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.
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   

2.
Periodate-oxidized NADP+ (dialdehyde-NADP+) inactivated soluble ferredoxin-NADP+ oxidoreductase and combined covalently to the enzyme. This inactivation was first order with respect to dialdehyde-NADP+ and followed saturation kinetics, indicating that the enzyme initially forms a reversible complex with the inactivator. NADP+ afforded complete protection against inactivation, while spinach ferredoxin was uneffective. In the presence of exogenous ferredoxin and illuminated thylakoids, the nucleotide analog functioned as a coenzyme for the reductase, although with rather lower efficiency than NADP+. It also acted as a competitive inhibitor with respect to NADPH in diaphorase activity. Incorporation of radioactivity from periodate-oxidized [3H]NADP+ gave a stoichiometry of 0.85 mol of reagent/mol of reductase, indicating that the modification of a single residue in the flavoprotein is responsible for the loss of enzymatic activity.  相似文献   

3.
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   

4.
Redox interconversion of glutathione reductase was studiedin situ withS. cerevisiae. The enzyme was more sensitive to redox inactivation in 24 hour-starved cells than in freshly-grown ones. While 5 μM NADPH or 100 μM NADH caused 50% inactivation in normal cells in 30 min, 0.75 μM NADPH or 50 μM NADH promoted a similar effect in starved cells. GSSG reactivated the enzyme previously inactivated by NADPH, ascertaining that the enzyme was subjected to redox interconversion. Low EDTA concentrations fully protected the enzyme from NADPH inactivation, thus confirming the participation of metals in such a process. Extensive inactivation was obtained in permeabilized cells incubated with glucose-6-phosphate or 6-phosphogluconate, in agreement with the very high specific activities of the corresponding dehydrogenases. Some inactivation was also observed with malate, L-lactate, gluconate or isocitrate in the presence of low NADP+ concentrations. The inactivation of yeast glutathione reductase has also been studiedin vivo. The activity decreased to 75% after 2 hours of growth with glucono-δ-lactone as carbon source, while NADPH rose to 144% and NADP+ fell to 86% of their initial values. Greater changes were observed in the presence of 1.5 μM rotenone: enzymatic activity descended to 23% of the control value, while the NADH/NAD+ and NADPH/NADP+ ratios rose to 171% and 262% of their initial values, respectively. Such results indicate that the lowered redox potential of the pyridine nucleotide pool existing when glucono-δ-lactone is oxidized promotesin vivo inactivation of glutathione reductase.  相似文献   

5.
The aim of the present study was the investigation of the occurrence of NADPH-generating pathways in the endoplasmic reticulum others then hexose-6-phosphate dehydrogenase. A significant isocitrate and a moderate malate-dependent NADP+ reduction were observed in endoplasmic reticulum-derived rat liver microsomes. The isocitrate-dependent activity was very likely attributable to the appearance of the cytosolic isocitrate dehydrogenase isozyme in the lumen. The isocitrate dehydrogenase activity of microsomes was present in the luminal fraction; it showed a strong preference towards NADP+versus NAD+, and it was almost completely latent. Antibodies against the cytosolic isoform of isocitrate dehydrogenase immunorevealed a microsomal protein of identical molecular weight; the microsomal enzyme showed similar kinetic parameters and oxalomalate inhibition as the cytosolic one. Measurable luminal isocitrate dehydrogenase activity was also present in microsomes from rat epididymal fat. The results suggest that isocitrate dehydrogenase is an important NADPH-generating enzyme in the endoplasmic reticulum.  相似文献   

6.
The Rhodospirillum rubrum pyridine dinucleotide transhydrogenase system is comprised of a membrane-bound component and an easily dissociable soluble factor. Active transhydrogenase complex was solubilized by extraction of chromatophores with lysolecithin. The membrane component was also extracted from membranes depleted of soluble factor. The solubilized membrane component reconstituted transhydrogenase activity upon addition of soluble factor. Various other ionic and non-ionic detergents, including Triton X-100, Lubrol WX, deoxycholate, and digitonin, were ineffectual for solubilization and/or inhibited the enzyme at higher concentrations. The solubilized membrane component was significantly less thermal stable than the membrane-bound component. None of the pyridine dinucleotide substrate affected the thermostability of the solubilized membrane-bound component, whereas NADP+ and NADPH afforded protection to membrane-bound component. NADPH stimulated trypsin inactivation of membrane-bound component to a greater extent than NADP+, but inactivation of solubilized membrane component was stimulated to the same extent by both pyridine dinucleotides. The solubilized membrane component appears to have a slightly higher affinity for soluble factor than does the membrane-bound component.Abbreviations AcPyAD+ oxidized 3-acetylpyridine adenine dinucleotide - BChl bacteriochlorophyll - CT-particles chromatophores depleted of soluble transhydrogenase factor and devoid of transhydrogenase activity This work was supported by Grant GM 22070 from the National Institutes of Health, United States Public Health Service. Paper I of this series is R. R. Fisher et al. (1975)  相似文献   

7.
Licia N.Y. Wu  Ronald R. Fisher 《BBA》1982,681(3):388-396
Modification of pyridine dinucleotide transhydrogenase with tetranitromethane resulted in inhibition of its activity. Development of a membrane potential in submitochondrial particles during the reduction of 3-acetylpyridine adenine dinucleotide (AcPyAD+) by NADPH decreased to nearly the same extent as the transhydrogenase rate on tetranitromethane treatment of the membrane. Kinetics of the inactivation of homogeneous transhydrogenase and the enzyme reconstituted into phosphatidylcholine liposomes indicate that a single essential residue was modified per active monomer. NADP+, NADPH and NADH gave substantial protection against tetranitromethane inactivation of both the nonenergy-linked and energy-linked transhydrogenase reactions of submitochondrial particles and the NADPH → AcPyAD+ reaction of reconstituted enzyme. NAD+ had no effect on inactivation. Tetranitromethane modification of reconstituted transhydrogenase resulted in a decrease in the rate of coupled H+ translocation that was comparable to the decrease in the rate of NADPH → AcPyAD+ transhydrogenation. It is concluded that tetranitromethane modification controls the H+ translocation process solely through its effect on catalytic activity, rather than through alteration of a separate H+-binding domain. Nitrotyrosine was not found in tetranitromethane-treated transhydrogenase. Both 5,5′-dithiobis(2-nitrobenzoate)-accessible and buried sulfhydryl groups were modified with tetranitromethane. NADH and NADPH prevented sulfhydryl reactivity toward tetranitromethane. These data indicate that the inhibition seen with tetranitromethane results from the modification of a cysteine residue.  相似文献   

8.
The mechanism of the inactivation of 6-phosphogluconate dehydrogenase from Candida utilis with two coenzyme analogues can be differentiated on the basis of kinetic studies and of the properties of the inactivated enzyme. 3-Chloroacetylpyridine--adenine dinucleotide phosphate is clearly an affinity label and 3-choloroacetylpyridine--adenine dinucleotide a second-order reagent. For 3-chloroacetylpyridine--adenine dinucleotide phosphate, there is a loss of one thiol per subunit at complete inactivation whereas for 3-chloroacetylpyridine--adenine dinucleotide 2.7 thiol groups are lost. The fluorescence of the protein is quenched after alkylation by 3-chloroacetylpyridine--adenine dinucleotide phosphate and there is no quenching after the inactivation with 3-chloroacetylpyridine--adenine dinucleotide.  相似文献   

9.
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   

10.
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   

11.
Klaus Lendzian  James A. Bassham 《BBA》1975,396(2):260-275
The activity of glucose-6-phosphate dehydrogenase (EC 1.1.1.49) from spinach chloroplasts is strongly regulated by the ratio of NADPH/NADP+, with the extent of this regulation controlled by the concentration of ribulose 1,5-diphosphate. Other metabolites of the reductive pentose phosphate cycle are far less effective in mediating the regulation of the enzyme activity by NADPH/NADP+ ratio. With a ratio of NADPH/NADP+ of 2, and a concentration of ribulose 1,5-diphosphate of 0.6 mM, the activity of the enzyme is completely inhibited.This level of ribulose 1,5-diphosphate is well within the concentration range which has been reported for unicellular green algae photosynthesizing in vivo. Ratios of NADPH/NADP+ of 2.0 have been measured for isolated spinach chloroplasts in the light and under physiological conditions.Since ribulose 1,5-diphosphate is a metabolite unique to the reductive pentose phosphate cycle and inhibits glucose-6-phosphate dehydrogenase in the presence of NADPH/NADP+ ratios found in chloroplasts in the light, it is proposed that regulation of the oxidative pentose phosphate cycle is accomplished in vivo by the levels of ribulose 1,5-diphosphate, NADPH, and NADP+.It already has been shown that several key reactions of the reductive pentose phosphate cycle in chloroplasts are regulated by levels of NADPH/NADP+ or other electron-carrying cofactors, and at least one key-regulated step, the carboxylation reaction is strongly affected by 6-phosphogluconate, the metabolite unique to the oxidative pentose phosphate cycle. Thus there is an interesting inverse regulation system in chloroplasts, in which reduced/oxidized coenzymes provide a general regulatory mechanism. The reductive cycle is activated at high NADPH/NADP+ ratios where the oxidative cycle is inhibited, and ribulose 1,5-diphosphate and 6-phosphogluconate provide further control of the cycles, each regulating the cycle in which it is not a metabolite.  相似文献   

12.
A sonicate of Achromobacter parvulus IFO-13182 produced NADPH from NADP+by an NADP+-linked malic enzyme [l-malate: NAD(P)+oxidoreductase, EC 1.1.1.39–40] reaction in the presence of l-malic acid and divalent metal ions. Malic enzyme of A. parvulus was stabilized by 5% l-malic acid, and activity was maintained at 60°C for 1 h. Contaminating phosphatase (orthophosphoricmonoester phosphohydrolase, EC 3.1.3.1–2) was completely inactivated by this treatment. Among the conditions tested, the optimum NADPH production was done using 36 μmol NADP+, 67 μmol l-malic acid, 63 μmol MgCl2 and 1 unit of the malic enzyme in 3 ml of 55 mm phosphate buffer (pH 7.8). Conversion ratio of NADPH from NADP+ reached 100% after 4 h incubation at 30°C and the amount of NADPH accumulated was ~12 μmol ml?1of the reaction mixture. No dephosphorylation of NADP+to NAD+or of NADPH to NADH was found by high performance liquid chromatography. The NADPH produced by such enzymatic reduction was purified by ethanol precipitation and dried in vacuo in powdered form with 97% purity, judged from the ratio of the absorbances at 340 and 260 nm. The purity of the NADPH produced was determined to be 95% from its coenzyme activity with NAD(P)+-linked glutathione reductase [NAD(P)H: oxidized-glutathione oxidoreductase, EC 1.6.4.2].  相似文献   

13.
NADP+- linked isocitrate dehydrogenase has been purified to apparent homogeneity from 36 h germinated mung beans by ammonium sulphate fractionation, heat treatment, acid treatment, and DEAE — Cellulose column chromatography. The enzyme was purified to 150 fold with 15% recovery. The preparation showed single protein band on native PAGE and was free from bound nucleotides and coloured pigments (A280/A260 = 1.4). The molecular weight was found to be 141,000 and was made of four identical subunits (mol wt 36,000). Thermal inactivation at 50, 53, and 55 °C revealed simple first order kinetics and t1/2 was found to be 38, 10, and 3 min, respectively. The enzyme exhibited absolute specificity for NADP+ and substrate. The Km for isocitrate and NADP+ was 28.57 µM and 70 µM, respectively. The enzyme appeared to be regulated by various metabolites of Krebs’ cycle intermediates.  相似文献   

14.
《BBA》2014,1837(2):251-263
Ferredoxin-nicotinamide–adenine dinucleotide phosphate (NADP+) reductase (FNR) catalyses the production of reduced nicotinamide–adenine dinucleotide phosphate (NADPH) in photosynthetic organisms, where its flavin adenine dinucleotide (FAD) cofactor takes two electrons from two reduced ferredoxin (Fd) molecules in two sequential steps, and transfers them to NADP+ in a single hydride transfer (HT) step. Despite the good knowledge of this catalytic machinery, additional roles can still be envisaged for already reported key residues, and new features are added to residues not previously identified as having a particular role in the mechanism. Here, we analyse for the first time the role of Ser59 in Anabaena FNR, a residue suggested by recent theoretical simulations as putatively involved in competent binding of the coenzyme in the active site by cooperating with Ser80. We show that Ser59 indirectly modulates the geometry of the active site, the interaction with substrates and the electronic properties of the isoalloxazine ring, and in consequence the electron transfer (ET) and HT processes. Additionally, we revise the role of Tyr79 and Ser80, previously investigated in homologous enzymes from plants. Our results probe that the active site of FNR is tuned by a H-bond network that involves the side-chains of these residues and that results to critical optimal substrate binding, exchange of electrons and, particularly, competent disposition of the C4n (hydride acceptor/donor) of the nicotinamide moiety of the coenzyme during the reversible HT event.  相似文献   

15.
M F Carlier  D Pantaloni 《Biochemistry》1976,15(21):4703-4712
The binding of reduced nicotinamide adenine dinucleotide phosphate (NADPH) to nicotinamide adenine dinucleotide phosphate (NADP) dependent isocitrate dehydrogenase from beef liver cytoplasm was studied by several equilibrium techniques (ultracentrifugation, molecular sieving, ultrafiltration, fluorescence). Two binding sites (per dimeric enzyme molecule) were found with slightly different dissociation constants (0.5 and 0.12 muM) and fluorescence yields (7.7 and 6.3). A ternary complex was formed between enzyme, isocitrate, and NADPH, in which NADPH dissociation constant was 5 muM. On the contrary, no binding of NADPH to the enzyme took place in the presence of magnesium isocitrate. Dialysis experiments showed the existence of 1 NADP binding site/dimer, with a dissociation constant of 26 muM. When NADPH was present with the enzyme in the proportion of 1 molecule/dimer, the dissociation constant of NADP was decreased fourfold, reaching a value quantitatively comparable to the Michaelis constant. The kinetics of coenzyme binding was followed using the stopped-flow technique with fluorescence detection. NADPH binding to the enzyme occurred through one fast reaction (k1 = 20 muM-1 s-1). Dissociation of NADPH took place upon NADP binding; however, equilibrium as well as kinetic data were incompatible with a simple competition scheme. Dissociation of NADPH from the enzyme upon magnesium isocitrate binding was preceded by the formation of a transitory ternary complex in which the fluorescence of NADPH was only about 30% of that in the enzyme-NADPH complex. Then interaction between the conenzymes and the involvement of ternary complexes in the catalytic mechanism are discussed in relation with what is known about the regulatory role of the coenzyme (Carlier, M. F., and Pantaloni, D. (1976), Biochemistry, 15, 1761-1766).  相似文献   

16.
Dehydroepiandrosterone (DHEA) treatment of rats decreases gain of body weight without affecting food intake; simultaneously, the activities of liver malic enzyme and cytosolic glycerol-3-P dehydrogenase are increased. In the present study experiments were conducted to test the possibility that DHEA enhances thermogenesis and decreases metabolic efficiency via trans-hydrogenation of cytosolic NADPH into mitochondrial FADH2 with a consequent loss of energy as heat. The following results provide evidence which supports the proposed hypothesis: (a) the activities of cytosolic enzymes involved in NADPH production (malic enzyme, cytosolic isocitrate dehydrogenase, and aconitase) are increased after DHEA treatment; (b) cytosolic glycerol-3-P dehydrogenase may use both NAD+ and NADP+ as coenzymes; (c) activities of both cytosolic and mitochondrial forms of glycerol-3-P dehydrogenase are increased by DHEA treatment; (d) cytosol obtained from DHEA-treated rats synthesizes more glycerol-3-P during incubation with fructose-1,6-P2 (used as source of dihydroxyacetone phosphate) and NADP+; the addition of citratein vitro further increases this difference; (e) mitochondria prepared from DHEA-treated rats more rapidly consume glycerol-3-P added exogenously or formed endogenously in the cytosol in the presence of fructose-1,6-P2 and NADP+.  相似文献   

17.
We observed a spot on two-dimensional (2-D) gel in the epileptic mutant strain El mice with a similar molecular weight but with a different isoelectric point of approximately 0.2, compared with its mother strain ddY mice. The collected protein from the El mice was identified as cytosolic NADP+-dependent isocitrate dehydrogenase by internal amino acid sequencing. The enzyme is known to be maximally active during the development of the brain and to play an important role in NADPH production for fatty acids and cholesterol synthesis. In addition, alterations in cholesterol synthesis early in the development of the mammalian brain have been reported to lead to chronic epilepsy. The results in the present study therefore suggest that cytosolic NADP+-dependent isocitrate dehydrogenase might be involved in the epileptogenesis of the El mouse.  相似文献   

18.
The kinetic mechanism of the reaction catalyzed by glucose-6-phosphate dehydrogenase (EC 1.1.1.49) from Dicentrarchus labrax liver was examined using initial velocity studies,NADPH and glucosamine 6-phosphate inhibition and alternate coenzyme experiments. The results are consistent with a steady-state ordered sequential mechanism in which NADP+ binds first to the enzyme and NADPH is released last. Replots of NADPH inhibition show an uncommon parabolic pattern for this enzyme that has not been previously described. A kinetic model is proposed in agreement with our kinetic results and with previously published structural studies (Bautista et al. (1988) Biochem. Soc. Trans. 16, 903–904). The kinetic mechanism presented provides a possible explanation for the regulation of the enzyme by the [NADPH]/[NADP+] ratio.  相似文献   

19.
Arginyl residues in the NADPH-binding sites of phenol hydroxylase   总被引:1,自引:0,他引:1  
Phenol hydroxylase was inactivated by the arginine reagents 2,3-butanedione, 1,2-cyclohexanedione, and phenylglyoxal. The cosubstrate NADPH, as well as NADP+ and several analogues thereof, protected the enzyme against inactivation. Phenol did not protect the activity against any of the reagents used, nor did modification by 2,3-butanedione affect the binding of phenol. We propose the presence of arginyl residues in the binding sites for the adenosine phosphate part of NADPH.  相似文献   

20.
Malic enzyme (ME; NADP+-dependent; EC 1.1.40) provides NADPH for lipid biosynthesis in oleaginous microorganisms. Its role in vivo depends on there being an adequate supply of NADH to drive malate dehydrogenase to convert oxaloacetate to malate as a component of a cycle of three reactions: pyruvate → oxaloacetate → malate and, by the action of ME, back to pyruvate. However, the availability of cytosolic NADH is limited and, consequently, ancillary means of producing NADPH are necessary. Stoichiometries are given for the conversion of glucose to triacylglycerols involving ME with and without the reactions of the pentose phosphate pathway (PPP) as an additional source of NADPH. Some oleaginous microorganisms (such as Yarrowia lipolytica), however, lack a cytosolic ME and, if the PPP is the sole provider of NADPH, the theoretical yield of triacylglycerol from glucose falls to 27.6 % (w/w) from 31.6 % when ME is present. An alternative route for NADPH generation via a cytosolic isocitrate dehydrogenase (NADP+-dependent) is then discussed.  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