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.
Gupta VK  Singh R 《Plant physiology》1988,87(3):741-744
NADP+-isocitrate dehydrogenase (threo-DS-isocitrate: NADP+ oxidoreductase [decarboxylating]; EC 1.1.1.42) (IDH) from pod walls of chickpea (Cicer arietinum L.) was purified 192-fold using ammonium sulfate fractionation, ion exchange chromatography on DEAE-Sephadex A-50, and gel filtration through Sephadex G-200. The purified enzyme, having a molecular weight of about 126,000, exhibited a broad pH optima from 8.0 to 8.6. It was quite stable at 4°C and had an absolute requirement for a divalent cation, either Mg2+ or Mn2+, for its activity. Typical hyperbolic kinetics was obtained with increasing concentrations of NADP+, dl-isocitrate, Mn2+, and Mg2+. Their Km values were 15, 110, 15, and 192 micromolar, respectively. The enzyme activity was inhibited by sulfhydryl reagents. Various amino acids, amides, organic acids, nucleotides, each at a concentration of 5 millimolar, had no effect on the activity of the enzyme. The activity was not influenced by adenylate energy charge but decreased linearly with increasing ratio of NADPH to NADP+. Initial velocity studies indicated kinetic mechanism to be sequential. NADPH inhibited the forward reaction competitively with respect to NADP+ at fixed saturating concentration of isocitrate, whereas 2-oxoglutarate inhibited the enzyme noncompetitively at saturating concentrations of both NADP+ and isocitrate, indicating the reaction mechanism to be random sequential. Results suggest that the activity of NADP+-IDH in situ is likely to be controlled by intracellular NADPH to NADP+ ratio as well as by the concentration of various substrates and products.  相似文献   

2.
The subcellular location of NADP+-isocitrate dehydrogenase was investigated by preparing protoplasts from leaves of pea seedlings. Washed protoplasts were gently lysed and the whole lysate separated on sucrose gradients by a rate-zonal centrifugation. Organelles were located by marker enzymes and chlorophyll analysis. Most of the NADP+-isocitrate dehydrogenase was in the soluble fraction. About 10% of the NADP+-isocitrate dehydrogenase was present in the chloroplasts as a partially latent enzyme. Less than 1% of the activity was found associated with the peroxisome fraction. NADP+-isocitrate dehydrogenase was partially characterized from highly purified chloroplasts isolated from shoot homogenates. The enzyme exhibited apparent Km values of 11 micromolar (NADP+), 35 micromolar (isocitrate), 78 micromolar (Mn2+), 0.3 millimolar (Mg2+) and showed optimum activity at pH 8 to 8.5 with Mn2+ and 8.8 to 9.2 with Mg2+. The NADP+-isocitrate dehydrogenase activity previously claimed in the peroxisomes by other workers is probably due to isolation procedures and/or nonspecific association. The NADP+-isocitrate dehydrogenase activity in the chloroplasts might help supply α-ketoglutarate for glutamate synthase action.  相似文献   

3.
The steady state kinetics of glyceraldehyde 3-phosphate:NADP+ oxidoreductase (GNR) (EC 1.2.1.9) have been investigated. The enzyme exhibits hyperbolic behavior over a wide range of substrate concentrations. Double-reciprocal plots are nearly parallel or distantly convergent with limiting Km values of 2 to 5 micromolar for NADP+ and 20 to 40 micromolar for D-glyceraldehyde 3-phosphate (G3P). The velocity response to NADP+ as the varied substrate is however sigmoidal if G3P concentration exceeds 10 micromolar, whereas the response to G3P may show inhibition above this concentration. This `G3P-inhibited state' is alleviated by saturating amounts of NADP+ or NADPH. Product inhibition patterns indicate NADPH as a potent competitive inhibitor to NADP+ (Ki 30 micromolar) and mixed inhibitor towards G3P, and 3-phosphoglycerate (3PGA) as mixed inhibitor to both NADP+ and G3P (Ki 10 millimolar). The data, and those obtained with dead-end inhibitors, are consistent with a nonrapid equilibrium random mechanism with two alternative kinetic pathways. Of these, a rapid kinetic sequence (probably ordered with NADP+ binding first and G3P binding as second substrate) is dominant in the range of hyperbolic responses. A reverse reaction with 3PGA and NADPH as substrates is unlikely, and was not detected. Of a number of compounds tested, erythrose 4-phosphate (Ki 7 micromolar) and Pi (Ki 2.4 millimolar) act as competitive inhibitors to G3P (uncompetitive towards NADP+) and are likely to affect the in vivo activity. Ribose 5-phosphate, phosphoenolpyruvate, ATP, and ADP are also somewhat inhibitory. Full GNR activity in the leaf seems to be allowed only under high photosynthesis conditions, when levels of several inhibitors are low and substrate is high. We suggest that a main function of leaf GNR is to supply NADPH required for photorespiration, the reaction product 3PGA being cycled back to chloroplasts.  相似文献   

4.
The kinetics of a Mn2+-requiring, NADP+-specific isocitrate dehydrogenase from Salmonella typhimurium have been examined by the measurement of initial velocity rates in the presence and absence of the reaction products. The binding of each of the cosubstrates, isocitrate, and NADP+, is not independent of the other, and the isocitrate-Mn2+ complex is the kinetically important substrate species. All of the reaction products, α-ketoglutarate, CO2, and NADPH are competitive with both cosubstrates and the mechanism appears to be of the rapid equilibrium random type. The enzyme has been purified to homogeneity and has an isoelectric point at pH 4.0–4.2, and an apparent molecular weight of 102,000.  相似文献   

5.
The kinetics of the NADP+- and phosphate-dependent oxidation of glutamic acid 5-semialdehyde are consistent with a rapid-equilibrium random order mechanism. The Km for dl-pyrroline-5-carboxylic acid is 2.5 mM, for NADP+ is 0.05 mM and for phosphate is 0.35 mM. TheVmax is approx. 8.0 units per mg protein. The reaction is highly specific for the dl-pyrroline-5-carboxylic acid and NADP+, but a number of divalent anions can substitute for phosphate. NADPH is competitive with respect to all three substrates and an analog of γ-glutamyl phosphate, 3-(phosphonoacetylamido)-l-alanine, is competitive with respect to dl-pyrroline-5-carboxylic acid and non-competitive with respect to NADP+ and phosphate, suggesting dead-end complex formation.  相似文献   

6.
NADP+:isocitrate dehydrogenase has been purified to homogeneity from lactating bovine mammary gland. Purification was achieved through the use of affinity and DEAE-cellulose chromatography. The isolated enzyme gives one band when stained for protein or enzyme activity on discontinuous alkaline gel electrophoresis. The enzyme has a molecular weight of 55,000 as estimated by sodium dodecyl sulfate-gel electrophoresis and a Stokes radius of 4.1 nm as measured by gel chromatography. The enzyme will not use NAD+ in place of NADP+ and has an absolute requirement for divalent cations. The apparent Km values for dl-isocitrate, Mn2+, and NADP+ were found to be 8, 6, and 11 μm, respectively. The Mn2+-ds-isocitrate complex is the most likely substrate for the mammary enzyme with a Km of 3 μm. The properties of mammary NADP+:isocitrate dehydrogenase are compared with those of the homologous enzymes from pig heart and bovine liver, and its characteristics are discussed with respect to the function of the enzyme in lactating mammary gland.  相似文献   

7.
Glucose-6-phosphate dehydrogenase from Leuconostoc mesenteroides utilizes either NAD+ or NADP+ as coenzyme. Kinetic studies showed that NAD+ and NADP+ interact with different enzyme forms (Olive, C., Geroch, M. E., and Levy, H. R. (1971) J. Biol. Chem.246, 2047–2057). In the present study the techniques of fluorescence quenching and fluorescence enhancement were used to investigate the interaction between Leuconostoc mesenteroides glucose-6-phosphate dehydrogenase and coenzymes. In addition, kinetic studies were performed to examine interaction between the enzyme and various coenzyme analogs. The maximum quenching of protein fluorescence is 5% for NADP+ and 50% for NAD+. The dissociation constant for NADP+, determined from fluorescence quenching measurements, is 3 μm, which is similar to the previously determined Km of 5.7 μm and Ki of 5 μm. The dissociation constant for NAD+ is 2.5 mm, which is 24 times larger than the previously determined Km of 0.106 mm. Glucose 1-phosphate, a substrate-competitive inhibitor, lowers the dissociation constant and maximum fluorescence quenching for NAD+ but not for NADP+. This suggests that glucose 6-phosphate may act similarly and thus play a role in enabling the enzyme to utilize NAD+ under physiological conditions. When NADPH binds to the enzyme its fluorescence is enhanced 2.3-fold. The enzyme was titrated with NADPH in the absence and presence of NAD+; binding of these two coenzymes is competitive. The dissociation constant for NADPH from these measurements is 24 μm; the previously determined Ki is 37.6 μm. The dissociation constant for NAD′ is 2.8 mm, in satisfactory agreement with the value obtained from protein fluorescence quenching measurements. Various compounds which resemble either the adenosine or the nicotinamide portion of the coenzyme structure are coenzyme-competitive inhibitors; 2′,5′-ADP, the most inhibitory analog tested, gives NADP+-competitive and NAD+-noncompetitive inhibition, consistent with the kinetic mechanism previously proposed. By using pairs of coenzyme-competitive inhibitors it was shown in kinetic studies that the two portions of the NAD+ structure cannot be accommodated on the enzyme simultaneously unies they are covalently linked. Fluorescence studies showed that there are both “buried” and “exposed” tryptophan residues in the enzyme structure.  相似文献   

8.
NAD+-dependent and NADP+-dependent glyceraldehyde-3-phosphate (G-3-P) dehydrogenases were isolated from Euglena gracilis and characterized as to their physical and chemical parameters. NAD+-G-3-P dehydrogenase was found to have a strong resemblance to similar enzymes from muscle tissue. It has a molecular weight of about 140,000, four subunits of identical size and charge, and a single species of NH2-terminal amino acid. Two sulfhydryl groups per subunit are present, one of which is directly involved in the catalytic activity and is rapidly titratable. The enzyme also exhibits the “half the sites reactivity” of sulfhydryl groups as defined by O. P. Malhotra and S. A. Bernhard ((1968) J. Biol. Chem. 243, 1243). The pH and temperature optima are also similar to those of the enzymes from muscle tissue, as are the reaction kinetics and the strict specificity for NAD+.NADP+-dependent G-3-P dehydrogenase is different in many respects. Its molecular weight is slightly lower (~136,000) than that of the NAD+ enzyme, though it also consists of four subunits. It has a higher affinity for the reverse reaction substrates, in line with its probable function in vivo in CO2 fixation. There is only one sulfhydryl group per subunit, and that is not involved in activity, suggesting a difference in reaction mechanisms between the two enzymes. The NADP+-dependent enzyme exhibits activation by ATP, whereas the NAD+-dependent enzyme is competitively inhibited by this nucleotide.The greatest difference observed is in the physical characteristics of the enzymes. NADP+-G-3-P dehydrogenase was highly hydrophobic. Its solubility in a 10% aqueous solution of p-dioxane was approximately four to five times that of the NAD+-enzyme. Isolation of the enzyme was accomplished by fractionation in 1,2-dimethoxyethane, which also stabilized the enzymatic activity, as did aqueous p-dioxane. The high axial ratio of the NADP+-enzyme (~9) coupled with its very low degree of hydration as well as the high degree of amidation of the dicarboxylic amino acids (>90%) indicates that the exterior of the enzyme molecule is probably hydrophobic in nature. This is in agreement with its in vivo hydrophobic environment in the chloroplast membrane and explains the lability of the enzyme once extracted into an aqueous environment as well as its stabilization in solvents.  相似文献   

9.
6-Phosphogluconate dehydrogenase (6PGDH), the third enzyme of the pentose phosphate pathway, catalyzes the oxidative decarboxylation of 6-phosphogluconate, making ribulose 5-phosphate, along with the reduction of NADP+ to NADPH and the release of CO2. Here, we report the first apo-form crystal structure of the pathogenic Klebsiella pneumoniae 6PGDH (Kp6PGDH) and the structures of the highly homologous Escherichia coli K12 6PGDH (Ec6PGDH) complexed with substrate, substrate/NADPH and glucose at high resolution. The binding of NADPH to one subunit of the homodimeric structure triggered a 10° rotation and resulting in a 7 Å movement of the coenzyme-binding domain. The coenzyme was thus trapped in a closed enzyme conformation, in contrast to the open conformation of the neighboring subunit. Comparison of our Ec/Kp6PGDH structures with those of other species illustrated how the domain conformation can be affected upon binding of the coenzyme, which in turn gives rise to concomitant movements of two important NADP+-interacting amino acids, M14 and N102. We propose that the catalysis follows an ordered binding mechanism with alternating conformational changes in the corresponding subunits, involving several related amino acid residues.  相似文献   

10.
The nicotinamide adenine dinucleotide-specific glutamate dehydrogenase (l-glutamate:NAD+ oxidoreductase, EC 1.4.1.2) of Chlorella sorokiniana was purified 1,000-fold to electrophoretic homogeneity. The native enzyme was shown to have a molecular weight of 180,000 and to be composed of four identical subunits with a molecular weight of 45,000. The N-terminal amino acid was determined to be lysine. The pH optima for the aminating and deaminating reactions were approximately 8 and 9, respectively. The Km values for α-ketoglutarate, NADH, NH4+, NAD+, and l-glutamate were 2 mm, 0.15 mm, 40 mm, 0.15 mm, and 60 mm, respectively. Whereas the Km for α-ketoglutarate and l-glutamate increased 10-fold, 1 pH unit above or below the pH optima for the aminating or deaminating reactions, respectively, the Km values for NADH and NAD+ were independent of change in pH from 7 to 9.6. By initial velocity, product inhibition, and equilibrium substrate exchange studies, the kinetic mechanism of enzyme was shown to be consistent with a bi uni uni uni ping-pong addition sequence. Although this kinetic mechanism differs from that reported for any other glutamate dehydrogenase, the chemical mechanism still appears to involve the formation of a Schiff base between α-ketoglutarate and an ε-amino group of a lysine residue in the enzyme. The physical, chemical, and kinetic properties of this enzyme differ greatly from those reported for the NH4+-inducible glutamate dehydrogenase in this organism.  相似文献   

11.
The cofactor-binding site of the NAD+-dependent Arabidopsis thaliana aldehyde dehydrogenase ALDH3H1 was analyzed to understand structural features determining cofactor-specificity. Homology modeling and mutant analysis elucidated important amino acid residues. Glu149 occupies a central position in the cofactor-binding cleft, and its carboxylate group coordinates the 2′- and 3′-hydroxyl groups of the adenosyl ribose ring of NAD+ and repels the 2′-phosphate moiety of NADP+. If Glu149 is mutated to Gln, Asp, Asn or Thr the binding of NAD+ is altered and rendered the enzyme capable of using NADP+. This change is attributed to a weaker steric hindrance and elimination of the electrostatic repulsion force of the 2′-phosphate of NADP+. Simultaneous mutations of Glu149 and Ile200, which is situated opposite of the cofactor binding cleft, improved the enzyme efficiency with NADP+. The double mutant ALDH3H1Glu149Thr/Ile200Val showed a good catalysis with NADP+. Subsequently a triple mutation was generated by replacing Val178 by Arg in order to create a “closed” cofactor binding site. The cofactor specificity was shifted even further in favor of NADP+, as the mutant ALDH3H1E149T/V178R/I200V uses NADP+ with almost 7-fold higher catalytic efficiency compared to NAD+. Our experiments suggest that residues occupying positions equivalent to 149, 178 and 200 constitute a group of amino acids in the ALDH3H1 protein determining cofactor affinity.  相似文献   

12.
Alcohol dehydrogenase of tomato (Lycopersicon esculentum) has been purified to homogeneity, using affinity chromatography on Cibacron F3GA-agarose. The enzyme is a dimer, Mr 90,000–100,000. The coenzyme is NAD+; no NADP+-dependent activity was detected even in crude extracts. Among saturated substrates, ethanol and acetaldehyde show the lowest apparent Km values (2.67 and 0.174 mm, respectively) and highest V values, supporting a primary role in acetaldehyde metabolism, with action also on “flavor aldehydes”; 2-unsaturated alcohols show still lower Km values, probably due to a more favorable Keq. This enzyme and other plant alcohol dehydrogenases form a definite class, intermediate in specificity between liver and yeast alcohol dehydrogenases: they differ from the former in being essentially inactive on secondary and aromatic substrates, from the latter in showing only a mild decrease in V with increasing chain length of alkyl substrates, and from both in showing the lowest Km as well as highest V on ethanol and acetaldehyde. The tomato enzyme differs from other reported plant enzymes in showing substantial activity on geraniol. Kinetic studies are in agreement with an ordered sequential mechanism. The enzyme is inhibited slowly by iodoacetamide, and reversibly by acetamide and zinc-chelating compounds.  相似文献   

13.
Characteristics of the three major ammonia assimilatory enzymes, glutamate dehydrogenase (GDH), glutamine synthetase (GS) and glutamate synthase (GOGAT) in Corynebacterium callunae (NCIB 10338) were examined. The GDH of C. callunae specifically required NADPH and NADP+ as coenzymes in the amination and deamination reactions, respectively. This enzyme showed a marked specificity for -ketoglutarate and glutamate as substrates. The optimum pH was 7.2 for NADPH-GDH activity (amination) and 9.0 for NADP+-GDH activity (deamination). The results showed that NADPH-GDH and NADP+-GDH activities were controlled primarily by product inhibition and that the feedback effectors alanine and valine played a minor role in the control of NADPH-GDH activity. The transferase activity of GS was dependent on Mn+2 while the biosynthetic activity of the enzyme was dependent on Mg2+ as essential activators. The pH optima for transferase and biosynthetic activities were 8.0 and 7.0, respectively. In the transfer reaction, the K m values were 15.2 mM for glutamine, 1.46 mM for hydroxylamine, 3.5×10-3 mM for ADP and 1.03 mM for arsenate. Feedback inhibition by alanine, glycine and serine was also found to play an important role in controlling GS activity. In addition, the enzyme activity was sensitive to ATP. The transferase activity of the enzyme was responsive to ionic strength as well as the specific monovalent cation present. GOGAT of C. callunae utilized either NADPH or NADH as coenzymes, although the latter was less effective. The enzyme specifically required -ketoglutarate and glutamine as substrates. In cells grown in a medium with glutamate as the nitrogen source, the optimum pH was 7.6 for NADPH-GOGAT activity and 6.8 for NADH-GOGAT activity. Findings showed that NADPH-GOGAT and NADH-GOGAT activities were controlled by product inhibition caused by NADP+ and NAD+, respectively, and that ATP also had an important role in the control of NADPH-GOGAT activity. Both activities of GOGAT were found to be inhibited by azaserine.Abbreviations GDH glutamate dehydrogenase - GOGAT glutamate synthase - GS glutamine synthetase  相似文献   

14.
Enzymatic studies have been performed on a local strain of Aspergillus niger to find a correlation with citric acid accumulation. The activity of aconitase [aconitate hydratase, citrate(isocitrate) hydrolyase, EC 4.2.1.3] and isocitrate dehydrogenase (NADP+) [threo-ds-isocitrate:NADP+ oxidoreductase (decarboxylating) EC 1.1.1.42] decreased after 4 days whereas that of citrate synthase [citrate oxaloacetate-lyase (pro-3S-CH2COO?acetylCoA), EC 4.1.3.7] did so after 8 days, when citric acid accumulation in the medium reached a maximum (45.9 mg ml?1). In vitro studies with mycelial cell-free extracts demonstrated inhibition of citrate synthase activity by sodium azide and potassium ferricyanide on both the 4th and 8th days. Aconitase was inhibited by sodium arsenate, sodium fluoride, iodoacetic acid and potassium ferricyanide only on the 4th day. Isocitrate dehydrogenase (NADP+) activity on the 4th and 8th days was inhibited by iodoacetic acid but was stimulated by potassium ferricyanide. The possible existence of isozyme species of these enzymes is discussed.  相似文献   

15.
In general, eukaryotic glucose-6-phosphate dehydrogenases (G6PDHs) are structurally stabilized by NADP+. Here we show by spectrofluorometric analysis, thermal and urea denaturation, and trypsin proteolysis, that a different mechanism stabilizes the enzyme from Pseudomonas aeruginosa (PaG6PDH) (EC 1.1.1.363). The spectrofluorometric analysis of the emission of 8-anilino-1-naphthalenesulfonic acid (ANS) indicates that this stabilization is the result of a structural change in the enzyme caused by G6P. The similarity between the Kd values determined for the PaG6PDH-G6P complex (78.0 ± 7.9 μM) and the K0.5 values determined for G6P (57.9 ± 2.5 and 104.5 ± 9.3 μM in the NADP+- and NAD+-dependent reactions, respectively) suggests that the structural changes are the result of G6P binding to the active site of PaG6PDH. Modeling of PaG6PDH indicated the residues that potentially bind the ligand. These results and a phylogenetic analysis of the amino acid sequences of forty-four G6PDHs, suggest that the stabilization observed for PaG6PDH could be a characteristic that distinguishes this and other G6PDHs that use NAD+ and NADP+ from those that use NADP+ only or preferentially, such as those found in eukaryotes. This characteristic could be related to the metabolic roles these enzymes play in the organisms to which they belong.  相似文献   

16.
Product inhibition of lysyl hydroxylase (peptidyllysine, 2-oxoglutarate:oxygen 5-oxidoreductase, EC 1.14.11.4) was studied with succinate, CO2, dehydroascorbate and hydroxylysine-rich polypeptide chains. The product inhibition patterns and addition data are consistent with a reaction mechanism involving an ordered binding of Fe2+, α-ketoglutarate, O2 and the peptide substrate to the enzyme in this order, and an ordered release of the hydroxylated peptide, CO2, succinate and Fe2+, in which Fe2+ need not leave the enzyme during each catalytic cycle and in which the order of release of the hydroxylated peptide and CO2 is uncertain. Ascorbate probably reacts by a substitution mechanism, either after the release of the hydroxylated peptide, CO2 and succinate or after the release of all products, including Fe2+, and dehydroascorbate is released before the binding of Fe2+. It is suggested that the ascorbate reaction is required to reduce either the enzyme-iron complex or the free enzyme, which may be oxidized by a side-reaction during some catalytic cycles, but not the majority. The mechanisms of the prolyl 4-hydroxylase and lysyl hydroxylase reactions are suggested to be identical.Zn2+, several citric acid cycle intermediates, nitroblue tetrazolium and homogentisic acid inhibited lysyl hydroxylase competitively with regard to Fe2+, α-ketoglutarate, O2 and ascorbate respectively, and epinephrine noncompetitively with regard to all cosubstrates. Apparent Ki values are given for the product and other inhibitors.  相似文献   

17.
Sepiapterin reductase from rat erythrocyte hemolysate was purified 2000-fold to apparent homogeneity with 30% yield. The specific activity of the purified enzyme was 18 units/mg protein, and its molecular weight was 55 000. The enzyme consists of two identical subunits, each of which has a molecular weight of 27 500. The enzyme showed a single peak by isoelectric focusing with a pI of 4.9 and partial specific volume of 0.73 cm3/g. The amino acid composition was determined. pH optimum of the enzyme was 5.5. The equilibrium constant of 2.2·109 of the enzyme showed that the equilibrium lies much in favor of dihydrobiopterin formation from sepiapterin in rat erythrocytes. From steady-state kinetic measurements, ordered bi-bi mechanism was proposed to the reaction of sepiapterin reductase in which NADPH binds to free enzyme and sepiapterin binds next. NADP+ is released after the release of dihydrobiopterin. The Km values for sepiapterin and NADPH were 15.4 μM and 1.7 μM, respectively, and the Vmax value was 21.7 μmol/min per mg.  相似文献   

18.
NADP+-dependent cytoplasmic malic enzyme was purified to homogeneity from mouse kidneys by a two-step procedure involving 8-(6-aminohexyl)-amino-2, 5-ADP-Sepharose affinity chromatography and DEAE-Sephadex ion exchange chromatography. The biochemical properties of the purified enzyme from DBA/2J mice were characterized. These include the determination of molecular weight and amino acid compositions, steady-state kinetics, thermal stability and inactivations by iodoacetate and urea. The native enzyme is a tetramer with a molecular weight of 270,000.Km's for NADP+, l-malate, NADPH and pyruvate were determined to be 3.3 µm,, 50 µm, 10.5 gm respectively. Similar to the pigeon liver enzyme, the mouse enzyme exhibits an ordered kinetic mechanism proceeding with the binding of coenzyme first. The enzyme is only weakly inhibited by ATP and other cellular metabolites. A remarkable similarity in amino acid compositions was found between the mouse and rat liver malic enzymes.Abbreviations DTNB 5,5-dithio, bis-nitrobenzoic acid  相似文献   

19.
Serine hydroxymethyltransferase (SHMT) catalyzes the interconversion of serine and tetrahydrofolate (THF) to glycine and methylenetetrahydrofolate. cDNA encoding Bombyx mori SHMT (bmSHMT) was cloned and sequenced. The deduced amino acid sequence consisted of 465 amino acids and was found to share homology with other SHMTs. Recombinant bmSHMT was overexpressed in Escherichia coli and purified to homogeneity. The enzyme showed optimum activity at pH 3.0 and 30°C and was stable under acidic conditions. The Km and kcat/Km values for THF in the presence of Nicotinamide adenine dinucleotide phosphate (NADP+) were 0.055 mM and 0.081 mM?1 s?1, respectively, whereas those toward NADP+ were 0.16 mM and 0.018 mM?1 s?1 and toward l ‐serine were 1.8 mM and 0.0022 mM?1 s?1, respectively. Mutagenesis experiments revealed that His119, His132, and His135 are important for enzymatic activity. Our results provide insight into the roles and regulation mechanism of one‐carbon metabolism in the silkworm B. mori.  相似文献   

20.
Isocitrate dehydrogenase (IDH) is one of the key enzymes in tricarboxylic acid cycle, widely distributed in Archaea, Bacteria and Eukarya. Here, we report for the first time the cloning, expression and characterization of a monomeric NADP+-dependent IDH from Streptomyces diastaticus No. 7 strain M1033 (SdIDH). Molecular mass of SdIDH was about 80 kDa and showed high amino acid sequence identity with known monomeric IDHs. Maximal activity of SdIDH was observed at pH 8.0 (Mn2+) and 9.0 (Mg2+), and the optimal temperature was 40 °C (Mn2+) and 37 °C (Mg2+). Heat-inactivation studies showed that SdIDH remained about 50 % activity after 20 min of incubation at 47 °C. SdIDH displayed a 19,000 and 32,000-fold (k cat/K m) preference for NADP+ over NAD+ with Mn2+ and Mg2+, respectively. Our work implicate that SdIDH is a divalent metal ion-dependent monomeric IDH with remarkably high coenzyme preference for NADP+. This work may provide fundamental information for further investigation on the catalytic mechanism of monomeric IDH and give a clue to disclose the real cause of IDH monomerization.  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