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.
Acetylpyridine NADP replaced NADP in promoting the Mn2+ ion-requiring mitochondrial "malic" enzyme of Hymenolepis diminuta. Disrupted mitochondria displayed low levels of an apparent oxaloacetate-forming malate dehydrogenase activity when NAD or acetylpyridine NAD served as the coenzyme. Significant malate-dependent reduction of acetylpyridine NAD by H. diminuta mitochondria required Mn2+ ion and NADP, thereby indicating the tandem operation of "malic" enzyme and NADPH:NAD transhydrogenase. Incubation of mitochondrial preparations with oxaloacetate resulted in a non-enzymatic decarboxylation reaction. Coupling of malate oxidation with electron transport via the "malic" enzyme and transhydrogenase was demonstrated by polarographic assessment of mitochondrial reduced pyridine nucleotide oxidase activity.  相似文献   

2.
NAD kinase catalyzes the phosphorylation of NAD+ to synthesize NADP+, whereas NADH kinase catalyzes conversion of NADH to NADPH. The mitochondrial protein Pos5 of Saccharomyces cerevisiae shows much higher NADH kinase than NAD kinase activity and is therefore referred to as NADH kinase. To clarify the structural determinant underlying the high NADH kinase activity of Pos5 and its selectivity for NADH over NAD+, we determined the tertiary structure of Pos5 complexed with NADH at a resolution of 2.0 Å. Detailed analysis, including a comparison of the tertiary structure of Pos5 with the structures of human and bacterial NAD kinases, revealed that Arg-293 of Pos5, corresponding to His-351 of human NAD kinase, confers a positive charge on the surface of NADH-binding site, whereas the corresponding His residue does not. Accordingly, conversion of the Arg-293 into a His residue reduced the ratio of NADH kinase activity to NAD kinase activity from 8.6 to 2.1. Conversely, simultaneous changes of Ala-330 and His-351 of human NAD kinase into Ser and Arg residues significantly increased the ratio of NADH kinase activity to NAD kinase activity from 0.043 to 1.39; human Ala-330 corresponds to Pos5 Ser-272, which interacts with the side chain of Arg-293. Arg-293 and Ser-272 were highly conserved in Pos5 homologs (putative NADH kinases), but not in putative NAD kinases. Thus, Arg-293 of Pos5 is a major determinant of NADH selectivity. Moreover, Ser-272 appears to assist Arg-293 in achieving the appropriate conformation.  相似文献   

3.
Aging, a major risk factor in Alzheimer's disease (AD), is associated with an oxidative redox shift, decreased redox buffer protection, and increased free radical reactive oxygen species (ROS) generation, probably linked to mitochondrial dysfunction. While NADH is the ultimate electron donor for many redox reactions, including oxidative phosphorylation, glutathione (GSH) is the major ROS detoxifying redox buffer in the cell. Here, we explored the relative importance of NADH and GSH to neurodegeneration in aging and AD neurons from nontransgenic and 3xTg‐AD mice by inhibiting their synthesis to determine whether NADH can compensate for the GSH loss to maintain redox balance. Neurons stressed by either depleting NAD(P)H or GSH indicated that NADH redox control is upstream of GSH levels. Further, although depletion of NAD(P)H or GSH correlated linearly with neuron death, compared with GSH depletion, higher neurodegeneration was observed when NAD(P)H was extrapolated to zero, especially in old age, and in the 3xTg‐AD neurons. We also observed an age‐dependent loss of gene expression of key redox‐dependent biosynthetic enzymes, NAMPT (nicotinamide phosphoribosyltransferase), and NNT (nicotinamide nucleotide transhydrogenase). Moreover, age‐related correlations between brain NNT or NAMPT gene expression and NADPH levels suggest that these genes contribute to the age‐related declines in NAD(P)H. Our data indicate that in aging and more so in AD‐like neurons, NAD(P)H redox control is upstream of GSH and an oxidative redox shift that promotes neurodegeneration. Thus, NAD(P)H generation may be a more efficacious therapeutic target upstream of GSH and ROS.  相似文献   

4.
The 2.1 A resolution crystal structure of flavin reductase P with the inhibitor nicotinamide adenine dinucleotide (NAD) bound in the active site has been determined. NAD adopts a novel, folded conformation in which the nicotinamide and adenine rings stack in parallel with an inter-ring distance of 3.6 A. The pyrophosphate binds next to the flavin cofactor isoalloxazine, while the stacked nicotinamide/adenine moiety faces away from the flavin. The observed NAD conformation is quite different from the extended conformations observed in other enzyme/NAD(P) structures; however, it resembles the conformation proposed for NAD in solution. The flavin reductase P/NAD structure provides new information about the conformational diversity of NAD, which is important for understanding catalysis. This structure offers the first crystallographic evidence of a folded NAD with ring stacking, and it is the first enzyme structure containing an FMN cofactor interacting with NAD(P). Analysis of the structure suggests a possible dynamic mechanism underlying NADPH substrate specificity and product release that involves unfolding and folding of NADP(H).  相似文献   

5.
Ian M. Mller  John M. Palmer 《BBA》1981,638(2):225-233
(1) The optimum pH for the oxidation of exogenous NADH by mitochondria from both Jerusalem artichoke (Helianthus tuberosus) tubers and Arum maculatum spadices was 7.0–7.1. NADPH oxidation had a lower optimum pH of 6.6 in Arum and 6.0 in Jerusalem artichoke mitochondria. In both types of mitochondria the rates of NADH and NADPH oxidation were identical below pH 6.0–5.5. (2) It is shown conclusively that neither a phosphatase converting NADPH to NADH nor a nicotinamide nucleotide transhydrogenase was involved in the oxidation of NADPH by these mitochondria. (3) Palmitoyl-CoA, an inhibitor of transhydrogenase activity in mammalian mitochondria, inhibits both NADH and NADPH oxidation by plant mitochondria with a Ki of about 10 μM. (4) It is concluded that the known properties of NAD(P)H oxidation are best explained by assuming the presence of a second dehydrogenase specific for NADPH. At low pH, electron flow from the two dehydrogenases to oxygen shares a common rate-limiting step.  相似文献   

6.
Transhydrogenase couples the redox reaction between NAD(H) and NADP(H) to proton translocation across a membrane. In membrane vesicles from Escherichia coli and Rhodospirillum rubrum, the transhydrogenase reaction (measured in the direction driving inward proton translocation) was inhibited by Zn(2+) and Cd(2+). However, depending on pH, the metal ions either had no effect on, or stimulated, "cyclic" transhydrogenation. They must, therefore, interfere specifically with steps involving binding/release of NADP(+)/NADPH: the steps thought to be associated with proton translocation. It is suggested that Zn(2+) and Cd(2+) bind in the proton-transfer pathway and block inter-conversion of states responsible for changing NADP(+)/NADPH binding energy.  相似文献   

7.
Mitochondria from the muscle of Ascaris lumbricoides var. suis function anaerobically. NADH is generated in the intermembrane space as a consequence of the "malic" enzyme reaction. It has been suggested that this reducing equivalent in the form of hydride ion, would be translocated across the inner membrane in order to mediate ATP generation via the fumarate reductase reaction. In accord with this suggestion, intact Ascaris mitochondria showed appreciable NADH oxidase activity. Sonication resulted in an approximately 2-fold increase in NADH oxidase activity, whereas "malic" enzyme, fumarase, and NADH:NAD+ transhydrogenase activities increased approximately 7- to 14-fold, respectively. Phosphorylation capabilities and permeability toward pyridine nucleotides also indicated the intactness of the mitochondria. Ascaris mitochondria incubated anaerobically in the presence of fumarate, and [14C]NADH catalyzed a rapid reduction of the fumarate to succinate with the concomitant formation of equivalent quantities of extramitochondrial NAD+. However, very little isotope was recovered from the washed mitochondria, indicating the possibility of hydride ion translocation in the absence of nucleotide translocation. NADH:NAD+ transhydrogenase has been isolated from the muscle mitochondria of the intestinal nematode, Ascaris lumbricoides var. suis. The enzyme seems to have been solubilized from the mitochondrial membrane fraction by treatment with sodium deoxycholate followed by dialysis and subsequent adsorption by and elution from alumina C gamma. No NADPH:NAD+ transhydrogenase activity was detectable, making the Ascaris system unique over others reported. Activity was protected by L-cysteine, reduced glutathione and dithioerythritol, but strongly inhibited by low concentrations of p-chloromercuribenzoate or silver nitrate. The thionicotinamide derivative of NAD+ (thioNAD+) was employed to accept hydride ions from NADH in order to assay spectrophotometrically at 398 nm. Apparent Km values for thioNAD+ and NADH were 1 X 10(-4) M and 8 X 10(-6) M, respectively. That the physiological nucleotide, could act as hydride ion acceptor from NADH was indicated by the findings that NAD+ competitively inhibited the reduction of thioNAD+ when assayed at 398 nm. The additional finding of a noncompetitive inhibition between NAD+ and NADH suggested at least two binding sites on the enzyme, one for NADH and another common site for NAD+ and thioNAD+. More conclusive evidence indicating the participation of NAD+ as acceptor was obtained by incubation of the enzyme with NADH and [14C]NAD+ and demonstrating a rapid formation of [14C]NADH. These findings, in conjunction with those discussed above, suggest a physiological function of this enzyme in hydride ion translocation.  相似文献   

8.
ATP and respiration (NADH)-driven NAD(P)+ transhydrogenase (EC 1.6.1.1) activities are low in membranes from Escherichia coli cultured on yeast extract medium (17 and 21 nmol/min × mg) but high on glucose (82 and 142 nmol/min × mg). The ATPase and respiratory activities in both cases appeared comparable. Growth of the bacteria in yeast extract medium followed by washing and replacement into a glucose medium showed that after 3 h the energy-linked and energy-independent NAD(P)+ transhydrogenase (reduction of acetylpyridine NAD+ by NADPH) activities had appeared simultaneously. Incorporation of chloramphenicol or omission of glucose in the induction medium resulted in no increase in these activities indicating that de novo protein synthesis is required for the induction of energy-linked and -independent NAD(P)+ transhydrogenase. It was found that the Km values for acetylpyridine NAD+ and NADPH for the energy-independent reaction in membranes from glucose grown cells (143 and 62 μm) were similar to those in membranes from cells grown on glucose-yeast extract (135 and 45 μm), respectively, but the maximum velocity at infinite acetyl pyridine NAD+ and NADPH increased from 353 to 2175 nmol/min × mg. Furthermore, the membrane-bound NAD(P)+ transhydrogenase in glucose-yeast extract grown cells showed substrate inhibition at high NADPH and low acetyl pyridine NAD+ levels. Further kinetic data demonstrate that the mechanism of the energy-independent NAD(P)+ transhydrogenase in E. coli is similar to that of the mitochondrial enzyme and exhibits similar responses to competitive inhibitors at the NAD+ and NADPH sites.  相似文献   

9.
The control of isocitrate oxidation by rat liver mitochondria   总被引:3,自引:1,他引:2  
1. The factors capable of affecting the rate of isocitrate oxidation in intact mitochondria include the rate of isocitrate penetration, the activity of the NAD-specific and NADP-specific isocitrate dehydrogenases, the activity of the transhydrogenase acting from NADPH to NAD(+), the rate of NADPH oxidation by the reductive synthesis of glutamate and the activity of the respiratory chain. A quantitative assessment of these factors was made in intact mitochondria. 2. The kinetic properties of the NAD-specific and NADP-specific isocitrate dehydrogenases extracted from rat liver mitochondria were examined. 3. The rate of isocitrate oxidation through the respiratory chain in mitochondria with coupled phosphorylation is approximately equal to the maximal of the NAD-specific isocitrate dehydrogenase but at least ten times as great as the transhydrogenase activity from NADPH to NAD(+). 4. It is concluded that the energy-dependent inhibition of isocitrate oxidation by palmitoylcarnitine oxidation is due to an inhibition of the NAD-specific isocitrate dehydrogenase. 5. Kinetic studies of NAD-specific isocitrate dehydrogenase demonstrated that its activity could be inhibited by one or more of the following: an increased reduction of mitochondrial NAD, an increased phosphorylation of mitochondrial adenine nucleotides or a fall in the mitochondrial isocitrate concentration. 6. Uncoupling agents stimulate isocitrate oxidation by an extent equal to the associated stimulation of transhydrogenation from NADPH to NAD(+). 7. A technique is described for continuously measuring with a carbon dioxide electrode the synthesis of glutamate from isocitrate and ammonia.  相似文献   

10.
Membranes isolated from vegetative cells of Anabaena variabilis (ATCC 29413) oxidine NADPH and NADH in a cyanide-sensitive reaction, NADPH being the better donor. In addition, both act as reductants for photosystem I in the light and in the presence of DCMU. An NADH-regenerating system has been introduced experimentally, functioning as donor system for photosystem I of Anabaena membranes. This NADH dehydrogenation is strictly light-dependent, sensitive to DBMIB, establishes a transhydrogenase system with NADP+ was hydrogen acceptor and competes with oxygen. Kinetic analysis of oxygen uptake with reference to nucleotide concentration exhibited different Km- and Vmax-values for NADPH and NADH in the light and in the dark.  相似文献   

11.
Rosalind Wilson 《BBA》2006,1757(3):215-223
All transhydrogenases appear to have three components: dI, which binds NAD(H), and dIII, which binds NADP(H), protrude from the membrane, and dII spans the membrane. However, the polypeptide composition of the enzymes varies amongst species. The transhydrogenases of Mycobacterium tuberculosis and of Rhodospirillum rubrum have three polypeptides. Sequence analysis indicates that an ancestral three-polypeptide enzyme evolved into transhydrogenases with either two polypeptides (such as the Escherichia coli enzyme) or one polypeptide (such as the mitochondrial enzyme). The fusion steps in each case probably led to the development of an additional transmembrane helix. A hybrid transhydrogenase was constructed from the dI component of the M. tuberculosis enzyme and the dII and dIII components of the R. rubrum enzyme. The hybrid catalyses cyclic transhydrogenation but not the proton-translocating, reverse reaction. This shows that nucleotide-binding/release at the NAD(H) site, and hydride transfer, are fully functional but that events associated with NADP(H) binding/release are compromised. It is concluded that sequence mismatch in the hybrid prevents a conformational change between dI and dIII which is essential for the step accompanying proton translocation.  相似文献   

12.
Summary The single-channel current recording technique has been used to study the influences that the pyridine nucleotides NAD, NADH, NADP and NADPH have on the gating of ATP-sensitive K+ channels in an insulin-secreting cell line (RINm5F). The effects of the nucleotides were studied at the intracellular surface using either excised inside-out membrane patches or permeabilized cells. All four pyridine nucleotides were found to evoke similar effects. At low concentrations, 100 m and less, each promoted channel opening whereas high concentrations, 500 m and above, evoked channel closure. The degree of K+ channel activation by pyridine nucleotides (low conc.) was found to be similar to that evoked by the same concentrations of ADP or GTP, whereas the degree of K+ channel inhibition (high conc.) was less marked than that evoked by the same concentrations of ATP, and never resulted in refreshment of K+ channels following removal. The effects of NAD, NADH, NADP and NADPH seemed to interact with those of ATP and ADP. In the presence of 1mm ADP and 4mm ATP, 10 to 100 m concentrations of the pyridine nucleotides could not evoke channel opening, whereas concentrations of 500 m and above were found to evoke channel closure. In the presence of 2mm ATP and 0.5mm ADP, however, 10 to 100 m concentrations of the pyridine nucleotides were able to activate K+ channels.  相似文献   

13.
S R Earle  S G O'Neal  R R Fisher 《Biochemistry》1978,17(22):4683-4690
Chemical-modification studies on submitochondrial particle pyridine dinucleotide transhydrogenase (EC 1.6.1.1) demonstrate the presence of one class of sulfhydryl group in the nicotinamide adenine dinucleotide phosphate (NADP) site and another peripheral to the active site. Reaction of the peripheral sulfhydryl group with N-ethylmaleimide, or both classes with 5,5'-dithiobis(2-nitrobenzoic acid), completely inactivated transhydrogenase. NADP+ or NADPH nearly completely protected against 5,5'-dithiobis(2-nitrobenzoic acid) inactivation and modification of both classes of sulfhydryl groups, while NADP+ only partially protected against and NADPH substantially stimulated N-ethylmaleimide inactivation. Methyl methanethiolsulfonate treatment resulted in methanethiolation at both classes of sulfhydryl groups, and either NADP+ or NADPH protected only the NADP site group. S-Methanethio and S-cyano transhydrogenases were active derivatives with pH optima shifted about 1 unit lower than that of the native enzyme. These experiments indicate that neither class of sulfhydryl group is essential for transhydrogenation. Lack of involvement of either sulfhydryl group in energy coupling to transhydrogenation is suggested by the observations that S-methanethio transhydrogenase is functional in (a) energy-linked transhydrogenation promoted by phenazine methosulfate mediated ascorbate oxidation and (b) the generation of a membrane potential during the reduction of NAD+ by reduced nicotinamide adenine dinucleotide phosphate (NADPH).  相似文献   

14.
Insertion mutant Ins2 of the cyanobacterium Synechocystis sp. PCC 6803, lacking NAD(P)H:quinone oxidoreductase (NQR) encoded by drgA gene, was characterized by higher sensitivity to quinone-type inhibitors (menadione and plumbagin) than wild type (WT) cells. In photoautotrophically grown cyanobacterial cells more than 60% of NADPH:quinone-reductase activity, as well as all NADPH:dinoseb-reductase activity, was associated with the function of NQR. NQR activity was observed only in soluble fraction of cyanobacterial cells, but not in membrane fraction. The effects of menadione and menadiol on the reduction of Photosystem I reaction center (P700(+)) after its photooxidation in the presence of DCMU were studied using the EPR spectroscopy. The addition of menadione increased the rate of P700(+) reduction in WT cells, whereas in Ins2 mutant the reduction of P700(+) was strongly inhibited. In the presence of menadiol the reduction of P700(+) was accelerated both in WT and Ins2 mutant cells. These data suggest that NQR protects the cyanobacterial cells from the toxic effect of exogenous quinones by their reduction to hydroquinones. These data may also indicate the probable functional homology of Synechocystis sp. PCC 6803 NQR with mammalian and plant NAD(P)H:quinone oxidoreductases (DT-diaphorases).  相似文献   

15.
Kinetic, circular dichroism, and NADH and NADPH fluorescence quenching studies indicate that these compounds interact with the antimalarial drug primaquine (PQ). The affinity of both pyridine nucleotides for PQ is similar. The data are in contrast with a previous report (Thornalley et al. (1983) Biochem. Pharmacol. 32, 3571-3575) suggesting specificity for the interaction with NADPH. The complex was seen to facilitate electron transfer from NAD(P)H to oxygen, generating oxygen-free radicals which were detected by the spin-trapping technique and to flavin nucleotides, giving rise to flavin semiquinone radicals which were demonstrated by direct ESR spectroscopy under anaerobic conditions. A twofold increase in oxygen uptake and hydroxyl radical generation by the NAD(P)H-PQ complex was observed in the presence of hemoglobin. This effect was independent of heme concentration (in the range 1 X 10(-5)-1 X 10(-4) M) and oxidation state of the iron. Under anaerobic conditions, the NAD(P)H-PQ complex reduces Fe-III to Fe-II hemoglobin, and under aerobic conditions about 65% of the heme chromophore is irreversibly destroyed. Superoxide dismutase inhibits hydroxyl radical generation by the NAD(P)H-PQ pair; this effect is not observed in the presence of hemoglobin. In the presence of microsomes there is a 10-fold increase in both oxygen consumption and hydroxyl radical generation by the NAD(P)H-PQ pair. The fact that both pyridine nucleotides are active, and the inability of SKF 525A in decreasing hydroxyl radical generation, suggests that microsomal reductases are involved in the catalysis.  相似文献   

16.
Using isolated chloroplasts or purified thylakoids from photoautotrophically grown cells of the chromophytic alga Pleurochloris meiringensis (Xanthophyceae) we were able to demonstrate a membrane bound NAD(P)H dehydrogenase activity. NAD(P)H oxidation was detectable with menadione, coenzyme Q0, decylplastoquinone and decylubiquinone as acceptors in an in vitro assay. K m-values for both pyridine nucleotides were in the molar range (K m[NADH]=9.8 M, K m[NADPH]=3.2 M calculated according to Lineweaver-Burk). NADH oxidation was optimal at pH 9 while pH dependence of NADPH oxidation showed a main peak at 9.8 and a smaller optimum at pH 7.5–8. NADH oxidation could be completely inhibited with rotenone, an inhibitor of mitochondrial complex I dehydrogenase, while NADPH oxidation revealed the typical inhibition pattern upon addition of oxidized pyridine nucleotides reported for ferredoxin: NADP+ reductase. Partly-denaturing gel electrophoresis followed by NAD(P)H dehydrogenase activity staining showed that NADPH and NADH oxidizing proteins had different electrophoretic mobilities. As revealed by denaturing electrophoresis, the NADH oxidizing enzyme had one main subunit of 22 kDa and two further polypeptides of 29 and 44 kDa, whereas separation of the NADPH depending protein yielded five bands of different molecular weight. Measurement of oxygen consumption due to PS I mediated methylviologen reduction upon complete inhibition of PS II showed that the NAD(P)H dehydrogenase is able to catalyze an input of electrons from NADH to the photosynthetic electron transport chain in case of an oxidized plastoquinone-pool. We suggest ferredoxin: NADP+ reductase to be the main NADPH oxidizing activity while a thylakoidal NAD(P)H: plastoquinone oxidoreductase involved in the chlororespiratory pathway in the dark acts mainly as an NADH oxidizing enzyme.Abbreviations Coenzyme Q0-2,3-dimethoxy-5-methyl-1,4-benzoquinone - FNR ferredoxin: NADP+ reductase - MD menadione - MV methylviologen - NDH NAD(P)H dehydrogenase - PQ plastoquinone - PQ10 decylplastoquinone - SDH succinate dehydrogenase - UQ10 decylubiquinone (2,3-dimethoxy-5-methyl-6-decyl-1,4-benzoquinone)  相似文献   

17.
Midgut and fatbody mitochondria from fifth larval instar Manduca sexta display a membrane-associated transhydrogenase that catalyzes a reversible hydride ion transfer between NADP(H) and NAD(H). The NADPH-forming activity occurs as a nonenergy- or energy-linked activity with energy for the latter derived from either electron transport-dependent NADH or succinate utilization, or ATP hydrolysis by Mg++-dependent ATPase. During the ten-day developmental period preceding the larval-pupal molt (fifth larval instar), significant peaks in the mitochondrial transhydrogenase activities of midgut and fatbody tissues were noted and these peaks were coincident with the onset of wandering behavior and with the fifty-fold increase in ecdysone 20-monooxygenase (E20-M) activity previously reported for M. sexta midgut. Since E20-M preferentially uses NADPH in catalyzing ecdysone conversion to the physiologically active molting hormone, 20-hydroxyecdysone, the physiological and developmental significance of the mitochondrial, NADPH-forming energy-linked transhydrogenations were made apparent. Moreover, that the increases in all transhydrogenase activities resulted from de novo enzyme synthesis were indicated by the cycloheximide-dependent reductions in these activities.  相似文献   

18.
Arabidopsis mitochondria contain two NAD(+)-malic enzymes, NAD-ME1 and NAD-ME2. These proteins have similar affinity for their substrates but display opposite regulation by fumarate, which strongly stimulates NAD-ME1 but inhibits NAD-ME2 activity. Here, the interaction of NAD-ME1 and -2 with fumarate was investigated by kinetic approaches, urea denaturation assays and intrinsic fluorescence quenching, in the absence and presence of NAD(+). Fumarate inhibited NAD-ME2 at saturating, but not at low, levels of NAD(+), and it behaved as competitive inhibitor with respect to L-malate. In contrast, NAD-ME1 fumarate activation was higher at suboptimal NAD(+) concentrations. In the absence of cofactor, the fluorescence of both NAD-ME1 and -2 is quenched by fumarate. However, for NAD-ME2 the quenching arises from a collisional phenomenon, while in NAD-ME1 the fluorescence decay can be explained by a static process that involves fumarate binding to the protein. Furthermore, the residue Arg84 of NAD-ME1 is essential for fumarate binding, as the mutant protein R84A exhibits a collisional quenching by this metabolite. Together, the results indicate that the differential fumarate regulation of Arabidopsis NAD-MEs, which is further modulated by NAD(+) availability, is related to the gaining of an allosteric site for fumarate in NAD-ME1 and an active site-associated inhibition by this C(4)-organic acid in NAD-ME2.  相似文献   

19.
Pyridine nucleotide transhydrogenases of bacterial cytosolic membranes and mitochondrial inner membranes are proton pumps in which hydride transfer between NADP(+) and NAD(+) is coupled to proton translocation across cytosolic or mitochondrial membranes. The pyridine nucleotide transhydrogenase of Escherichia coli is composed of two subunits (alpha and beta). Three domains are recognized. The extrinsic cytosolic domain 1 of the amino-terminal region of the alpha subunit bears the NAD(H)-binding site. The NADP(H)-binding site is present in domain 3, the extrinsic cytosolic carboxyl-terminal region of the beta subunit. Domain 2 is composed of the membrane-intrinsic carboxyl-terminal region of the alpha subunit and the membrane-intrinsic amino-terminal region of the beta subunit. Treatment of the transhydrogenase of E. coli with 4-chloro-7-nitrobenzo-2-oxa-1,3-diazole (NBD chloride) inhibited enzyme activity. Analysis of inhibition revealed that several sites on the enzyme were involved. NBD chloride modified two (betaCys-147 and betaCys-260) of the seven cysteine residues present in the transhydrogenase. Modification of betaCys-260 in domain 2 resulted in inhibition of enzyme activity. Modification of residues other than cysteine residues also resulted in inhibition of transhydrogenation as shown by use of a cysteine-free mutant enzyme. The beta subunit was modified by NBD chloride to a greater extent than the alpha subunit. Reaction of domain 2 and domain 3 was prevented by NADPH. Modification of domain 3 is probably not associated with inhibition of enzyme activity. Modification of domain 2 of the beta subunit resulted in a decreased binding affinity for NADPH at its binding site in domain 3. The product resulting from the reaction of NBD chloride with NADPH was a very effective inhibitor of transhydrogenation. In experiments with NBD chloride in the presence of NADPH it is likely that all of the sites of reaction described above will contribute to the inhibition observed. The NBD-NADPH adduct will likely be more useful than NBD chloride in investigations of the pyridine nucleotide transhydrogenase.  相似文献   

20.
NAD(H) kinase catalyzes the phosphorylation of NAD(H) to form NADP(H) using ATP or inorganic polyphosphate as a phosphoryl donor. While the enzyme is conserved throughout prokaryotes and eukaryotes, remarkable differences in kinetic parameters including substrate preference, cation dependence, and physiological roles exist among the organisms. In the present study, we biochemically characterized NAD(H) kinase from the anaerobic/microaerophilic fermentative protozoan parasite Entamoeba histolytica, which lacks the conventional mitochondria capable of oxidative phosphorylation, leading to ATP. The kinetic properties of E. histolytica NAD(H) kinase recombinantly produced in Escherichia coli showed remarkable differences from those in bacteria and higher eukaryotes. Entamoeba NAD(H) kinase preferred NADH to NAD+ as the phosphoryl acceptor, utilized nucleoside triphosphates including ATP, GTP and deoxyATP, but not nucleoside di-, mono-phosphates, or inorganic polyphosphates, as the phosphoryl donor. To further understand the physiological roles in E. histolytica, we generated a stable transformant overexpressing NAD(H) kinase. Overexpression of NAD(H) kinase resulted in a 1.6–2 fold increase in the NADPH and NADP+ concentrations, a 40% reduction of the intracellular concentration of reactive oxygen species, and also led to increased tolerance toward hydrogen peroxide. These data, together with the essentially of NAD(H) kinase gene, underscore its significance as an NADP(H)-producing enzyme in this organism, and should help in designing of drugs targeting this enzyme.  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