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 20 毫秒
1.
Assay conditions for the 11beta-hydroxysteroid dehydrogenase have been optimized by adding phospholipids in the media buffer to increase and stabilize the enzymatic activity. The presence of phospholipids greatly facilitates the study of the binding of cortisone and NADPH at the enzyme catalytic site. Kinetic analyses conducted with the human and rabbit enzyme isoforms suggest that both enzymes behave according to an ordered sequential bi-bi mechanism where the NADPH is the first to bind at the active site followed by cortisone. The equilibrium dissociation constant, K(i)a as well as the apparent Michaelis-Menten constants K(m)a, K(m)b, k(cat)a, and k(cat)b for NADPH and cortisone, have been determined to be 147.5 microM, 14.4 microM, 43.8 nM, 0.21 min(-1), and 0.27 min(-1), respectively, for the human enzyme and 41.1 microM, 3.1 microM, 161.7 nM, 0.49 min(-1), and 0.52min(-1), respectively, for the rabbit enzyme.  相似文献   

2.
The kinetic properties of placental glucose-6-phosphate dehydrogenase were studied, since this enzyme is expected to be an important component of the placental protection system. In this capacity it is also very important for the health of the fetus. The placental enzyme obeyed "Rapid Equilibrium Ordered Bi Bi" sequential kinetics with K(m) values of 40+/-8 microM for glucose-6-phosphate and 20+/-10 microM for NADP. Glucose-6-phosphate, 2-deoxyglucose-6-phosphate and galactose-6-phosphate were used with catalytic efficiencies (k(cat)/K(m)) of 7.4 x 10(6), 4.89 x 10(4) and 1.57 x 10(4) M(-1).s(-1), respectively. The K(m)app values for galactose-6-phosphate and for 2-deoxyglucose-6-phosphate were 10+/-2 and 0.87+/-0.06 mM. With galactose-6-phosphate as substrate, the same K(m) value for NADP as glucose-6-phosphate was obtained and it was independent of galactose-6-phosphate concentration. On the other hand, when 2-deoxyglucose-6-phosphate used as substrate, the K(m) for NADP decreased from 30+/-6 to 10+/-2 microM as the substrate concentration was increased from 0.3 to 1.5 mM. Deamino-NADP, but not NAD, was a coenzyme for placental glucose-6-phosphate dehydrogenase. The catalytic efficiencies of NADP and deamino-NADP (glucose-6-phosphate as substrate) were 1.48 x 10(7) and 4.80 x 10(6) M(-1)s(-1), respectively. With both coenzymes, a hyperbolic saturation and an inhibition above 300 microM coenzyme concentration, was observed. Human placental glucose-6-phosphate dehydrogenase was inhibited competitively by 2,3-diphosphoglycerate (K(i)=15+/-3 mM) and NADPH (K(i)=17.1+/-3.2 microM). The small dissociation constant for the G6PD:NADPH complex pointed to tight enzyme:NADPH binding and the important role of NADPH in the regulation of the pentose phosphate pathway.  相似文献   

3.
Pollock VV  Barber MJ 《Biochemistry》2001,40(5):1430-1440
Rhodobacter sphaeroides f. sp. denitrificans biotin sulfoxide reductase catalyzes the reduction of d-biotin d-sulfoxide (BSO) to biotin. Initial rate studies of the homogeneous recombinant enzyme, expressed in Escherichia coli, have demonstrated that the purified protein utilizes NADPH as a facile electron donor in the absence of any additional auxiliary proteins. We have previously shown [Pollock, V. V., and Barber, M. J. (1997) J. Biol. Chem. 272, 3355-3362] that, at pH 8 and in the presence of saturating concentrations of BSO, the enzyme exhibits, a marked preference for NADPH (k(cat,app) = 500 s(-1), K(m,app) = 269 microM, and k(cat,app)/K(m,app) = 1.86 x 10(6) M(-1) s(-1)) compared to NADH (k(cat,app) = 47 s(-1), K(m,app) = 394 microM, and k(cat,app)/K(m,app) = 1.19 x 10(5) M(-1) s(-1)). Production of biotin using NADPH as the electron donor was confirmed by both the disk biological assay and by reversed-phase HPLC analysis of the reaction products. The purified enzyme also utilized ferricyanide as an artificial electron acceptor, which effectively suppressed biotin sulfoxide reduction and biotin formation. Analysis of the enzyme isolated from tungsten-grown cells yielded decreased reduced methyl viologen:BSO reductase, NADPH:BSO reductase, and NADPH:FR activities, confirming that Mo is required for all activities. Kinetic analyses of substrate inhibition profiles revealed that the enzyme followed a Ping Pong Bi-Bi mechanism with both NADPH and BSO exhibiting double competitive substrate inhibition. Replots of the 1/v-axes intercepts of the parallel asymptotes obtained at several low concentrations of fixed substrate yielded a K(m) for BSO of 714 and 65 microM for NADPH. In contrast, utilizing NADH as an electron donor, the replots yielded a K(m) for BSO of 132 microM and 1.25 mM for NADH. Slope replots of data obtained at high concentrations of BSO yielded a K(i) for BSO of 6.10 mM and 900 microM for NADPH. Kinetic isotope studies utilizing stereospecifically deuterated NADPD indicated that BSO reductase uses specifically the 4R-hydrogen of the nicotinamide ring. Cyanide inhibited NADPH:BSO and NADPH:FR activities in a reversible manner while diethylpyrocarbonate treatment resulted in complete irreversible inactivation of the enzyme concomitant with molybdenum cofactor release, indicating that histidine residues are involved in cofactor-binding.  相似文献   

4.
Ge L  Seah SY 《Journal of bacteriology》2006,188(20):7205-7210
Pseudomonas aeruginosa is an opportunistic pathogen that produces the siderophore pyoverdine, which enables it to acquire the essential nutrient iron from its host. Formation of the iron-chelating hydroxamate functional group in pyoverdine requires the enzyme PvdA, a flavin-dependent monooxygenase that catalyzes the N(5) hydroxylation of l-ornithine. pvdA from P. aeruginosa was successfully overexpressed in Escherichia coli, and the enzyme was purified for the first time. The enzyme possessed its maximum activity at pH 8.0. In the absence of l-ornithine, PvdA has an NADPH oxidase activity of 0.24 +/- 0.02 micromol min(-1) mg(-1). The substrate l-ornithine stimulated this activity by a factor of 5, and the reaction was tightly coupled to the formation of hydroxylamine. The enzyme is specific for NADPH and flavin adenine dinucleotide (FAD(+)) as cofactors, as it cannot utilize NADH and flavin mononucleotide. By fluorescence titration, the dissociation constants for NADPH and FAD(+) were determined to be 105.6 +/- 6.0 microM and 9.9 +/- 0.3 microM, respectively. Steady-state kinetic analysis showed that the l-ornithine-dependent NADPH oxidation obeyed Michaelis-Menten kinetics with apparent K(m) and V(max) values of 0.58 mM and 1.34 micromol min(-1) mg(-1). l-Lysine was a nonsubstrate effector that stimulated NADPH oxidation, but uncoupling occurred and hydrogen peroxide instead of hydroxylated l-lysine was produced. l-2,4-Diaminobutyrate, l-homoserine, and 5-aminopentanoic acid were not substrates or effectors, but they were competitive inhibitors of the l-ornithine-dependent NADPH oxidation reaction, with K(ic)s of 3 to 8 mM. The results indicate that the chemical nature of effectors is important for simulation of the NADPH oxidation rate in PvdA.  相似文献   

5.
In plant, the first and the third steps of the synthesis of methionine and threonine are catalyzed by a bifunctional enzyme, aspartate kinase-homoserine dehydrogenase (AK-HSDH). In this study, we report the first purification and characterization of a highly active threonine-sensitive AK-HSDH from plants (Arabidopsis thaliana). The specific activities corresponding to the forward reaction of AK and reverse reaction of HSDH of AK-HSDH were 5.4 micromol of aspartyl phosphate produced min(-1) mg(-1) of protein and 18.8 micromol of NADPH formed min(-1) mg(-1) of protein, respectively. These values are 200-fold higher than those reported previously for partially purified plant enzymes. AK-HSDH exhibited hyperbolic kinetics for aspartate, ATP, homoserine, and NADP with K(M) values of 11.6 mM, 5.5 mM, 5.2 mM, and 166 microM, respectively. Threonine was found to inhibit both AK and HSDH activities by decreasing the affinity of the enzyme for its substrates and cofactors. In the absence of threonine, AK-HSDH behaved as an oligomer of 470 kDa. Addition of the effector converted the enzyme into a tetrameric form of 320 kDa.  相似文献   

6.
An arylketone monooxygenase was purified from Pseudomonas putida JD1 by ion exchange and affinity chromatography. It had the characteristics of a Baeyer-Villiger-type monooxygenase and converted its substrate, 4-hydroxyacetophenone, into 4-hydroxyphenyl acetate with the consumption of one molecule of oxygen and oxidation of one molecule of NADPH per molecule of substrate. The enzyme was a monomer with an M(r) of about 70,000 and contained one molecule of flavin adenine dinucleotide (FAD). The enzyme was specific for NADPH as the electron donor, and spectral studies showed rapid reduction of the FAD by NADPH but not by NADH. Other arylketones were substrates, including acetophenone and 4-hydroxypropiophenone, which were converted into phenyl acetate and 4-hydroxyphenyl propionate, respectively. The enzyme displayed Michaelis-Menten kinetics with apparent K(m) values of 47 microM for 4-hydroxyacetophenone, 384 microM for acetophenone, and 23 microM for 4-hydroxypropiophenone. The apparent K(m) value for NADPH with 4-hydroxyacetophenone as substrate was 17.5 microM. The N-terminal sequence did not show any similarity to other proteins, but an internal sequence was very similar to part of the proposed NADPH binding site in the Baeyer-Villiger monooxygenase cyclohexanone monooxygenase from an Acinetobacter sp.  相似文献   

7.
We overexpressed and purified from Escherichia coli the dihydrofolate reductase (DHFR) of the gammaherpesviruses human herpesvirus 8 (HHV-8), herpesvirus saimiri (HVS), and rhesus rhadinovirus (RRV). All three enzymes proved catalytically active. The K(m) value of HHV-8 DHFR for dihydrofolate (DHF) was 2.02+/-0.44 microM, that of HVS DHFR was 4.31+/-0.56 microM, and that of RRV DHFR is 7.09+/-0.11 microM. These values are approximately 5-15-fold higher than the K(m) value reported for the human DHFR. The K(m) value of HHV-8 DHFR for NADPH was 1.31+/-0.23 microM, that of HVS DHFR was 3.78+/-0.61 microM, and that of RRV DHFR was 7.47+/-0.59 microM. These values are similar or slightly higher than the corresponding K(m) value of the human enzyme. Methotrexate, aminopterin, trimethoprim, pyrimethamine, and N(alpha)-(4-amino-4-deoxypteroyl)-N(delta)-hemiphthaloyl-L-ornithine (PT523), all well-known folate antagonists, inhibited the DHFR activity of the three gammaherpesviruses competitively with respect to DHF but proved markedly less inhibitory to the viral than towards the human enzyme.  相似文献   

8.
Glucose-6-phosphate dehydrogenase (G6PD) catalyses the first step of the pentose phosphate pathway which generates NADPH for anabolic pathways and protection systems in liver. G6PD was purified from dog liver with a specific activity of 130 U x mg(-1) and a yield of 18%. PAGE showed two bands on protein staining; only the slower moving band had G6PD activity. The observation of one band on SDS/PAGE with M(r) of 52.5 kDa suggested the faster moving band on native protein staining was the monomeric form of the enzyme.Dog liver G6PD had a pH optimum of 7.8. The activation energy, activation enthalpy, and Q10, for the enzymatic reaction were calculated to be 8.96, 8.34 kcal x mol(-1), and 1.62, respectively.The enzyme obeyed "Rapid Equilibrium Random Bi Bi" kinetic model with Km values of 122 +/- 18 microM for glucose-6-phosphate (G6P) and 10 +/- 1 microM for NADP. G6P and 2-deoxyglucose-6-phosphate were used with catalytic efficiencies (kcat/Km) of 1.86 x 10(6) and 5.55 x 10(6) M(-1) x s(-1), respectively. The intrinsic Km value for 2-deoxyglucose-6-phosphate was 24 +/- 4mM. Deamino-NADP (d-NADP) could replace NADP as coenzyme. With G6P as cosubstrate, Km d-ANADP was 23 +/- 3mM; Km for G6P remained the same as with NADP as coenzyme (122 +/- 18 microM). The catalytic efficiencies of NADP and d-ANADP (G6P as substrate) were 2.28 x 10(7) and 6.76 x 10(6) M(-1) x s(-1), respectively. Dog liver G6PD was inhibited competitively by NADPH (K(i)=12.0 +/- 7.0 microM). Low K(i) indicates tight enzyme:NADPH binding and the importance of NADPH in the regulation of the pentose phosphate pathway.  相似文献   

9.
The gene fprA of Mycobacterium tuberculosis, encoding a putative protein with 40% identity to mammalian adrenodoxin reductase, was expressed in Escherichia coli and the protein purified to homogeneity. The 50-kDa protein monomer contained one tightly bound FAD, whose fluorescence was fully quenched. FprA showed a low ferric reductase activity, whereas it was very active as a NAD(P)H diaphorase with dyes. Kinetic parameters were determined and the specificity constant (kcat/Km) for NADPH was two orders of magnitude larger than that of NADH. Enzyme full reduction, under anaerobiosis, could be achieved with a stoichiometric amount of either dithionite or NADH, but not with even large excess of NADPH. In enzyme titration with substoichiometric amounts of NADPH, only charge transfer species (FAD-NADPH and FADH2-NADP+) were formed. At NADPH/FAD ratios higher than one, the neutral FAD semiquinone accumulated, implying that the semiquinone was stabilized by NADPH binding. Stabilization of the one-electron reduced form of the enzyme may be instrumental for the physiological role of this mycobacterial flavoprotein. By several approaches, FprA was shown to be able to interact productively with [2Fe-2S] iron-sulfur proteins, either adrenodoxin or plant ferredoxin. More interestingly, kinetic parameters of the cytochrome c reductase reaction catalyzed by FprA in the presence of a 7Fe ferredoxin purified from M. smegmatis were determined. A Km value of 30 nm and a specificity constant of 110 microM(-1) x s(-1) (10 times greater than that for the 2Fe ferredoxin) were determined for this ferredoxin. The systematic name for FprA is therefore NADPH-ferredoxin oxidoreductase.  相似文献   

10.
Microsomal preparations from cell suspension cultures of the Indian plant Rauvolfia serpentina catalyze the hydroxylation of deoxysarpagine under formation of sarpagine. The newly discovered enzyme is dependent on NADPH and oxygen. It can be inhibited by typical cytochrome P450 inhibitors such as cytochrome c, ketoconazole, metyrapone, tetcyclacis and carbon monoxide. The CO-effect is reversible with light (450 nm). The data indicate that deoxysarpagine hydroxylase is a novel cytochrome P450-dependent monooxygenase. A pH optimum of 8.0 and a temperature optimum of 35 degrees C were determined. K(m) values were 25 microM for NADPH and 7.4 microM for deoxysarpagine. Deoxysarpagine hydroxylase activity was stable in presence of 20% sucrose at -25 degrees C for >3 months. The analysis of presence of the hydroxylase in nine cell cultures of seven different families indicates a very limited taxonomic distribution of this enzyme.  相似文献   

11.
Q X Chen  C D Nancarrow  F Sweet 《Steroids》1987,49(6):477-496
3 beta,20 alpha-Hydroxysteroid oxidoreductase has been isolated from ovine fetal blood by a 2,370-fold purification scheme of ammonium sulfate fractionation, calcium phosphate gel adsorption, affinity chromatography, and fast performance liquid chromatography. A new high performance liquid chromatography-based assay for measuring 20 alpha-reductase activity is described. The enzyme is a monomer with a molecular weight of 35,000 and uses NADPH as a cofactor for reductase activity. It reduces progesterone to 4-pregnen-20 alpha-ol-3-one or 5 alpha-dihydrotestosterone to 5 alpha-androstan-3 beta,17 beta-diol with kinetic characteristics of Km = 30.8 microM and Vmax = 0.7 nmol min-1 (nmol of enzyme)-1 or Km = 74 microM and Vmax = 1.3 nmol min-1 (nmol of enzyme)-1, respectively. 5 alpha-Dihydrotestosterone competitively inhibits 20 alpha-reductase activity with a Ki value of 102 microM.  相似文献   

12.
Affinity labeling studies of NADP(+)-glutamate dehydrogenase from Salmonella typhimurium have shown that the peptide Leu-282-Lys-286 is located near the coenzyme site [Haeffner-Gormley et al. (1991) J. Biol. Chem. 266, 5388-5394]. The present study was undertaken to evaluate the role of lysine-286. The mutant enzymes K286R, K286Q, and K286E were prepared by site-directed mutagenesis, expressed in Escherichia coli, and purified. The Vmax values (micromoles of NADPH per minute per milligram of protein) were similar for WT (270), K286R (529), K296Q (409), and K286E (382) enzymes. As measured at pH 7.9, the Km value for NADPH was much greater for K286E (280 microM) than for WT (9.8 microM), K286R (30 microM), or K286Q (66 microM) enzymes. The efficiencies (kcat/Km) of the WT and K286R mutant were similar (1.2 x 10(3) min-1 microM-1 and 1.0 x 10(3) min-1 microM-1, respectively) while those of K286Q (0.30 x 10(3) min-1 microM-1) and K286E (0.07 x 10(3) min-1 microM-1) were greatly reduced. The decreased efficiency of the K286E mutant results from the increase in Km-NADPH, consistent with a role for a basic residue at position 286 which enhances the binding of NADPH. Plots of Vmax vs pH showed the pH optima to be 8.1-8.3 for all enzymes at saturating NADPH concentrations. A 40-fold increase in Km-NADPH for K286E was observed as the pH increased from 5.98 to 8.08, from which a unique pKe of 6.5 was calculated.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

13.
Crude soluble extracts of Corynebacterium cyclohexanicum, grown on cyclohexanecarboxylic acid, were found to contain 4-hydroxybenzoate 3-hydroxylase which functions with NADH as well as NADPH. The purified enzyme preparation was electrophoretically homogeneous and contained FAD as prosthetic group. The relative molecular mass of the enzyme was estimated to be about 47000 by native and denaturated acrylamide gel electrophoresis, indicating that it is monomeric. The enzyme was stable at 60 degrees C for 10 min. The enzyme was highly specific for p-hydroxybenzoate. The activity was inhibited by several aromatic analogues of p-hydroxybenzoate such as p-aminobenzoate, p-fluorobenzoate, o-hydroxybenzoate, m-hydroxybenzoate, 2,4-dihydroxygenzoate, and 2,5-dihydroxybenzoate. The Km value for NADH was fairly constant, about 45 microM, in the pH range 7.0-8.4, whereas the Km value for NADPH increased from 63 microM to 170 microM as the pH rose from 7.0 to 8.4. V values in the same pH range, however, were approximately constant in both cases; about 30 mumol min-1 mg-1 for NADH, and 26 mumol min-1 mg-1 for NADPH. Mg2+ was required for full activity of the enzyme in low concentrations of phosphate buffer. The enzyme was inhibited by C1- which was non-competitive with respect to NADH, NADPH and p-hydroxybenzoate.  相似文献   

14.
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   

15.
The 3-hydroxypropionate cycle is a new autotrophic CO(2) fixation pathway in Chloroflexus aurantiacus and some archaebacteria. The initial step is acetyl-coenzyme A (CoA) carboxylation to malonyl-CoA by acetyl-CoA carboxylase, followed by NADPH-dependent reduction of malonyl-CoA to 3-hydroxypropionate. This reduction step was studied in Chloroflexus aurantiacus. A new enzyme was purified, malonyl-CoA reductase, which catalyzed the two-step reduction malonyl-CoA + NADPH + H(+) --> malonate semialdehyde + NADP(+) + CoA and malonate semialdehyde + NADPH + H(+) --> 3-hydroxypropionate + NADP(+). The bifunctional enzyme (aldehyde dehydrogenase and alcohol dehydrogenase) had a native molecular mass of 300 kDa and consisted of a single large subunit of 145 kDa, suggesting an alpha(2) composition. The N-terminal amino acid sequence was determined, and the incomplete gene was identified in the genome database. Obviously, the enzyme consists of an N-terminal short-chain alcohol dehydrogenase domain and a C-terminal aldehyde dehydrogenase domain. No indication of the presence of a prosthetic group was obtained; Mg(2+) and Fe(2+) stimulated and EDTA inhibited activity. The enzyme was highly specific for its substrates, with apparent K(m) values of 30 microM malonyl-CoA and 25 microM NADPH and a turnover number of 25 s(-1) subunit(-1). The specific activity in autotrophically grown cells was 0.08 micromol of malonyl-CoA reduced min(-1) (mg of protein)(-1), compared to 0.03 micromol min(-1) (mg of protein)(-1) in heterotrophically grown cells, indicating downregulation under heterotrophic conditions. Malonyl-CoA reductase is not required in any other known pathway and therefore can be taken as a characteristic enzyme of the 3-hydroxypropionate cycle. Furthermore, the enzyme may be useful for production of 3-hydroxypropionate and for a coupled spectrophotometric assay for activity screening of acetyl-CoA carboxylase, a target enzyme of potent herbicides.  相似文献   

16.
The effects of benzyl (BITC) and phenethyl isothiocyanate (PEITC) on the activity of a P450 2E1 mutant where the conserved threonine at position 303 was replaced with an alanine residue (P450 2E1 T303A) were examined. PEITC inactivated the mutant enzyme with a K(I) of 1.6 microM. PEITC also inactivated the wild-type P450 2E1 as efficiently with a K(I) of 2.7 microM. The inactivation was entirely dependent on NADPH and followed pseudo-first-order kinetics. Previously we reported the mechanism-based inactivation of wild-type P450 2E1 by BITC with a K(I) of 13 microM. In contrast to the wild-type enzyme, the P450 2E1 T303A mutant was not inactivated by BITC but it was inhibited in a competitive manner with a K(i) of 3 microM. The binding constants determined by spectral binding studies were similar for both enzymes. The binding of BITC produced characteristic Type I spectral changes in the wild-type and mutant enzyme. A radiolabeled BITC metabolite bound to P450 2E1 and to P450 2E1 T303A when both enzymes were incubated with [(14)C]BITC and NADPH. Whole protein electrospray ion trap mass spectrometry indicated that a mass consistent with one molecule of benzylisocyanate and oxygen was adducted to the wild-type enzyme. The mass adducted to the T303A mutant was consistent with the addition of one hydroxylated BITC or of one benzylisocyanate moiety and one sulfur molecule. Analysis of the metabolites of BITC indicated that each enzyme produced similar metabolites but that the mutant enzyme generated significantly higher amounts of benzaldehyde and benzoic acid when compared to the wild-type enzyme.  相似文献   

17.
J L Gabriel  G W Plaut 《Biochemistry》1984,23(12):2773-2778
The activity of NAD-dependent isocitrate dehydrogenase from bovine heart was inhibited by NADH (apparent Ki about 4.3 microM) and NADPH (Ki about 9.8 microM) at subsaturating substrate concentrations at pH 7.4. The inhibition by NADH or NADPH was reversed competitively by magnesium isocitrate in the presence of ADP, but not without ADP. Reversal of inhibition by NADH or NADPH with respect to NAD+ was competitive or of the linear mixed type depending on whether ADP was absent or present. ADP3- (0.2 mM) increased the Ki(app) for NADPH from 9.8 to 27.1 microM; further addition of Ca2+ (0.2 mM) raised the Ki(app) to 127 microM. For the modification of NADPH inhibition by ADP, S0.5 for Ca2+ was approximately 48 microM. This compares to the Km for Ca2+ of 0.3-1 microM for the activation of the enzyme without NADPH [Denton, R. M., Richards, D. A., & Chin, J. G. (1978) Biochem. J. 176, 899-906; Aogaichi, T., Evans, J., Gabriel, J., & Plaut, G. W. E. (1980) Arch. Biochem. Biophys. 204, 350-360]. ADP did not affect the Ki for NADH. Magnesium citrate, which was about 100-fold more effective as a positive modifier of the enzyme with ADP than without ADP [Gabriel, J. L., & Plaut, G. W. E. (1983) Fed. Proc., Fed. Am. Soc. Exp. Biol. 42, 2082], reversed competitively the inhibition by NADPH in the presence of ADP, but not without ADP. Magnesium citrate did not reverse NADH inhibition.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

18.
Dihydrofolate reductase (DHFR) from extracts of Mycobacterium smegmatis strain mc2(6) and trimethoprim-resistant mutant mc2(26) was purified to homogeneity. In crude extracts, the specific activity of the enzyme from the trimethoprim resistant strain was comparable to that from the sensitive strain. The DHFR from both sources was purified using affinity chromatography on MTX-Sepharose followed by Mono Q FPLC. The enzyme has an apparent molecular mass of 23 kDa from gel filtration on Sephadex G-100 and from SDS-PAGE. Amino terminal sequence analysis showed homology with DHFRs from a subset of other gram-positive organisms. The purified enzyme from the trimethoprim-sensitive organism exhibited Km values for H2folate and NADPH of 0.68 +/- 0.2 microM and 21 +/- 4 microM, respectively. The Km values for H2folate and NADPH for the enzyme from the drug-resistant organism were 1.8 +/- 0.4 microM and 5.3 +/- 1.5 microM, respectively. A kcat of 4.5 sec-1 was determined for the DHFR from both sources. The enzyme from both sources was competitively inhibited by pyrimethamine and trimethoprim. The Ki value of trimethoprim, for the enzyme from the drug-resistant organism was about six-fold higher than for the enzyme from drug-sensitive strain. Our data suggest that mutation of DHFR contributes to trimethoprim resistance in the mc2(26) strain of M. smegmatis.  相似文献   

19.
An NADPH-dependent membrane-bound flavoprotein dehydrogenase, assayed as a catalyst of electron transfer from NADPH to cytochrome c, was extracted from membranes of rabbit peritoneal neutrophils with Triton X-100 and sodium deoxycholate in the presence of diisopropylfluorophosphate as antiprotease, and purified to electrophoretic homogeneity. The purified enzyme in detergent was able to enhance the rate of formation of the superoxide anion O2- in a cell-free system, consisting of membrane and cytosolic fractions from resting neutrophils complemented with arachidonic acid, guanosine 5'-[gamma- thio]triphosphate and Mg2+. This suggested that the NADPH dehydrogenase was a component of the rabbit neutrophil oxidase complex. The purification factor of the enzyme with respect to the membrane fraction was close to 1000 and the recovery of activity was 33%. FMN and FAD were associated with the enzyme in a molar ratio close to 1. On SDS/PAGE, the enzyme migrated with a molecular mass of 77 kDa. A similar mass was determined by filtration on a molecular sieve. The isoelectric point of this enzyme was 4.7 +/- 0.1. Its activity was maximal between pH 7.5 and pH 8.5, and depended on the ionic strength of the medium, with a maximum at an ionic strength of 0.5. Reduction of cytochrome c by NADPH obeyed Michaelis-Menten kinetics with a KM value of 15 microM for cytochrome c. When NADPH was the variable substrate, a KM value of 1.9 microM for NADPH was found, but a significant deviation from Michaelis-Menten kinetics was observed at high concentrations of NADPH. Mersalyl strongly inhibited the reductase activity when added to the enzyme prior to NADPH; preincubation of the enzyme with NADPH considerably reduced the inhibitory efficiency of mersalyl. A partially proteolyzed water-soluble, active, form of enzyme with a molecular mass of 67 kDa was prepared. The proteolyzed enzyme exhibited the same specificity, and kinetic behavior with respect to NADPH, and the same dependency on the ionic strength, as the native enzyme.  相似文献   

20.
2-C-Methyl-D-erythritol-4-phosphate synthase (MEP synthase) catalyzes the rearrangement/reduction of 1-D-deoxyxylulose-5-phosphate (DXP) to methylerythritol-4-phosphate (MEP) as the first pathway-specific reaction in the MEP biosynthetic pathway to isoprenoids. Recombinant E. coli MEP was purified by chromatography on DE-52 and phenyl-Sepharose, and its steady-state kinetic constants were determined: k(cat) = 116 +/- 8 s(-1), K(M)(DXP) = 115 +/- 25 microM, and K(M)(NADPH) = 0.5 +/- 0.2 microM. The rearrangement/reduction is reversible; K(eq) = 45 +/- 6 for DXP and MEP at 150 microM NADPH. The mechanism for substrate binding was examined using fosmidomycin and dihydro-NADPH as dead-end inhibitors. Dihydro-NADPH gave a competitive pattern against NADPH and a noncompetitive pattern against DXP. Fosmidomycin was an uncompetitive inhibitor against NADPH and gave a pattern representative of slow, tight-binding competitive inhibition against DXP. These results are consistent with an ordered mechanism where NADPH binds before DXP.  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