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 63 毫秒
1.
Metabolic engineering of Saccharomyces cerevisiae for xylose fermentation has often relied on insertion of a heterologous pathway consisting of nicotinamide adenine dinucleotide (phosphate) NAD(P)H-dependent xylose reductase (XR) and NAD+-dependent xylitol dehydrogenase (XDH). Low ethanol yield, formation of xylitol and other fermentation by-products are seen for many of the S. cerevisiae strains constructed in this way. This has been ascribed to incomplete coenzyme recycling in the steps catalyzed by XR and XDH. Despite various protein-engineering efforts to alter the coenzyme specificity of XR and XDH individually, a pair of enzymes displaying matched utilization of NAD(H) and NADP(H) was not previously reported. We have introduced multiple site-directed mutations in the coenzyme-binding pocket of Galactocandida mastotermitis XDH to enable activity with NADP+, which is lacking in the wild-type enzyme. We describe four enzyme variants showing activity for xylitol oxidation by NADP+ and NAD+. One of the XDH variants utilized NADP+ about 4 times more efficiently than NAD+. This is close to the preference for NADPH compared with NADH in mutants of Candida tenuis XR. Compared to an S. cerevisiae-reference strain expressing the genes for the wild-type enzymes, the strains comprising the gene encoding the mutated XDH in combination a matched XR mutant gene showed up to 50% decreased glycerol yield without increase in ethanol during xylose fermentation.  相似文献   

2.
Two inducible NADP+-dependent glycerol dehydrogenase (GlcDH) activities were identified in Mucor circinelloides strain YR-1. One of these, denoted iGlcDH2, was specifically induced by n-decanol when it was used as sole carbon source in the culture medium, and the second, denoted iGlcDH1, was induced by alcohols and aliphatic or aromatic hydrocarbons when glycerol was used as the only substrate. iGlcDH2 was found to have a much broader substrate specificity than iGlcDH1, with a low activity as an ethanol dehydrogenase with NAD+ or NADP+ as cofactor. Both isozymes showed an optimum pH for activity of 9.0 in Tris-HCl buffer and are subject to carbon catabolite repression. In contrast, the constitutive NADP+-dependent glycerol dehydrogenases (GlcDHI, II, and III) were only present in cell extracts when the fungus was grown in glycolytic carbon sources or glycerol under oxygenation, and their optimum pH was 7.0 in Tris-HCl buffer. In addition to these five NADP+-dependent glycerol dehydrogenases, a NAD+-dependent alcohol dehydrogenase is also present in glycerol or n-decanol medium; this enzyme was found to have weak activity as a glycerol dehydrogenase.  相似文献   

3.
Cell-free extracts of the xylose fermenting yeast Pichia stipitis exhibited xylitol dehydrogenase activity with NAD+ and NADP+. During the purification step on DEAE-sephadex A-50 a NAD+-dependent xylitol dehydrogenase could be separated from a NADP+-dependent. The NAD+-xylitol dehydrogenase was further purified to electrophoretic homogeneity via gel and affinity chromatography. The purified enzyme was most active at pH 9 and 35°C. Its molecular weight was determined to be 63,000 dalton by Sephadex G-200 column chromatography, and that of its subunit was 32,000 dalton by sodium dodecyl sulphate polyacrylamide gel electrophoresis. From the results of substrate specificity, the enzyme should be named l-iditol:NAD+-5-oxidoreductase (EC 1.1.1.14, sorbitol dehydrogenase).  相似文献   

4.
DEAE-cellulose chromatography of extracts of free-living Rhizobium meliloti cells revealed separate NAD+-dependent and NADP+-dependent malic enzyme activities. The NAD+ malic enzyme exhibited more activity with NAD+ as cofactor, but also showed some activity with NADP+. The NADP+ malic enzyme only showed activity when NADP+ was supplied as cofactor. Three independent transposon-induced mutants of R. meliloti which lacked NADP+ malic enzyme activity (dme) but retained NADP+ malic enzyme activity were isolated. In an otherwise wild-type background, the dme mutations did not alter the carbon utilization phenotype; however, nodules induced by these mutants failed to fix N2. Structurally, these nodules appeared to develop like wild-type nodules up to the stage where N2-fixation would normally begin. These results support the proposal that NAD+ malic enzyme, together with pyruvate dehydrogenase, functions in the generation of acetyl-CoA required for TCA cycle function in N2-fixing bacteroids which metabolize C4-dicarboxylic acids supplied by the plant.  相似文献   

5.
The subcellular distribution of NADP+ and NAD+-dependent glucose-6-phosphate and galactose-6-phosphate dehydrogenases were studied in rat liver, heart, brain, and chick brain. Only liver particulate fractions oxidized glucose-6-phosphate and galactose-6-phosphate with either NADP+ or NAD+ as cofactor. While all of the tissues examined had NADP+-dependent glucose-6-phosphate dehydrogenase activity, only rat liver and rat brain soluble fractions had NADP+-dependent galactose-6-phosphate dehydrogenase activity. Rat liver microsomal and rat brain soluble galactose-6-phosphate dehydrogenase activities were kinetically different (Km's 0.5 mm and 10 mm, respectively, for galactose-6-phosphate), although their reaction products were both 6-phosphogalactonate. Rat brain subcellular fractions did not oxidize 6-phosphogalactonate with either NADP+ or NAD+ cofactors but phosphatase activities hydrolyzing 6-phosphogalactonate, galactose-6-phosphate and galactose-1-phosphate were found in crude brain homogenates. In addition, galactose-6-phosphate and 6-phosphogalactonate were tested as inhibitors of various enzymes, with largely negative results, except that 6-phosphogalactonate was a competitive inhibitor (Ki = 0.5 mM) of rat brain 6-phosphogluconate dehydrogenase.  相似文献   

6.
NAD+ and NADP+, chemically similar and with almost identical standard oxidation–reduction potentials, nevertheless have distinct roles, NAD+ serving catabolism and ATP generation whereas NADPH is the biosynthetic reductant. Separating these roles requires strict specificity for one or the other coenzyme for most dehydrogenases. In many organisms this holds also for glutamate dehydrogenases (GDH), NAD+-dependent for glutamate oxidation, NADP+-dependent for fixing ammonia. In higher animals, however, GDH has dual specificity. It has been suggested that GDH in mitochondria reacts only with NADP(H), the NAD+ reaction being an in vitro artefact. However, contrary evidence suggests mitochondrial GDH not only reacts with NAD+ but maintains equilibrium using the same pool as accessed by β-hydroxybutyrate dehydrogenase. Another complication is the presence of an energy-linked dehydrogenase driving NADP+ reduction by NADH, maintaining the coenzyme pools at different oxidation–reduction potentials. Its coexistence with GDH makes possible a futile cycle, control of which is not yet properly explained. Structural studies show NAD+-dependent, NADP+-dependent and dual-specificity GDHs are closely related and a few site-directed mutations can reverse specificity. Specificity for NAD+ or for NADP+ has probably emerged repeatedly during evolution, using different structural solutions on different occasions. In various GDHs the P7 position in the coenzyme-binding domain plays a key role. However, whereas in other dehydrogenases an acidic P7 residue usually hydrogen bonds to the 2′- and 3′-hydroxyls, dictating NAD+ specificity, among GDHs, depending on detailed conformation of surrounding residues, an acidic P7 may permit binding of NAD+ only, NADP+ only, or in higher animals both.  相似文献   

7.
Two types of 15-hydroxyprostaglandin dehydrogenase (NAD+ and NADP+ dependent) were demonstrated in bovine mesentric arteries and veins. The 15-hydroxyprostaglandin dehydrogenase activity was found in the high-speed supernatant, suggesting that these enzymes are associated with the cytoplasmic fraction of the blood vessels. The levels of activities of both NAD+- and NADP+-dependent dehydrogenases were similar in mesentric blood vessels. Prostaglandin F was preferred to the prostaglandin E2 as subtrate by both NAD+ and NADP+ dependent enzymes. The presence of 15-hydroxyprostaglandin dehydrogenase in blood vessels may play a siginificant role in the regulation of intracellular levels of prostaglandins of the E and F series in blood vessels.  相似文献   

8.
Several denitrifying Pseudomonas strains contained an NADP+-specific 2-oxoglutarate dehydrogenase, in contrast to an NAD+-specific pyruvate dehydrogenase, if the cells were grown anaerobically with aromatic compounds. With non-aromatic substrates or after aerobic growth the coenzyme specificity of 2-oxoglutarate dehydrogenase changed to NAD+-specificity. The reaction stoichiometry and the apparent K m-values of the enriched enzymes were determined: pyruvate 0.5 mM, coenzyme A 0.05 mM, NAD+ 0.25 mM; 2-oxoglutarate 0.6 mM, coenzyme A 0.05 mM, NADP+ 0.03 mM. Isocitrate dehydrogenase was NADP+-specific. The findings suggest that these strains contained at least two lipoamide dehydrogenases, one NAD+-specific, the other NADP+-specific.  相似文献   

9.
The autotrophic nitrifying bacterium Nitrosomonas europaea does not synthesize 2-oxoglutarate (α-ketoglutarate) dehydrogenase under aerobic conditions and so has an incomplete citric acid cycle. l-malate (S-malate) dehydrogenase (MDH) from N. europaea was predicted to show similarity to the NADP+-dependent enzymes from chloroplasts and was separated from the NAD+-dependent proteins from most other bacteria or mitochondria. MDH activity in a soluble fraction from N. europaea ATCC 19718 was measured spectrophotometrically and exhibited simple Michaelis–Menten kinetics. In the reductive direction, activity with NADH increased from pH 6.0 to 8.5 but activity with NADPH was consistently lower and decreased with pH. At pH 7.0, the K m for oxaloacetate was 20 μM; the K m for NADH was 22 μM but that for NADPH was at least 10 times higher. In the oxidative direction, activity with NAD+ increased with pH but there was very little activity with NADP+. At pH 7.0, the K m for l-malate was 5 mM and the K m for NAD+ was 24 μM. The reductive activity was quite insensitive to inhibition by l-malate but the oxidative activity was very sensitive to oxaloacetate. MDH activity was not strongly activated or inhibited by glycolytic or citric acid cycle metabolites, adenine nucleotides, NaCl concentrations, or most metal ions, but increased with temperature up to about 55 °C. The reductive activity was consistently 10–20 times higher than the oxidative activity. These results indicate that the l-malate dehydrogenase in N. europaea is similar to other NAD+-dependent MDHs (EC 1.1.1.37) but physiologically adapted for its role in a reductive biosynthetic sequence.  相似文献   

10.
NADP-Utilizing Enzymes in the Matrix of Plant Mitochondria   总被引:9,自引:4,他引:5       下载免费PDF全文
Purified potato tuber (Solanum tuberosum L. cv Bintie) mitochondria contain soluble, highly latent NAD+- and NADP+-isocitrate dehydrogenases, NAD+- and NADP+-malate dehydrogenases, as well as an NADPH-specific glutathione reductase (160, 25, 7200, 160, and 16 nanomoles NAD(P)H per minute and milligram protein, respectively). The two isocitrate dehydrogenase activities, but not the two malate dehydrogenase activities, could be separated by ammonium sulfate precipitation. Thus, the NADP+-isocitrate dehydrogenase activity is due to a separate matrix enzyme, whereas the NADP+-malate dehydrogenase activity is probably due to unspecificity of the NAD+-malate dehydrogenase. NADP+-specific isocitrate dehydrogenase had much lower Kms for NADP+ and isocitrate (5.1 and 10.7 micromolar, respectively) than the NAD+-specific enzyme (101 micromolar for NAD+ and 184 micromolar for isocitrate). A broad activity optimum at pH 7.4 to 9.0 was found for the NADP+-specific isocitrate dehydrogenase whereas the NAD+-specific enzyme had a sharp optimum at pH 7.8. Externally added NADP+ stimulated both isocitrate and malate oxidation by intact mitochondria under conditions where external NADPH oxidation was inhibited. This shows that (a) NADP+ is taken up by the mitochondria across the inner membrane and into the matrix, and (b) NADP+-reducing activities of malate dehydrogenase and the NADP+-specific isocitrate dehydrogenase in the matrix can contribute to electron transport in intact plant mitochondria. The physiological relevance of mitochondrial NADP(H) and soluble NADP(H)-consuming enzymes is discussed in relation to other known mitochondrial NADP(H)-utilizing enzymes.  相似文献   

11.
A sensitive isotope exchange method was developed to assess the requirements for and compartmentation of pyruvate and oxalacetate production from malate in proliferating and nonproliferating human fibroblasts. Malatedependent pyruvate production (malic enzyme activity) in the particulate fraction containing the mitochondria was dependent on either NAD+ or NADP+. The production of pyruvate from malate in the soluble, cytosolic fraction was strictly dependent on NADP+. Oxalacetate production from malate (malate dehydrogenase, EC 1.1.1.37) in both the particulate and soluble fraction was strictly dependent on NAD+. Relative to nonproliferating cells, NAD+-linked malic enzyme activity was slightly reduced and the NADP+-linked activity was unchanged in the particulate fraction of serum-stimulated, exponentially proliferating cells. However, a reduced activity of particulate malate dehydrogenase resulted in a two-fold increase in the ratio of NAD(P)+-linked malic enzyme to NAD+-linked malate dehydrogenase activity in the particulate fraction of proliferating fibroblasts. An increase in soluble NADP+-dependent malic enzyme activity and a decrease in NAD+-linked malate dehydrogenase indictated an increase in the ratio of pyruvate-producing to oxalacetate-producing malate oxidase activity in the cytosol of proliterating cells. These coordinate changes may affect the relative amount of malate that is oxidized to oxalacetate and pyruvate in proliferating cells and, therefore, the efficient utilization of glutamine as a respiratory fuel during cell proliferation.  相似文献   

12.
Summary From tested yeast-like organisms, onlyGeotrichum candidum showed the same activity of glucose-6-phosphate dehydrogenase with both NAD+ and NADP+. i. e. 0.017–0.019 mol NADH/min. mg dry weight of cell free extracts. Omission of Mg++ in the reaction mixture did not influence the activity of the enzyme in the presence of NAD+. Cell free extracts ofEndomyces magnusii showed only low activity of this enzyme and the ratio of its activity in the presence of NAD+ and NADP+, respectively, varied in individual cultures.Rhodotorula glutinis showed only an NADP+-dependent activity.  相似文献   

13.
The gene encoding isocitrate dehydrogenase (IDH) of Methylococcus capsulatus (McIDH) was cloned and overexpressed in Escherichia coli. The purified enzyme was NAD+-dependent with a thermal optimum for activity at 55–60°C and an apparent midpoint melting temperature (T m) of 70°C. Analytical ultracentrifugation (AUC) revealed a homotetrameric state, and McIDH thus represents the first homotetrameric NAD+-dependent IDH that has been characterized. Based on a structural alignment of McIDH and homotetrameric homoisocitrate dehydrogenase (HDH) from Thermus thermophilus (TtHDH), we identified the clasp-like domain of McIDH as a likely site for tetramerization. McIDH showed moreover, higher sequence identity (48%) to TtHDH than to previously characterized IDHs. Putative NAD+-IDHs with high sequence identity (48–57%) to McIDH were however identified in a variety of bacteria showing that NAD+-dependent IDHs are indeed widespread within the domain, Bacteria. Phylogenetic analysis including these new sequences revealed a close relationship with eukaryal allosterically regulated NAD+-IDH and the subfamily III of IDH was redefined to include bacterial NAD+- and NADP+-dependent IDHs. This apparent relationship suggests that the mitochondrial genes encoding NAD+-IDH are derived from the McIDH-like IDHs.  相似文献   

14.
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   

15.
Structural analysis of glucose dehydrogenase from Haloferax mediterranei revealed that the adenosine 2′-phosphate of NADP+ was stabilized by the side chains of Arg207 and Arg208. To investigate the structural determinants for coenzyme specificity, several mutants involving residues Gly206, Arg207 and Arg208 were engineered and kinetically characterized. The single mutants G206D and R207I were less efficient with NADP+ than the wild type, and the double and triple mutants G206D/R207I and G206D/R207I/R208N showed no activity with NADP+.In the single mutant G206D, the relation kcat/KNAD+ was 1.6 times higher than in the wild type, resulting in an enzyme that preferred NAD+ over NADP+. The single mutation was sufficient to modify coenzyme specificity, whereas other dehydrogenases usually required more than one or two mutations to change coenzyme specificity. However, the highest reaction rates were reached with the double mutant G206D/R207I and with coenzyme NAD+, where the kcat was 1.6 times higher than the kcat of the wild-type enzyme with NADP+. However, catalytic efficiency with NAD+ was lower, as the Km value for coenzyme was 77 times higher than the wild type with NADP+.  相似文献   

16.
Sixteen Tn916-induced mutants of Clostridium acetobutylicum were selected that were defective in the production of acetone and butanol. Formation of ethanol, however, was only partially affected. The strains differed with respect to the degree of solvent formation ability and could be assigned to three different groups. Type I mutants (2 strains) were completely defective in acetone and butanol production and contained one or three copies of Tn916 in the chromosome. Analysis of the mutants for enzymes responsible for solvent production revealed the presence of a formerly unknown, specific acetaldehyde dehydrogenase. The data obtained also strongly indicate that the NADP+-dependent alcohol dehydrogenase is in vivo reponsible for ethanol formation, whereas the NAD+-dependent alcohol dehydrogenase is probably involved in butanol production. No activity of this enzyme together with all other enzymes in the acetone and butanol pathway could be found in type I strains. All tetracycline-resistant mutants obtained did no longer sporulate.Non-standard abbreviations AADC acetoacetate decarboxylase - AcaDH acetaldehyde dehydrogenase - BuaDH butyraldehyde dehydrogenase - CoA-TF acetoacetyl coenzyme A: acetate/butyrate: coenzyme A transferase - NAD-ADH, NAD+ dependent alcohol dehydrogenase - NADP-ADH, NADP+ dependent alcohol dehydrogenase  相似文献   

17.
The NAD+‐dependent lactate dehydrogenase from Bacillus subtilis (BsLDH) catalyzes the enantioselective reduction of pyruvate to lactate. BsLDH is highly specific to NAD+ and exhibits only a low activity with NADP+ as cofactor. Based on the high activity and good stability of LDHs, these enzymes have been frequently used for the regeneration of NAD+. While an application in the regeneration of NADP+ is not sufficient due to the cofactor preference of the BsLDH. In addition, NADP+‐dependent LDHs have not yet been found in nature. Therefore, a structure‐based approach was performed to predict amino acids involved in the cofactor specificity. Methods of site‐saturation mutagenesis were applied to vary these amino acids, with the aim to alter the cofactor specificity of the BsLDH. Five constructed libraries were screened for improved NADP+ acceptance. The mutant V39R was identified to have increased activity with NADP+ relative to the wild type. V39R was purified and biochemically characterized. V39R showed excellent kinetic properties with NADP(H) and NAD(H), for instance the maximal specific activity with NADPH was enhanced 100‐fold to 90.8 U/mg. Furthermore, a 249‐fold increased catalytic efficiency was observed. Surprisingly, the activity with NADH was also significantly improved. Overall, we were able to successfully apply V39R in the regeneration of NADP+ in an enzyme‐coupled approach combined with the NADP+‐dependent alcohol dehydrogenase from Lactobacillus kefir. We demonstrate for the first time an application of an LDH in the regeneration of NADP+.  相似文献   

18.
Summary A quantitative cytochemical assay for NAD+ kinase-like activity in the guinea-pig thyroid gland is described. The NADP+ produced by the activity of the kinase was used to drive the NADP+-dependent enzyme glucose-6-phosphate dehydrogenase which is endogenous to the tissue. The activity of glucose-6-phosphate dehydrogenase is greatly in excess of that of the kinase and was unaffected by the constituents of the kinase incubation medium (ATP, Mg2+ and NAD+) either alone or in combination. Kinase activity was dependent both on ATP and Mg2+, with maximal activity seen when the Mg-ATP ratio was between 1:1 and 4:1. Free ATP inhibited the activity of the enzyme. Enzyme activity was exhibited over a broad pH range (7–9) with a peak at pH 8.2. The sulphhydryl-blocking agents,p-chloromercuribenzoate, iodoacetate and iodoacetamide (at 1 mM), completely abolished kinase activity but were without effect on glucose-6-phosphate dehydrogenase activity.N-ethylmaleimide and citrate (both at 1 mM) had no effect on either kinase or glucose-6-phosphate dehydrogenase activities.  相似文献   

19.
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   

20.
Isocitrate dehydrogenase (IDH) activities were measured in mitochondria isolated from aerial parts of 21-day-old spruce (Picea abies L. Karst.) seedlings. Mitochondria were purified by two methods, involving continuous and discontinuous Percoll gradients. Whatever the method of purification, the mitochondrial outer membrane was about 69% intact, and the mitochondria contained very low cytosolic, chloroplastic and peroxisomal contaminations. Nevertheless, as judged by the recovery of fumarase activity, purification on a continuous 28% Percoll gradient gave the best yield in mitochondria, which exhibited a high degree of inner membrane intactness (91%). The purified mitochondria oxidized succinate and malate with good respiratory control and ADP/O ratios. The highest oxidation rate was obtained with succinate as substrate, and malate oxidation was improved (+ 60%) by addition of exogenous NAD+. Experiments using standard respiratory chain inhibitors indicated that, in spruce mitochondria, the alternative pathway was present. Both NAD+-isocitrate dehydrogenase (EC 1.1.1.41) and NADP+-isocitrate dehydrogenase (EC 1.1.1.42) were present in the mitochondrial matrix fraction, and NAD+-IDH activity was about 2-fold higher than NADP+-IDH activity. The NAD+-IDH showed sigmoidal kinetics in response to isocitrate and standard Michaelis-Menten kinetics for NAD+ and Mg2+. The NADP+-IDH, in contrast, displayed lower Km values. For NAD+-IDH the pH optimum was at 7.4, whereas NADP+-IDH exhibited a broad pH optimum between 8.3 and 9. In addition, NAD+-IDH was more thermolabile. Adenine nucleotides and 2-oxoglutarate were found to inhibit NAD(P)+-IDH activities only at high concentrations.  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