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 46 毫秒
1.
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   

2.
The FAD-dependent NAD(P)H oxidase from Lactobacillus sanfrancisensis (L.san-Nox2) catalyzes the oxidation of 2 equivalents of either NADH or NADPH and reduces 1 equivalent of O(2) to yield 2 equivalents of water. During steady-state turnover only 0.5% of the reducing equivalents are detected in solution as hydrogen peroxide, suggesting that it is not released from the enzyme after the oxidation of the first equivalent of NAD(P)H and reaction with O(2). Here we report the crystal structure of L.san-Nox2 to 1.8 A resolution. The enzyme crystallizes as a dimer with each monomer consisting of a FAD binding domain (residues 1-120), a NAD(P)H binding domain (residues 150-250), and a dimerization domain (residues 325-451). The electron density for the redox-active Cys42 residue located adjacent to the si-face FAD is consistent with oxidation to the sulfenic acid (Cys-SOH) state. The side chain of Cys42 is also observed in two conformations; in one the sulfenic acid is hydrogen bonded to His10 and in the other it hydrogen bonds with the FAD O2' atom. Surprisingly, the NAD(P)H binding domains each contain an ADP ligand as established by electron density maps and MALDI-TOF analysis of the ligands released from heat-denatured enzyme. The ADP ligand copurifies with the enzyme, and its presence does not inhibit enzyme activity. Consequently, we hypothesize that either NADPH or NADH substrates bind via a long channel that extends from the enzyme exterior and terminates at the FAD re-face. A homology model of the NADH oxidase from Lactococcus lactis (L.lac-Nox2) was also generated using the crystal structure of L.san-Nox2, which reveals several important similarities and differences between the two enzymes. HPLC analysis of ligands released from denatured L.lac-Nox2 indicates that it does not bind ADP, which correlates with the specificity of the enzyme for oxidation of NADH.  相似文献   

3.
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   

4.
To elucidate the structural basis for the alteration of coenzyme specificity from NADH toward NADPH in a malate dehydrogenase mutant EX7 from Thermus flavus, we determined the crystal structures at 2.0 A resolution of EX7 complexed with NADPH and NADH, respectively. In the EX7-NADPH complex, Ser42 and Ser45 form hydrogen bonds with the 2'-phosphate group of the adenine ribose of NADPH, although the adenine moiety is not seen in the electron density map. In contrast, although Ser42 and Ser45 occupy a similar position in the EX7-NADH complex structure, both the adenine and adenine ribose moieties of NADH are missing in the map. These results and kinetic analysis of site-directed mutant enzymes indicate (1) that the preference of EX7 for NADPH over NADH is ascribed to the recognition of the 2'-phosphate group by two Ser and Arg44, and (2) that the adenine moiety of NADPH is not recognized in this mutant.  相似文献   

5.
Type II NADH dehydrogenase of Corynebacterium glutamicum (NDH-2) was purified from an ndh overexpressing strain. Purification conferred 6-fold higher specific activity of NADH:ubiquinone-1 oxidoreductase with a 3.5-fold higher recovery than that previously reported (K. Matsushita et al., 2000). UV-visible and fluorescence analyses of the purified enzyme showed that NDH-2 of C. glutamicum contained non-covalently bound FAD but not covalently bound FMN. This enzyme had an ability to catalyze electron transfer from NADH and NADPH to oxygen as well as various artificial quinone analogs at neutral and acidic pHs respectively. The reduction of native quinone of C. glutamicum, menaquinone-2, with this enzyme was observed only with NADH, whereas electron transfer to oxygen was observed more intensively with NADPH. This study provides evidence that C. glutamicum NDH-2 is a source of the reactive oxygen species, superoxide and hydrogen peroxide, concomitant with NADH and NADPH oxidation, but especially with NADPH oxidation. Together with this unique character of NADPH oxidation, phylogenetic analysis of NDH-2 from various organisms suggests that NDH-2 of C. glutamicum is more closely related to yeast or fungal enzymes than to other prokaryotic enzymes.  相似文献   

6.
Two different NAD(P)H dehydrogenases could be demonstrated in the blue-green alga, Aphanocapsa. Both function as quinone reductases using benzoquinone as electron acceptor. One, which was found in the soluble fraction, was NADH specific and showed high sensitivity to rotenone, thenoyltrifluoroacetone and o-phenanthroline. The second dehydrogenase was membrane-bound and used NADH as well as NADPH as substrates. Inhibition by rotenone and o-phenanthroline was less pronounced with the bound enzyme than with the soluble enzyme. Based on studies with NADH or NADPH, the membrane-bound enzyme apparently was associated with a low-temperature EPR signal at g=1.92 in the reduced state, indicative of an iron-sulfur center. The membrane-bound dehydrogenase was solubilized with Triton X-100 and partially purified. This preparation was used for studies of enzyme kinetics and acceptor specificity.Abbreviations DBMIB 2,5-dibromo methyl isopropylbenzoquinone - TTFA thenoyltrifluoroacetone - E m midpoint redox potential  相似文献   

7.
Quantum-mechanical electron density calculations reveal that a significant polarization is induced in the cofactor NADPH (reduced nicotinamide adenine dinucleotide phosphate) on binding to the enzyme dihydrofolate reductase. The calculations indicate that electron density corresponding to approximately 0.7 electron charges is shifted within the molecule, extending over more than 20 A. Further calculations on proposed enzyme mutants show that the polarization of NADPH on binding to DHFR is, in large part, induced by a motif of three positively charged residues. This motif was also identified to be directly responsible for the positive electrostatic potential surrounding the cofactor binding site in the enzyme. The possibility of this long-range polarization of NADPH was originally proposed based on a previous study of ligand binding to DHFR where a conserved structural motif of three positively charged residues was found to play a major role in polarizing the substrate folate over its entire length of 18 A.  相似文献   

8.
G F Leanz  G G Hammes 《Biochemistry》1986,25(19):5617-5624
The ionic strength dependence of the second-order rate constant for the association of reduced nicotinamide adenine dinucleotide phosphate (NADPH) and chicken liver fatty acid synthase was determined. This rate constant is 7.2 X 10(7) M-1 s-1 at zero ionic strength and 25 degrees C; the effective charge at the cofactor binding sites is +0.8. The conformations of nicotinamide adenine dinucleotide phosphate (NADP+) and NADPH bound to the beta-ketoacyl and enoyl reductase sites were determined from transferred nuclear Overhauser effect measurements. Covalent modification of the enzyme with pyridoxal 5'-phosphate abolished cofactor binding at the enoyl reductase site; this permitted the cofactor conformations at the beta-ketoacyl and enoyl reductase sites to be distinguished. For NADP+ bound to the enzyme, the conformation of the nicotinamide-ribose bond is anti at the enoyl reductase site and syn at the beta-ketoacyl reductase site; the adenine-ribose bond is anti, and the sugar puckers are C3'-endo. Nicotinamide-adenine base stacking was not detected. Structural models of NADP+ at the beta-ketoacyl and enoyl reductase sites were constructed by using the distances calculated from the observed nuclear Overhauser effects. Because of the overlap of the resonances of several nonaromatic NADPH protons with the resonances of HDO and ribose protons, less extensive structural information was obtained for NADPH bound to the enzyme. However, the conformations of NADPH bound to the two reductases are qualitatively the same as those of NADP+, except that the nicotinamide moiety of NADPH is closer to being fully anti at the enoyl reductase site.  相似文献   

9.
The reduced nicotinamide adenine dinucleotide phosphate (NADPH) is pivotal to the cellular anti-oxidative defence strategies in most organisms. Although its production mediated by different enzyme systems has been relatively well-studied, metabolic networks dedicated to the biogenesis of NADPH have not been fully characterized. In this report, a metabolic pathway that promotes the conversion of reduced nicotinamide adenine dinucleotide (NADH), a pro-oxidant into NADPH has been uncovered in Pseudomonas fluorescens exposed to oxidative stress. Enzymes such as pyruvate carboxylase (PC), malic enzyme (ME), malate dehydrogenase (MDH), malate synthase (MS), and isocitrate lyase (ICL) that are involved in disparate metabolic modules, converged to create a metabolic network aimed at the transformation of NADH into NADPH. The downregulation of phosphoenol carboxykinase (PEPCK) and the upregulation of pyruvate kinase (PK) ensured that this metabolic cycle fixed NADH into NADPH to combat the oxidative stress triggered by the menadione insult. This is the first demonstration of a metabolic network invoked to generate NADPH from NADH, a process that may be very effective in combating oxidative stress as the increase of an anti-oxidant is coupled to the decrease of a pro-oxidant.  相似文献   

10.
The 2',3'-dialdehyde nicotinamide ribose derivatives of NAD (oNAD) and NADH (oNADH) have been prepared enzymatically from the corresponding 2',3'-dialdehyde analogs of NADP and NADPH. Pig heart NAD-dependent isocitrate dehydrogenase requires NAD as coenzyme but binds NADPH, as well as NADH, ADP, and ATP, at regulatory sites. Incubation of 1-3 mM oNAD or oNADH with this isocitrate dehydrogenase causes a time-dependent decrease in activity to a limiting value 40% that of the initial enzyme, suggesting that reaction does not occur at the catalytic coenzyme site. Upon varying the concentration of oNAD or oNADH from 0.2 to 3 mM, the inactivation rate constants increase in a nonlinear manner, consistent with reversible binding of oNAD and oNADH to the enzyme prior to covalent reaction. Inactivation is accompanied by incorporation of radioactive reagent with extrapolation to 0.54 mol [14C]oNAD or 0.45 mol [14C]oNADH/mol average enzyme subunit (or about 2 mol reagent/mol enzyme tetramer) when the enzyme is maximally inactivated; this value corresponds to the number of reversible binding sites for each of the natural ligands of isocitrate dehydrogenase. The protection against oNAD or oNADH inactivation by NADH, NADPH, and ADP (but not by isocitrate, NAD, or NADP) indicates that reaction occurs in the region of a nucleotide regulatory site. In contrast to the effects of oNAD and oNADH, oNADP and oNADPH cause total inactivation of the NAD-dependent isocitrate dehydrogenase, concomitant with incorporation, respectively, of about 3.5 mol [14C]oNADP or 1.3 mol [14C]oNADPH/mol average subunit. Reaction rates exhibit a linear dependence on [oNADP] or [oNADPH] and protection by natural ligands against inactivation is not striking. These results imply that oNADP and oNADPH are acting in this case as general chemical modifiers and indicate the importance of the free adenosine 2'-OH of oNAD and oNADH for specific labeling of the NAD-dependent isocitrate dehydrogenase. The new availability of 2',3'-dialdehyde nicotinamide ribose derivatives of NAD, NADH, NADP, and NADPH may allow selection of the appropriate reactive coenzyme analog for affinity labeling of a variety of dehydrogenases.  相似文献   

11.
The reduced nicotinamide adenine dinucleotide phosphate (NADPH) is pivotal to the cellular anti-oxidative defence strategies in most organisms. Although its production mediated by different enzyme systems has been relatively well-studied, metabolic networks dedicated to the biogenesis of NADPH have not been fully characterized. In this report, a metabolic pathway that promotes the conversion of reduced nicotinamide adenine dinucleotide (NADH), a pro-oxidant into NADPH has been uncovered in Pseudomonas fluorescens exposed to oxidative stress. Enzymes such as pyruvate carboxylase (PC), malic enzyme (ME), malate dehydrogenase (MDH), malate synthase (MS), and isocitrate lyase (ICL) that are involved in disparate metabolic modules, converged to create a metabolic network aimed at the transformation of NADH into NADPH. The downregulation of phosphoenol carboxykinase (PEPCK) and the upregulation of pyruvate kinase (PK) ensured that this metabolic cycle fixed NADH into NADPH to combat the oxidative stress triggered by the menadione insult. This is the first demonstration of a metabolic network invoked to generate NADPH from NADH, a process that may be very effective in combating oxidative stress as the increase of an anti-oxidant is coupled to the decrease of a pro-oxidant.  相似文献   

12.
A gene encoding an sn‐glycerol‐1‐phosphate dehydrogenase (G1PDH) was identified in the hyperthermophilic archaeon Pyrobaculum calidifontis. The gene was overexpressed in Escherichia coli, and its product was purified and characterized. In contrast to conventional G1PDHs, the expressed enzyme showed strong preference for NADH: the reaction rate (Vmax) with NADPH was only 2.4% of that with NADH. The crystal structure of the enzyme was determined at a resolution of 2.45 Å. The asymmetric unit consisted of one homohexamer. Refinement of the structure and HPLC analysis showed the presence of the bound cofactor NADPH in subunits D, E, and F, even though it was not added in the crystallization procedure. The phosphate group at C2’ of the adenine ribose of NADPH is tightly held through the five biased hydrogen bonds with Ser40 and Thr42. In comparison with the known G1PDH structure, the NADPH molecule was observed to be pushed away from the normal coenzyme binding site. Interestingly, the S40A/T42A double mutant enzyme acquired much higher reactivity than the wild‐type enzyme with NADPH, which suggests that the biased interactions around the C2’‐phosphate group make NADPH binding insufficient for catalysis. Our results provide a unique structural basis for coenzyme preference in NAD(P)‐dependent dehydrogenases. Proteins 2016; 84:1786–1796. © 2016 Wiley Periodicals, Inc.  相似文献   

13.
The nicotinamide adenine dinucleotide (NADH)/nicotinamide adenine dinucleotide phosphate (NADPH) oxidase and the xanthine oxidase (XOD) systems generate reactive oxygen species (ROS). In the present study, to characterize the difference between the two systems, the kinetics of ROS generated by both the NADH oxidase and XOD systems were analysed by an electron spin resonance (ESR) spin trapping method using 5,5-dimethyl-1-pyrroline-N-oxide (DMPO), 5-(diethoxyphosphoryl)-5-methyl-pyrroline N-oxide (DEPMPO) and 5-(2,2-dimethyl-1,3-propoxy cyclophosphoryl)-5-methyl-1-pyrroline N-oxide (CYPMPO). As a result, two major differences in ROS kinetics were found between the two systems: (i) the kinetics of (?)OH and (ii) the kinetics of hydrogen peroxide. In the NADH oxidase system, the interaction of hydrogen peroxide with each component of the enzyme system (NADPH, NADH oxidase and FAD) was found to generate (?)OH. In contrast, (?)OH generation was found to be independent of hydrogen peroxide in the XOD system. In addition, the hydrogen peroxide level in the NADPH-NADH oxidase system was much lower than measured in the XOD system. This lower level of free hydrogen peroxide is most likely due to the interaction between hydrogen peroxide and NADPH, because the hydrogen peroxide level was reduced by ~90% in the presence of NADPH.  相似文献   

14.
Crystalline complexes of tyrosyl tRNA synthetase were prepared with the following substrates and substrate analogues: ATP, AMP, α-β methylene ATP, tyrosine and tyrosinyl adenylate. Using 14C-labelled ligands, the binding constants for tyrosine and ATP to crystals were shown to be similar to those observed in solution. Two tyrosine molecules were found to bind to the symmetrical dimer in the crystalline enzyme, while only one tyrosine binds with high affinity in solution. Electron density difference maps show that tyrosine and the AMP derivatives all bind at the same site, in a cleft 10 Å deep at one side of the pleated sheet, tyrosine binding over 100 times more strongly. The phosphate groups of AMP and ATP are not unambiguously observed in the difference electron density maps. Tyrosinyl adenylate is clearly delineated in the electron density difference map, with the tyrosyl side-chain occupying the site previously observed. The adenosine group is in a wide cup-like depression outside the pocket, lying between the carboxyl-terminal continuations of strands 3 and 5 of the pleated sheet. The adenine ring is lying against an α-helix. The binding of tyrosinyl adenylate causes no detectable conformational changes of the enzyme.  相似文献   

15.
Nitric oxide reductase (Nor) cytochrome P450nor (P450nor) is unique because it is catalytically self-sufficient, receiving electrons directly from NADH or NADPH. However, little is known about the direct binding of NADH to cytochrome. Here, we report that oxidized pyridine nucleotides (NAD(+) and NADP(+)) and an analogue induce a spectral perturbation in bound heme when mixed with P450nor. The P450nor isoforms are classified according to electron donor specificity for NADH or NADPH. One type (Fnor, a P450nor of Fusarium oxysporum) utilizes only NADH. We found that NAD(+) induced a type I spectral change in Fnor, whereas NADP(+) induced a reverse type I spectral change, although the K(d) values for both were comparable. In contrast, NADP(+) as well as NAD(+) caused a type I spectral change in Tnor, a P450nor isozyme from Trichosporon cutaneum that utilizes both NADH and NADPH as electron donors. The B' helix region of Tnor ((73)SAGGKAAA(80)) contains some Ala and Gly residues, whereas the sequence is replaced at a few sites with more bulky amino acid residues in Fnor ((73)SASGKQAA(80)). A single mutation (S75G) significantly improved the NADPH- dependent Nor activity of Fnor, and the overall activity was accelerated via the NADPH-enhanced reduction step. These results showed that pyridine nucleotide cofactors can bind P450nor and that only a few residues in the B' helix region determine cofactor specificity. We further showed that a poor electron donor (NADPH) could also bind Fnor, but an appropriate configuration for electron transfer is blocked by steric hindrance mainly by Ser(75) against the 2'-phosphate moiety. The present results along with previous observations together revealed a novel motif for cofactor binding.  相似文献   

16.
The cytosolic component of macrophage-derived superoxide generating NADPH oxidase was partially purified by affinity chromatography on 2′,5′-ADP-agarose. Elution was nonspecific by elevated phosphate molarity. A single step attains at least 40-fold enrichment of specific activity, the recovery being over 20%. Elution with various ligands in the concentration range 2–3.5 mM was also tested. The most effective ligands were: ATP, dATP, GTP, NADPH and 2′,5′-ADP. Ineffective were AMP, 2′-AMP, FMN, FAD and NADH. ADP was of medium potency. On the basis of the above and other results, we infer that the molecule (or complex) purified by us may contain the enzymatic NADPH binding site. This component is fully retained by a 100 kDa cutoff membrane and is labile at room temperature, the lability being cancelled by 2-mercaptoethanol.  相似文献   

17.
The complex between active site-specific metal-depleted horse liver alcohol dehydrogenase and NADH has been studied with X-ray crystallographic methods to 2.9 A resolution. The electron density maps revealed that only the catalytic zinc ions are removed, whereas the non-catalytic zinc sites ae fully occupied. A gross conformational change in the protein induced by co-enzyme binding takes place in this enzyme species despite the absence of the metal ion in the catalytic center. This circumstance is of great importance in the understanding and further analysis of the trigger mechanisms operating during the conformation transition in alcohol dehydrogenase, since the catalytic center is located at the hinge region for a domain rotation in the subunit, and the metal atom is essential for catalysis. The overall protein structure is the same as that of an NADH complex of the native zinc enzyme and the co-enzyme is bound in a similar manner. The local structural changes observed are restricted to the empty metal binding site.  相似文献   

18.
The role of coenzyme binding in regulating interflavin electron transfer in human cytochrome P450 reductase (CPR) has been studied using temperature-jump spectroscopy. Previous studies [Gutierrez, A., Paine, M., Wolf, C.R., Scrutton, N.S., & Roberts, G.C.K. Biochemistry (2002) 41, 4626-4637] have shown that the observed rate, 1/tau, of interflavin electron transfer (FADsq - FMNsq-->FADox - FMNhq) in CPR reduced at the two-electron level with NADPH is 55 +/- 2 s-1, whereas with dithionite-reduced enzyme the observed rate is 11 +/- 0.5 s-1, suggesting that NADPH (or NADP+) binding has an important role in controlling the rate of internal electron transfer. In relaxation experiments performed with CPR reduced at the two-electron level with NADH, the observed rate of internal electron transfer (1/tau = 18 +/- 0.7 s-1) is intermediate in value between those seen with dithionite-reduced and NADPH-reduced enzyme, indicating that the presence of the 2'-phosphate is important for enhancing internal electron transfer. To investigate this further, temperature jump experiments were performed with dithionite-reduced enzyme in the presence of 2',5'-ADP and 2'-AMP. These two ligands increase the observed rate of interflavin electron transfer in two-electron reduced CPR from 1/tau = 11 s-1 to 35 +/- 0.2 s-1 and 32 +/- 0.6 s-1, respectively. Reduction of CPR at the two-electron level by NADPH, NADH or dithionite generates the same spectral species, consistent with an electron distribution that is equivalent regardless of reductant at the initiation of the temperature jump. Spectroelectrochemical experiments establish that the redox potentials of the flavins of CPR are unchanged on binding 2',5'-ADP, supporting the view that enhanced rates of interdomain electron transfer have their origin in a conformational change produced by binding NADPH or its fragments. Addition of 2',5'-ADP either to the isolated FAD-domain or to full-length CPR (in their oxidized and reduced forms) leads to perturbation of the optical spectra of both the flavins, consistent with a conformational change that alters the environment of these redox cofactors. The binding of 2',5'-ADP eliminates the unusual dependence of the observed flavin reduction rate on NADPH concentration (i.e. enhanced at low coenzyme concentration) observed in stopped-flow studies. The data are discussed in the context of previous kinetic studies and of the crystallographic structure of rat CPR.  相似文献   

19.
The respiratory complex I couples the electron transfer from NADH to ubiquinone with a translocation of protons across the membrane. Its nucleotide-binding site is made up of a unique Rossmann fold to accommodate the binding of the substrate NADH and of the primary electron acceptor flavin mononucleotide. Binding of NADH includes interactions of the hydroxyl groups of the adenosine ribose with a conserved glutamic acid residue. Structural analysis revealed that due to steric hindrance and electrostatic repulsion, this residue most likely prevents the binding of NADPH, which is a poor substrate of the complex. We produced several variants with mutations at this position exhibiting up to 200-fold enhanced catalytic efficiency with NADPH. The reaction of the variants with NAD(P)H is coupled with proton translocation in an inhibitor-sensitive manner. Thus, we have created an energy-converting NADPH:ubiquinone oxidoreductase, an activity so far not found in nature. Remarkably, the oxidation of NAD(P)H by the variants leads to an enhanced production of reactive oxygen species.  相似文献   

20.
Glutathione reductase bound to an affinity matrix of immobilized GSSG was eluted by its coenzyme NADPH rather than by its substrate GSSG or by NADH. NADP+ could also elute the enzyme, but a high concentration was needed to release enzyme activity in a sharp peak. This chromatographic system exhibits an unusual form of biospecificity in which the enzyme is bound to an immobilized substrate but released by its soluble cofactor.  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