1.10)-shikimate: NADP oxidoreductase (SHORase, EC 1.1.1.25) has been purified 6500-fold to homogeneity from Pisum sativum shoot tissue. A rapid purification procedure using high performance liquid chromatography was used to isolate the enzyme from chloroplast preparations. The purified enzyme is monomeric with Mr 59 000. Chromatofocusing separates three isoenzymes, two of which are chloroplastic. DHQase and SHORase (forward reaction) show pH optima at pH 7 and apparent Km values of 2.7 x 10−5 M (dehydroquinate), 2.1 x 10−4 M (dehydroshikimate) and 1.5 x 10−5 M (NADPH). Chloride is a competitive inhibitor of DHQase. The SHORase reaction has an ordered (sequential) kinetic mechanism and is unaffected by the presence of DHQ.  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