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.
Lipoamide dehydrogenase catalyzes the reversible NAD(+)-dependent oxidation of the dihydrolipoyl cofactors that are covalently attached to the acyltransferase components of the pyruvate dehydrogenase, alpha-ketoglutarate dehydrogenase, and glycine reductase multienzyme complexes. It contains two redox centers: a tightly, but noncovalently, bound FAD and an enzymic disulfide, each of which can accommodate two electrons. In the two-electron-reduced enzyme (EH(2)), the disulfide is reduced while the FAD cofactor is oxidized. In the four-electron-reduced enzyme (EH(4)), both redox centers are reduced. Lipoamide dehydrogenase can also catalyze the NADH-dependent reduction of alternative electron acceptors such as 2,6-dichlorophenolindophenol, ferricyanide, quinones, and molecular oxygen (O(2)). To determine the mechanism of these "diaphorase" reactions, we generated the EH(2) and EH(4) forms of Mycobacterium tuberculosis lipoamide dehydrogenase and rapidly mixed these enzyme forms with d,l-lipoylpentanoate, 2,6-dimethyl-1,4-benzoquinone, and O(2), in a stopped-flow spectrophotometer at pH 7.5 and 4 degrees C. EH(2) reduced d,l-lipoylpentanoate >/=100 times faster than EH(4) did. Conversely, EH(4) reduced 2,6-dimethyl-1,4-benzoquinone and molecular oxygen 90 and 40 times faster than EH(2), respectively. Comparison of the rates of reduction of the above substrates by EH(2) and EH(4) with their corresponding steady-state kinetic parameters for kinetic competence leads to the conclusion that reduction of lipoyl substrates occurs with EH(2) while reduction of diaphorase substrates occurs with EH(4).  相似文献   

2.
Treatment of submitochondrial particles (ETP) with trypsin at 0 degrees destroyed NADPH leads to NAD (or 3-acetylpyridine adenine dinucleotide, AcPyAD) transhydrogenase activity. NADH oxidase activity was unaffected; NADPH oxidase and NADH leads to AcPyAD transhydrogenase activities were diminished by less than 10%. When ETP was incubated with trypsin at 30 degrees, NADPH leads to NAD transhydrogenase activity was rapidly lost, NADPH oxidase activity was slowly destroyed, but NADH oxidase activity remained intact. The reduction pattern by NADPH, NADPH + NAD, and NADH of chromophores absorbing at 475 minus 510 nm (flavin and iron-sulfur centers) in complex I (NADH-ubiquinone reductase) or ETP treated with trypsin at 0 degrees also indicated specific destruction of transhydrogenase activity. The sensitivity of the NADPH leads to NAD transhydrogenase reaction to trypsin suggested the involvement of susceptible arginyl residues in the enzyme. Arginyl residues are considered to be positively charged binding sites for anionic substrates and ligands in many enzymes. Treatment of ETP with the specific arginine-binding reagent, butanedione, inhibited transhydrogenation from NADPH leads to NAD (or AcPyAD). It had no effect on NADH oxidation, and inhibited NADPH oxidation and NADH leads to AcPyAD transhydrogenation by only 10 to 15% even after 30 to 60 min incubation of ETP with butanedione. The inhibition of NADPH leads to NAD transhydrogenation was diminished considerably when butanedione was added to ETP in the presence of NAD or NADP. When both NAD and NADP were present, the butanedione effect was completely abolished, thus suggesting the possible presence of arginyl residues at the nucleotide binding site of the NADPH leads to NAD transhydrogenase enzyme. Under conditions that transhydrogenation from NADPH to NAD was completely inhibited by trypsin or butanedione, NADPH oxidation rate was larger than or equal to 220 nmol min-1 mg-1 ETP protein at pH 6.0 and 30 degrees. The above results establish that in the respiratory chain of beef-heart mitochondria NADH oxidation, NADPH oxidation, and NADPH leads to NAD transhydrogenation are independent reactions.  相似文献   

3.
Adrenodoxin stimulated the oxidation of NADPH by 1,4-benzoquinone, catalyzed by NADPH:adrenodoxin reductase. It prevented the enzyme inhibition by NADPH and formed an additional pathway of benzoquinone reduction presumably via reduced adrenodoxin. In the presence of 100-400 microM NADP+, which increased the Km of NADPH, adrenodoxin acted as a partial competitive inhibitor for NADPH decreasing its TN/Km by a limiting factor of 3. Ki of adrenodoxin decreased on the NADP+ concentration decrease and was estimated to be about 10(-8) M in the absence of NADP+.  相似文献   

4.
Glutathione reductase has been found to catalyze an NAD(P)H-dependent electron transfer to 2,4,6-trinitrobenzenesulfonate (TNBS). In the presence of oxygen TNBS is not consumed in the reaction, but is rapidly reoxidized with concomitant production of hydrogen peroxide. Cytochrome c can replace oxygen as the final electron acceptor, indicating that a one-electron transfer takes place. The rate is slightly higher in the absence than in the presence of oxygen, ruling out superoxide anion as an obligatory intermediate in cytochrome c reduction. In the absence of oxygen (or cytochrome c), TNBS limits the reaction and accepts a total of four electrons. The TNBS-dependent NADPH (or NADH) oxidation is markedly stimulated by NADP+, and to a smaller extent also by NAD+. The TNBS-dependent reactions are inhibited by excess of NADPH but not by NADH. The kinetics of these reactions are consistent with a branching reaction mechanism in which a pathway including a ternary complex between the two-electron reduced enzyme and NADP+ has the highest turnover. NADPH-dependent reductions of ferricyanide or 2,6-dichloroindophenol catalyzed by glutathione reductase are also markedly influenced by NADP+. Evidently NADP+ facilitates a shift of the catalyzed reaction from the normal two-electron reduction of glutathione disulfide to a more unspecific one-electron reduction of other acceptors. Spectral as well as kinetic data suggest that the rate of radical formation limits the reactions with the artificial electron acceptors and that NADP+ promotes this rate-limiting step.  相似文献   

5.
Properties of glutamate dehydrogenase purified from Bacteroides fragilis   总被引:2,自引:0,他引:2  
The dual pyridine nucleotide-specific glutamate dehydrogenase [EC 1.4.1.3] was purified 37-fold from Bacteroides fragilis by ammonium sulfate fractionation, DEAE-Sephadex A-25 chromatography twice, and gel filtration on Sephacryl S-300. The enzyme had a molecular weight of approximately 300,000, and polymeric forms (molecular weights of 590,000 and 920,000) were observed in small amounts on polyacrylamide gel disc electrophoresis. The molecular weight of the subunit was 48,000. The isoelectric point of the enzyme was pH 5.1. This glutamate dehydrogenase utilized NAD(P)H and NAD(P)+ as coenzymes and showed maximal activities at pH 8.0 and 7.4 for the amination with NADPH and with NADH, respectively, and at pH 9.5 and 9.0 for the deamination with NADP+ and NAD+, respectively. The amination activity with NADPH was about 5-fold higher than that with NADH. The Lineweaver-Burk plot for ammonia showed two straight lines in the NADPH-dependent reactions. The values of Km for substrates were: 1.7 and 5.1 mM for ammonium chloride, 0.14 mM for 2-oxoglutarate, 0.013 mM for NADPH, 2.4 mM for L-glutamate, and 0.019 mM for NADP+ in NADP-linked reactions, and 4.9 mM for ammonium chloride, 7.1 mM for 2-oxoglutarate, 0.2 mM for NADH, 7.3 mM for L-glutamate, and 3.0 mM for NAD+ in NAD-linked reactions. 2-Oxoglutarate and L-glutamate caused substrate inhibition in the NADPH- and NADP+-dependent reactions, respectively, to some extent. NAD+- and NADH-dependent activities were inhibited by 50% by 0.1 M NaCl. Adenine nucleotides and dicarboxylic acids did not show remarkable effects on the enzyme activities.  相似文献   

6.
L A Sheean  R A Meigs 《Steroids》1983,42(1):77-91
The ability of NADH to function as an alternative cofactor for the support of estrogen biosynthesis was validated. NADH supported rates of aromatization of up to 80% of those obtained with NADPH, with an apparent Km of 0.70 mM, and stimulated the NADPH-supported reaction only when supplies of the normal cofactor were limiting, both additive and synergistic effects being observed. NADH-supported aromatization was inhibited competitively by NADP+ and 2'-AMP with Ki values of 5 microM and 22 microM, respectively. Support by both cofactors was lost in parallel with the selective removal of NADPH-cytochrome c reductase from microsomes by graded subtilisin treatment. NADH-supported aromatization was differentiated from NADPH-supported aromatization by its sensitivity to inhibition by NAD+ and its response to changes in ionic strength. NADH appears to function, at high concentrations, as a surrogate for NADPH at the reduced nucleotide-binding site of NADPH-cytochrome c reductase but additional roles for NADH are also suggested both when acting alone and as a supplement to NADPH. A common oxidase (cytochrome P-450) appears to catalyze both NADH- and NADPH-supported aromatization.  相似文献   

7.
The goal was to determine whether endogenous cytosolic NAD(P)H:quinone oxidoreductase 1 (NQO1) preferentially uses NADPH or NADH in intact pulmonary arterial endothelial cells in culture. The approach was to manipulate the redox status of the NADH/NAD(+) and NADPH/NADP(+) redox pairs in the cytosolic compartment using treatment conditions targeting glycolysis and the pentose phosphate pathway alone or with lactate, and to evaluate the impact on the intact cell NQO1 activity. Cells were treated with 2-deoxyglucose, iodoacetate, or epiandrosterone in the absence or presence of lactate, NQO1 activity was measured in intact cells using duroquinone as the electron acceptor, and pyridine nucleotide redox status was measured in total cell KOH extracts by high-performance liquid chromatography. 2-Deoxyglucose decreased NADH/NAD(+) and NADPH/NADP(+) ratios by 59 and 50%, respectively, and intact cell NQO1 activity by 74%; lactate restored NADH/NAD(+), but not NADPH/NADP(+) or NQO1 activity. Iodoacetate decreased NADH/NAD(+) but had no detectable effect on NADPH/NADP(+) or NQO1 activity. Epiandrosterone decreased NQO1 activity by 67%, and although epiandrosterone alone did not alter the NADPH/NADP(+) or NADH/NAD(+) ratio, when the NQO1 electron acceptor duroquinone was also present, NADPH/NADP(+) decreased by 84% with no impact on NADH/NAD(+). Duroquinone alone also decreased NADPH/NADP(+) but not NADH/NAD(+). The results suggest that NQO1 activity is more tightly coupled to the redox status of the NADPH/NADP(+) than NADH/NAD(+) redox pair, and that NADPH is the endogenous NQO1 electron donor. Parallel studies of pulmonary endothelial transplasma membrane electron transport (TPMET), another redox process that draws reducing equivalents from the cytosol, confirmed previous observations of a correlation with the NADH/NAD(+) ratio.  相似文献   

8.
The relatively slow reduction of NAD+ and NADP+ by sodium cyanoborohydride leads to formation of the enzymically active form of NADH and NADPH. This reaction could be useful as a simple procedure to enzymically introduce a specific label into substrates when tritiated or deuterated cynanoborohydride is used for obtaining the reduced nicotinamide adenine dinucleotide.  相似文献   

9.
We studied the physiological effect of the interconversion between the NAD(H) and NADP(H) coenzyme systems in recombinant Saccharomyces cerevisiae expressing the membrane-bound transhydrogenase from Escherichia coli. Our objective was to determine if the membrane-bound transhydrogenase could work in reoxidation of NADH to NAD+ in S. cerevisiae and thereby reduce glycerol formation during anaerobic fermentation. Membranes isolated from the recombinant strains exhibited reduction of 3-acetylpyridine-NAD+ by NADPH and by NADH in the presence of NADP+, which demonstrated that an active enzyme was present. Unlike the situation in E. coli, however, most of the transhydrogenase activity was not present in the yeast plasma membrane; rather, the enzyme appeared to remain localized in the membrane of the endoplasmic reticulum. During anaerobic glucose fermentation we observed an increase in the formation of 2-oxoglutarate, glycerol, and acetic acid in a strain expressing a high level of transhydrogenase, which indicated that increased NADPH consumption and NADH production occurred. The intracellular concentrations of NADH, NAD+, NADPH, and NADP+ were measured in cells expressing transhydrogenase. The reduction of the NADPH pool indicated that the transhydrogenase transferred reducing equivalents from NADPH to NAD+.  相似文献   

10.
Transhydrogenase couples the redox reaction between NAD(H) and NADP(H) to proton translocation across a membrane. Coupling is achieved through changes in protein conformation. Upon mixing, the isolated nucleotide-binding components of transhydrogenase (dI, which binds NAD(H), and dIII, which binds NADP(H)) form a catalytic dI(2).dIII(1) complex, the structure of which was recently solved by x-ray crystallography. The fluorescence from an engineered Trp in dIII changes when bound NADP(+) is reduced. Using a continuous flow device, we have measured the Trp fluorescence change when dI(2).dIII(1) complexes catalyze reduction of NADP(+) by NADH on a sub-millisecond scale. At elevated NADH concentrations, the first-order rate constant of the reaction approaches 21,200 s(-1), which is larger than that measured for redox reactions of nicotinamide nucleotides in other, soluble enzymes. Rather high concentrations of NADH are required to saturate the reaction. The deuterium isotope effect is small. Comparison with the rate of the reverse reaction (oxidation of NADPH by NAD(+)) reveals that the equilibrium constant for the redox reaction on the complex is >36. This high value might be important in ensuring high turnover rates in the intact enzyme.  相似文献   

11.
NADH acts as an incomplete competitive inhibitor for 5,8-dioxy-1,4-naphtoquinone during its rotenone-insensitive reduction by mitochondrial NADH:ubiquinone reductase. NAD+ and ADP-ribose act as incomplete mixed-type inhibitors. Ki of NAD+ and NADH towards quinone are about one order less than towards ferricyanide. The bimolecular rate constant of the reduction of the enzyme by NADH in the quinone reductase reaction is about 2 times less than that of ferricyanide reductase reaction. These data indicate that the reduction site of 5,8-dioxy-1,4-naphtoquinone is close to NAD+/NADH and ferricyanide binding site. It seems that during the steady-state reduction of ferricyanide and 5,8-dioxy-1,4-naphtoquinone these oxidizers react with NADH:ubiquinone reductase reduced to different extents.  相似文献   

12.
Glucose dehydrogenase from rat liver microsomes was found to react not only with glucose as a substrate but also with glucose 6-phosphate, 2-deoxyglucose 6-phosphate and galactose 6-phosphate. The relative maximum activity of this enzyme was 29% for glucose 6-phosphate, 99% for 2-deoxyglucose 6-phosphate, and 25% for galactose 6-phosphate, compared with 100% for glucose with NADP. The enzyme could utilize either NAD or NADP as a coenzyme. Using polyacrylamide gradient gel electrophoresis, we were able to detect several enzymatically active bands by incubation of the gels in a tetrazolium assay mixture. Each band had different Km values for the substrates (3.0 x 10(-5)M glucose 6-phosphate with NADP to 2.4M glucose with NAD) and for coenzymes (1.3 x 10(-6)M NAD with galactose 6-phosphate to 5.9 x 10(-5)M NAD with glucose). Though glucose 6-phosphate and galactose 6-phosphate reacted with glucose dehydrogenase, they inhibited the reaction of this enzyme only when either glucose or 2-deoxyglucose 6-phosphate was used as a substrate. The Ki values for glucose 6-phosphate with glucose as substrate were 4.0 x 10(-6)M with NAD, and 8.4 x 10(-6)M with NADP; for galactose 6-phosphate they were 6.7 x10(-6)M with NAD and 6.0 x 10(-6)M with NADP. The Ki values for glucose 6-phosphate with 2-deoxyglucose 6-phosphate as substrate were 6.3 x 10(-6)M with NAD and 8.9 x 10(-6)M with NADP; and for galactose 6-phosphate, 8.0 x 10(-6)M with NAD and 3.5 x 10(-6)M with NADP. Both NADH and NADPH inhibited glucose dehydrogenase when the corresponding oxidized coenzymes were used (Ki values: 8.0 x 10(-5)M by NADH and 9.1 x 10(-5)M by NADPH), while only NADPH inhibited cytoplasmic glucose 6-phosphate dehydrogenase (Ki: 2.4 x 10(-5)M). The results indicate that glucose dehydrogenase cannot directly oxidize glucose in vivo, but it might play a similar role to glucose 6-phosphate dehydrogenase. The differences in the kinetics of glucose dehydrogenase and glucose 6-phosphate dehydrogenase show that glucose 6-phosphate and galactose 6-phosphate could be metabolized in quite different ways in the microsomes and cytoplasm of rat liver.  相似文献   

13.
The content of NADH and NADPH was measured in the intact and regenerating rat liver. In the intact rat liver, the content of NAD+, NADH, NADP+ and NADPH was 235 +/- 6.4, 66.6 +/- 4.3, 73.3 +/- 2.5 and 148.0 +/- 4.6 micrograms/g crude liver weight, respectively. Seasonal alterations in the rat liver content of coenzymes were established. No changes were found in the content of nicotinamide coenzymes in the regenerating liver 4 and 18 h after operation. Twenty-four hours after operation, a 25.6% increase in the content of NAD+ and a 57.8% reduction in the NADH content were recorded in the liver of hepatectomized animals. At the same time the total content of NAD+ plus NADH changed but insignificantly (14.7%). The total content of NADP+ plus NADPH dropped by 29.8% (within the above period). Thirty-two hours after operation the content of all the nicotinamide coenzymes returned to the initial level.  相似文献   

14.
Slices of hypocotyls from 3-day-old seedlings of Vigna sesquipedalis (L.) Fruwirth in the germination stage were incubated under various gaseous conditions. The NADP+NADPH level in the hypocotyl slices changed with the oxygen tension. A high NADP+NADPH level was observed under aerobic conditions and a low NADP+NADPH level under anaerobic conditions.

The 100 × NADH/NAD+NADH ratio increased greatly under anaerobic conditions. In general a low NADP + NADPH level corresponded with a high 100 × NADH/NAD+NADH ratio. On the basis of the results given in the following paper, it was discussed that the slowness of NADH oxidation in hypocotyl tissue due to anaerobic conditions results in the inhibition of NADP formation.

The variation of the NADP+NADPH level was considered to produce a modification of the carbohydrate metabolism.

The NADP+NADPH level in E. coli cells suspended in glucose solution also changed with the oxygen tension.

  相似文献   

15.
Kinetic measurements indicate that the energy-independent transhydrogenation of 3-acetylpyridine-NAD+ by NADPH in membranes of Escherichia coli follows a rapid equilibrium random bireactant mechanism. Each substrate, although reacting preferentially with its own binding site, is able to interact with the binding site of the other substrate to cause inhibition of enzyme activity. 5'-AMP (and ADP) and 2'-AMP interact with the NAD+- and NADP+-binding sites, respectively. Phenylglyoxal and 2,3-butanedione in borate buffer inhibit transhydrogenase activity presumably by reacting with arginyl residues. Protection against inhibition by 2,3-butanedione is afforded by NADP+, NAD+, and high concentrations of NADPH and NADH. Low concentrations of NADPH and NADH increase the rate of inhibition by 2,3-butanedione. Similar effects are observed for the inactivation of the transhydrogenase by tryptic digestion in the presence of these coenzymes. It is concluded that there are at least two conformations of the active site of the transhydrogenase which differ in the extent to which arginyl residues are accessible to exogenous agents such as trypsin and 2,3-butanedione. One conformation is induced by low concentrations of NADH and NADPH. Under these conditions the coenzymes could be reacting at the active site or at an allosteric site. The stimulation of transhydrogenase activity by low concentrations of the NADH is consistent with the latter possibility.  相似文献   

16.
NAD+ reduction catalyzed by transhydrogenase (EC 1.6.1.1) from E. coli membrane particles at the expense of NADPH oxidation is coupled with phenyldicarbaundecaborate (PCB-) absorption by the particles. This process is inhibited by oxidative phosphorylation protonophorous uncouplers and by equilibration of concentrations of the substrates and products of the transhydrogenase reaction. Elimination of the water-soluble part of membrane ATPase results in the inhibition of PCB- absorption at the expense of the transhydrogenase reaction energy. Treatment of the particles by dicyclohexyl carbodiimide increases the transhydrogenase-coupled absorption of PCB-. The transhydrogenase-induced increase of pPCB in the suspension of particles is directly correlated with the ratio of ([NADPH].[NAD+])/([NADP+].[NADH]). When this value is equal to 1, no energy-dependent increase of pPCB was observed. NADP+ reduction at the expense of NADH oxidation leads to a decrease in the amount of PCB- absorbed by the particles at the expense of ATP hydrolysis energy. The experimental data suggest that NADPH oxidation in the course of the transhydrogenase reaction is coupled with the formation of a membrane potential with a positive charge localized inside the particles.  相似文献   

17.
NAD kinase was purified 93-fold from Escherichia coli. The enzyme was found to have a pH optimum of 7.2 and an apparent Km for NAD+, ATP, and Mg2+ of 1.9, 2.1, and 4.1 mM, respectively. Several compounds including quinolinic acid, nicotinic acid, nicotinamide, nicotinamide mononucleotide, AMP, ADP, and NADP+ did not affect NAD kinase activity. The enzyme was not affected by changes in the adenylate energy charge. In contrast, both NADH and NADPH were potent negative modulators of the enzyme, since their presence at micromolar concentrations resulted in a pronounced sigmoidal NAD+ saturation curve. In addition, the presence of a range of concentrations of the reduced nucleotides resulted in an increase of the Hill slope (nH) to 1.7 to 2.0 with NADH and to 1.8 to 2.1 with NADPH, suggesting that NAD kinase is an allosteric enzyme. These results indicate that NAD kinase activity is regulated by the availability of ATP, NAD+, and Mg2+ and, more significantly, by changes in the NADP+/NADPH and NAD+/NADH ratios. Thus, NAD kinase probably plays a role in the regulation of NADP turnover and pool size in E. coli.  相似文献   

18.
The interactions of calcium with NAD+, NADH, NADP+ and NADPH in a 50% (by volume) methanol/water mixture (pH 7, 25 degrees C) were studied by calorimetry. The association constants for 1:1 complex formation were found to be 6.6 +/- 0.2, 270 +/- 76, 18 +/- 3 and 98 +/- 10 for NAD+, NADH, NADP+ and NADPH, respectively. Comparing these to the association constants for an aqueous system reveals that as the polarity of the solvent system is decreased the interactions involving NAD+, NADP+ and NADPH are all decreased. In contrast, the interaction involving NADH is markedly increased. All the interactions were found to be endothermic.  相似文献   

19.
Regulation of NAD- and NADP-dependent isocitrate dehydrogenases (NAD-ICDH, EC 1.1.1.41, and NADP-ICDH, EC 1.1.1.42) by the level of reduced and oxidized pyridine nucleotides has been investigated in pea (Pisum sativum L.) leaves. The affinities of mitochondrial and cytosolic ICDH enzymes to substrates and inhibitors were determined on partially purified preparations in forward and reverse directions. From the kinetic data, it follows that NADP(+)- and NAD(+)-dependent isocitrate dehydrogenases in mitochondria represent a system strongly responding to the intramitochondrial NADPH and NADH levels. The NADPH, NADP(+), NADH and NAD(+) concentrations were determined by subcellular fractionation of pea leaf protoplasts using membrane filtration in mitochondria and cytosol in darkness and in the light under saturating and limiting CO(2) conditions. The cytosolic NADPH/NADP ratio was about 1 and almost constant both in darkness and in the light. In mitochondria, the NADPH/NADP ratio was low in darkness (0.2) and increased in the light, reaching 3 in limiting CO(2) conditions compared to 1 in saturating CO(2). At high reduction levels of NADP and NAD observed at limiting CO(2) in the light, i.e. when photorespiratory glycine is the main mitochondrial substrate, isocitrate oxidation in mitochondria will be suppressed and citrate will be transported to the cytosol ('citrate valve'), where the cytosolic NADP-ICDH supplies 2-oxoglutarate for the photorespiratory ammonia refixation.  相似文献   

20.
Transhydrogenase couples the redox reaction between NADH and NADP+ to proton translocation across a membrane. The enzyme comprises three components; dI binds NAD(H), dIII binds NADP(H), and dII spans the membrane. The 1,4,5,6-tetrahydro analogue of NADH (designated H2NADH) bound to isolated dI from Rhodospirillum rubrum transhydrogenase with similar affinity to the physiological nucleotide. Binding of either NADH or H2NADH led to closure of the dI mobile loop. The 1,4,5,6-tetrahydro analogue of NADPH (H2NADPH) bound very tightly to isolated R. rubrum dIII, but the rate constant for dissociation was greater than that for NADPH. The replacement of NADP+ on dIII either with H2NADPH or with NADPH caused a similar set of chemical shift alterations, signifying an equivalent conformational change. Despite similar binding properties to the natural nucleotides, neither H2NADH nor H2NADPH could serve as a hydride donor in transhydrogenation reactions. Mixtures of dI and dIII form dI2dIII1 complexes. The nucleotide charge distribution of complexes loaded either with H2NADH and NADP+ or with NAD+ and H2NADPH should more closely mimic the ground states for forward and reverse hydride transfer, respectively, than previously studied dead-end species. Crystal structures of such complexes at 2.6 and 2.3 A resolution are described. A transition state for hydride transfer between dihydronicotinamide and nicotinamide derivatives determined in ab initio quantum mechanical calculations resembles the organization of nucleotides in the transhydrogenase active site in the crystal structure. Molecular dynamics simulations of the enzyme indicate that the (dihydro)nicotinamide rings remain close to a ground state for hydride transfer throughout a 1.4 ns trajectory.  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