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 17 毫秒
1.
Progress in understanding the role of NAD(P)H oxidation in plant respiration is restricted by the lack of access to specific inhibitors of each of the unknown number of NAD(P)H dehydrogenases in the inner mitochondrial membrane. Platanetin (3,5,7,8-tetrahydroxy-6-isoprenyl flavone) is known to be an inhibitor of extermal NADH oxidation by plant mitochondria, while 7-iodo-acridone-4-carboxylic acid (IACA) is an inhibitor of an internal, rotenone-insensitive NAD(P)H dehydrogenase isolated from yeast mitochondria.
Here we show that platanetin inhibits external NAD(P)H oxidation by intact potato ( Solanum tuberosum L. cv. Bintje) tuber mitochondria, deamino-NADH oxidation by Complex I assayed using inside-out submitochondrial particles from these mitochondria, and rotenone-insensitive NAD(P)H oxidation by these submitochondrial particles. IACA was found to inhibit the oxidation of external NADH and succinate by intact mitochondria with similar efficiency. However, IACA also inhibited NADPH and duroquinol oxidation by intact mitochondria as well as deamino-NADH and NAD(P)H oxidation by inside-out submitochondrial particles. This indicates that IACA has several sites of inhibition in the electron transport chain. The lack of specificity of both platanetin and IACA prevents these inhibitors from being used to shed more light on the identity of the NAD(P)H dehydrogenases in plant mitochondria.  相似文献   

2.
Several NAD(P)+-dependent hydroxysteroid dehydrogenases, namely 3α-hydroxysteroid dehydrogenase, β-hydroxysteroid dehydrogenase, 7α-hydroxysteroid dehydrogenase, and 12α-hydroxysteroid dehydrogenase were separately immobilized on nylon tubes for the continuous-flow automated assay of hydroxysteroids. 3α-Hydroxysteroid dehydrogenase was also immobilized on pore glass. Spectrophotometric monitoring in the visible region, where blank values were markedly reduced, was achieved through the Meldola blue catalyzed transfer of hydrogen from NAD(P)H to a tetrazolium salt. Nylon-tube-immobilized enzymes maintained 45–55% of the original activity after 1 month of intermittent use. The operational range, using the “end point” approach, was 1–25 nmol of steroid and the assay speed 10–15 samples/h. Reliable results were obtained in the determination of 3α-hydroxysteroids and 3β,17β-hydroxysteroids in urine and total bile acids in serum.  相似文献   

3.
Synthesis and release of NAD(P)ase by Neurospora crassa wild type was studied in experiments in which mycelia grown in Vogel minimal medium were transferred to media containing protein as the only carbon source. Several results are presented suggesting that the NAD(P)ase may be induced by the presence of protein in the culture medium. Low concentrations of sucrose or glucose (0.1%), Casamino acids or some amino acids such as methionine, cysteine, phenylalanine and tryptophan strongly repressed the enzyme synthesis. Under induction conditions NAD(P)ase and alkaline protease appeared together in the culture medium. It would appear that NAD(P)ase and alkaline protease are coordinately regulated by a common control mechanism related to carbon catabolism.  相似文献   

4.
The mechanisms of nitric oxide (NO) synthesis in plants have been extensively investigated. NO degradation can be just as important as its synthesis in controlling steady-state levels of NO. Here, we examined NO degradation in mitochondria isolated from potato tubers and the contribution of the respiratory chain to this process. NO degradation was faster in mitochondria energized with NAD(P)H than with succinate or malate. Oxygen consumption and the inner membrane potential were transiently inhibited by NO in NAD(P)H-energized mitochondria, in contrast to the persistent inhibition seen with succinate. NO degradation was abolished by anoxia and superoxide dismutase, which suggested that NO was consumed by its reaction with superoxide anion (O2). Antimycin-A stimulated and myxothiazol prevented NO consumption in succinate- and malate-energized mitochondria. Although favored by antimycin-A, NAD(P)H-mediated NO consumption was not abolished by myxothiazol, indicating that an additional site of O2 generation, besides complex III, stimulated NO degradation. Larger amounts of O2 were generated in NAD(P)H- compared to succinate- or malate-energized mitochondria. NAD(P)H-mediated NO degradation and O2 production were stimulated by free Ca2+ concentration. Together, these results indicate that Ca2+-dependent external NAD(P)H dehydrogenases, in addition to complex III, contribute to O2 production that favors NO degradation in potato tuber mitochondria.  相似文献   

5.
不对称还原胺化反应是制备医药中间体手性胺结构单元的重要反应。目前已有许多不同种类的酶被应用于合成手性胺,其中NAD(P)H依赖型氧化还原酶催化的还原胺化反应最为引人注目,因为其能够一步将潜手性酮化合物完全转化为光学纯的手性胺化合物。文中以亚胺还原酶、氨基酸脱氢酶、冠瘿碱脱氢酶和还原性酮胺化酶为例,从NAD(P)H依赖型氧化还原酶的结构特征、作用机理、分子改造及催化应用等方面,综述了其在不对称还原胺化合成手性胺领域的研究进展。  相似文献   

6.
Hybridoma cells were grown at steady state under both reductiveand oxidative stress and the intracellular fluxes weredetermined by mass-balancing techniques. By decreasing the dissolved oxygen pressure (pO2) in the bioreactor, the reduced formof nicotinamide adenine nucleotide (NADH) was enhanced relativeto the oxidized form (NAD+). Oxidative stress, as a resultof which the NAP(P)+/NAD(P)H-ratio increases, was generatedby both the enhancement of the pO2 to 100% air saturationand by the addition of the artificial electron acceptorphenazine methosulphate (PMS) to the culture medium. It wasfound that fluxes of dehydrogenase reactions by which NAD(P)H isproduced decreased under hypoxic conditions. For example, thedegradation rates of arginine, isoleucine, lysine and theglutamate dehydrogenase flux were significantly lower at oxygenlimitation, and increased at higher pO2 levels and when PMSwas added to the culture medium. In contrast, the prolinesynthesis reaction, which requires NADPH, decreased under PMSstress. The flux of the NADH-requiring lactate dehydrogenase reaction also strongly decreased from 19 to 3,4 pmol/cell/day,under oxygen limitation and under PMS stress, respectively. Thedata show that metabolic-flux balancing can be used to determinehow mammalian respond to oxidative and reduction stress.  相似文献   

7.
The asymmetric bioreduction of alkenes bearing an electron-withdrawing group using flavin-dependent enzymes from the 'old yellow enzyme' family at the expense of NAD(P)H yields the corresponding non-racemic alkanes going in hand with the creation of up to two chiral carbon centres. To avoid external cofactor recycling, this intriguing biotransformation was hitherto performed using whole microbial cells, which frequently showed insufficient stereoselectivities and/or undesired side reactions because of the action of competing enzymatic activities. Co-expression of enoate reductases with the corresponding redox enzymes for NAD(P)H recycling in a suitable host enables to overcome these drawbacks to furnish highly stereoselective and 'clean' C=C bioreductions on a preparative scale that are difficult to perform by conventional means.  相似文献   

8.
Ribitol dehydrogenase from Zymomonas mobilis (ZmRDH) catalyzes the conversion of ribitol to d-ribulose and concomitantly reduces NAD(P)(+) to NAD(P)H. A systematic approach involving an initial sequence alignment-based residue screening, followed by a homology model-based screening and site-directed mutagenesis of the screened residues, was used to study the molecular determinants of the cofactor specificity of ZmRDH. A homologous conserved amino acid, Ser156, in the substrate-binding pocket of the wild-type ZmRDH was identified as an important residue affecting the cofactor specificity of ZmRDH. Further insights into the function of the Ser156 residue were obtained by substituting it with other hydrophobic nonpolar or polar amino acids. Substituting Ser156 with the negatively charged amino acids (Asp and Glu) altered the cofactor specificity of ZmRDH toward NAD(+) (S156D, [k(cat)/K(m)(,NAD)]/[k(cat)/K(m)(,NADP)] = 10.9, where K(m)(,NAD) is the K(m) for NAD(+) and K(m)(,NADP) is the K(m) for NADP(+)). In contrast, the mutants containing positively charged amino acids (His, Lys, or Arg) at position 156 showed a higher efficiency with NADP(+) as the cofactor (S156H, [k(cat)/K(m)(,NAD)]/[k(cat)/K(m)(,NADP)] = 0.11). These data, in addition to those of molecular dynamics and isothermal titration calorimetry studies, suggest that the cofactor specificity of ZmRDH can be modulated by manipulating the amino acid residue at position 156.  相似文献   

9.
The respiratory chain of plant mitochondria differs from that in mammalian mitochondria by containing several rotenone-insensitive NAD(P)H dehydrogenases. Two of these are located on the outer, cytosolic surface of the inner membrane. One is specific for NADH, the other for NADPH. Only the latter is inhibited by diphenyleneiodonium (DPI). Both of these enzymes are normally dependent upon Ca2+ for activity and this constitutes a potentially important mechanism by which the cell can regulate the oxidation of cytosolic NAD(P)H via the concentration of free Ca2+. This and other potential regulatory mechanisms such as the substrate concentration and polyamines are discussed.  相似文献   

10.
Phenylacetaldehyde dehydrogenase (PAD) and lactaldehyde dehydrogenase (ALD) share some structural and kinetic properties. One difference is that PAD can use NAD+ and NADP+, whereas ALD only uses NAD+. An acidic residue has been involved in the exclusion of NADP+ from the active site in pyridine nucleotide-dependent dehydrogenases. However, other factors may participate in NADP+ exclusion. In the present work, analysis of the sequence of the region involved in coenzyme binding showed that residue F180 of ALD might participate in coenzyme specificity. Interestingly, F180T mutation rendered an enzyme (ALD-F180T) with the ability to use NADP+. This enzyme showed an activity of 0.87 micromol/(min * mg) and K(m) for NADP+ of 78 microM. Furthermore, ALD-F180T exhibited a 16-fold increase in the V(m) /K(m) ratio with NAD+ as the coenzyme, from 12.8 to 211. This increase in catalytic efficiency was due to a diminution in K(m) for NAD+ from 47 to 7 microM and a higher V(m) from 0.51 to 1.48 micromol/(min * mg). In addition, an increased K(d) for NADH from 175 (wild-type) to 460 microM (mutant) indicates a faster product release and possibly a change in the rate-limiting step. For wild-type ALD it is described that the rate-limiting step is shared between deacylation and coenzyme dissociation. In contrast, in the present report the rate-limiting step in ALD-F180T was determined to be exclusively deacylation. In conclusion, residue F180 participates in the exclusion of NADP+ from the coenzyme binding site and disturbs the binding of NAD+.  相似文献   

11.
Exogenous NAD utilization or pyridine nucleotide cycle metabolism is used by many bacteria to maintain NAD turnover and to limit energy-dependent de novo NAD synthesis. The genus Haemophilus includes several important pathogenic bacterial species that require NAD as an essential growth factor. The molecular mechanisms of NAD uptake and processing are understood only in part for Haemophilus. In this report, we present data showing that the outer membrane lipoprotein e(P4), encoded by the hel gene, and an exported 5'-nucleotidase (HI0206), assigned as nadN, are necessary for NAD and NADP utilization. Lipoprotein e(P4) is characterized as an acid phosphatase that uses NADP as substrate. Its phosphatase activity is inhibited by compounds such as adenosine or NMN. The nadN gene product was characterized as an NAD-nucleotidase, responsible for the hydrolysis of NAD. H. influenzae hel and nadN mutants had defined growth deficiencies. For growth, the uptake and processing of the essential cofactors NADP and NAD required e(P4) and 5'-nucleotidase. In addition, adenosine was identified as a potent growth inhibitor of wild-type H. influenzae strains, when NADP was used as the sole source of nicotinamide-ribosyl.  相似文献   

12.
The conformation of NAD bound to diphtheria toxin (DT), an ADP-ribosylating enzyme, has been compared to the conformations of NAD(P) bound to 23 distinct NAD(P)-binding oxidoreductase enzymes, whose structures are available in the Brookhaven Protein Data Bank. For the oxidoreductase enzymes, NAD(P) functions as a cofactor in electron transfer, whereas for DT, NAD is a labile substrate in which the N-glycosidic bond between the nicotinamide ring and the N-ribose is cleaved. All NAD(P) conformations were compared by (1) visual inspection of superimposed molecules, (2) RMSD of atomic positions, (3) principal component analysis, and (4) analysis of torsion angles and other conformational parameters. Whereas the majority of oxidoreductase-bound NAD(P) conformations are found to be similar, the conformation of NAD bound to DT is found to be unusual. Distinctive features of the conformation of NAD bound to DT that may be relevant to DT''s function as an ADP-ribosylating enzyme include (1) an unusually short distance between the PN and N1N atoms, reflecting a highly folded conformation for the nicotinamide mononucleotide (NMN) portion of NAD, and (2) a torsion angle chi N approximately 0 degree about the scissile N-glycosidic bond, placing the nicotinamide ring outside of the preferred anti and syn orientations. In NAD bound to DT, the highly folded NMN conformation and torsion angle chi N approximately 0 degree could contribute to catalysis, possibly by orienting the C1''N atom of NAD for nucleophilic attack, or by placing strain on the N-glycosidic bond, which is cleaved by DT. The unusual overall conformation of NAD bound to DT is likely to reflect the structure of DT, which is unusual among NAD(P)-binding enzymes. In DT, the NAD binding site is formed at the junction of two antiparallel beta-sheets. In contrast, although the 24 oxidoreductase enzymes belong to at least six different structural classes, almost all of them bind NAD(P) at the C-terminal end of a parallel beta-sheet. The structural alignments and principal component analysis show that enzymes of the same structural class bind to particularly similar conformations of NAD(P), with few exceptions. The conformation of NAD bound to DT superimposes closely with that of an NAD analogue bound to Pseudomonas exotoxin A, an ADP-ribosylating toxin that is structurally homologous to DT. This suggests that all of the ADP-ribosylating enzymes that are structurally homologous to DT and ETA will bind a highly similar conformation of NAD.  相似文献   

13.
Mitochondria recently have emerged as important sites in controlling NO levels within the cell. In this study, the synthesis of nitric oxide (NO) from nitrite and its degradation by mitochondria isolated from Arabidopsis thaliana were examined. Oxygen and NO concentrations in the reaction medium were measured with specific electrodes. Nitrite inhibited the respiration of isolated A. thaliana mitochondria, in competition with oxygen, an effect that was abolished or potentiated when electron flow occurred via alternative oxidase (AOX) or cytochrome c oxidase (COX), respectively. The production of NO from nitrite was detected electrochemically only under anaerobiosis because of a superoxide-dependent process of NO degradation. Electron leakage from external NAD(P)H dehydrogenases contributed the most to NO degradation as higher rates of Amplex Red-detected H2O2 production and NO consumption were observed in NAD(P)H-energized mitochondria. Conversely, the NO-insensitive AOX diminished electron leakage from the respiratory chain, allowing the increase of NO half-life without interrupting oxygen consumption. These results show that the accumulation of nitric oxide derived from nitrite reduction and the superoxide-dependent mechanism of NO degradation in isolated A. thaliana mitochondria are influenced by the external NAD(P)H dehydrogenases and AOX, revealing a role for these alternative proteins of the mitochondrial respiratory chain in the control of NO levels in plant cells.  相似文献   

14.
There are multiple routes of NAD(P)H oxidation associated with the inner membrane of plant mitochondria. These are the phosphorylating NADH dehydrogenase, otherwise known as Complex I, and at least four other nonphosphorylating NAD(P)H dehydrogenases. Complex I has been isolated from beetroot, broad bean, and potato mitochondria. It has at least 32 polypeptides associated with it, contains FMN as its prosthetic group, and the purified enzyme is sensitive to inhibition by rotenone. In terms of subunit complexity it appears similar to the mammalian and fungal enzymes. Some polypeptides display antigenic similarity to subunits fromNeurospora crassa but little cross-reactivity to antisera raised against some beef heart complex I subunits. Plant complex I contains eight mitochondrial encoded subunits with the remainder being nuclear-encoded. Two of these mitochondrial-encoded subunits, nad7 and nad9, show homology to corresponding nuclear-encoded subunits inNeurospora crassa (49 and 30 kDa, respectively) and beef heart CI (49 and 31 kDa, respectively), suggesting a marked difference between the assembly of CI from plants and the fungal and mammalian enzymes. As well as complex I, plant mitochondria contain several type-II NAD(P)H dehydrogenases which mediate rotenone-insensitive oxidation of cytosolic and matrix NADH. We have isolated three of these dehydrogenases from beetroot mitochondria which are similar to enzymes isolated from potato mitochondria. Two of these enzymes are single polypeptides (32 and 55 kDa) and appear similar to those found in maize mitochondria, which have been localized to the outside of the inner membrane. The third enzyme appears to be a dimer comprised of two identical 43-kDa subunits. It is this enzyme that we believe contributes to rotenone-insensitive oxidation of matrix NADH. In addition to this type-II dehydrogenases, several observations suggest the presence of a smaller form of CI present in plant mitochondria which is insensitive to rotenone inhibition. We propose that this represents the peripheral arm of CI in plant mitochondria and may participate in nonphosphorylating matrix NADH oxidation.  相似文献   

15.
HMG-CoA reductase (HMGR) is an enzyme critical for cellular cholesterol synthesis in mammals and isoprenoid synthesis in certain eubacteria, catalyzing the NAD(P)H-dependent reduction of HMG-CoA to mevalonate. We have isolated the gene encoding HMG-CoA reductase from Listeria monocytogenes and expressed the recombinant 6x-His-tagged form in Escherichia coli. Using NAD(P)(H), the enzyme catalyzes HMG-CoA reduction approximately 200-fold more efficiently than mevalonate oxidation in vitro. The purified enzyme exhibits dual coenzyme specificity, utilizing both NAD(H) and NADP(H) in catalysis; however, catalytic efficiency using NADP(H) is approximately 200 times greater than when using NAD(H). The statins mevinolin and mevastatin are weak inhibitors of L. monocytogenes HMG-CoA reductase, requiring micromolar concentrations for inhibition. Three-dimensional modeling reveals that the overall structure of L. monocytogenes HMG-CoA reductase is likely similar to the known structure of the class II enzyme from Pseudomonas mevalonii. It appears that the enzyme has catalytic amino acids in analogous positions that likely play similar roles and also has a flap domain that brings a catalytic histidine into the active site. However, in L. monocytogenes HMG-CoA reductase histidine 143 and methionine 186 are present in the putative NAD(P)(H)-selective site, possibly interacting with the 2' phosphate of NADP(H) or 2' hydroxyl of NAD(H) and providing the active site architecture necessary for dual coenzyme specificity.  相似文献   

16.
17.
18.
Araucaria angustifolia (Bert.) O. Kuntze is a species critically endangered of extinction and its development and propagation is strongly affected by abiotic stress. We have previously shown the activation of uncoupling protein in A. angustifolia embryogenic stem cells subjected to cold stress. Now, we have furthered those studies by exposing these cells to cold stress (4?±?1?°C for either 24 or 48?h) and evaluating parameters associated with oxidative stress and alterations in the cellular and mitochondrial responses. Cold stress affect the H2O2 levels and lipid peroxidation increased after both stress condition, an effect associated with the decrease in the activities of peroxidases, catalase and ascorbate/dehydroascorbate ratio. On the other hand, the activities of ascorbate peroxidase, monodehydroascorbate and dehydroascorbate reductases increased as an indication of adaptation. Another important impact of cold stress conditions was the decrease of external alternative NAD(P)H dehydrogenases activity and the increase of mitochondrial mass. These results show that cold stress induces oxidative stress in A. angustifolia embryogenic cells, which results in activation of the glutathione-ascorbate cycle as a compensation for the decrease in the activities of catalase, peroxidases, and external NAD(P)H dehydrogenases. Our results contribute to the understanding of the pathways that gymnosperms employ to overcome oxidative stress, which must be explored in order to improve the methods of conservation and propagation of A. angustifolia.  相似文献   

19.
The heterolactic bacterium Oenococcus oeni ferments fructose by a mixed heterolactic/mannitol fermentation. For heterolactic fermentation of fructose, the phosphoketolase pathway is used. The excess NAD(P)H from the phosphoketolase pathway is reoxidized by fructose (yielding mannitol). It is shown here that, under conditions of C-limitation or decreased growth rates, fructose can be fermented by heterolactic fermentation yielding nearly stoichiometric amounts of lactate, ethanol and CO(2). Quantitative evaluation of NAD(P)H-producing (phosphoketolase pathway) and -reoxidizing (ethanol, mannitol and erythritol pathways) reactions demonstrated that at high growth rates or in batch cultures the ethanol pathway does not have sufficient capacity for NAD(P)H reoxidation, requiring additional use of the mannitol pathway to maintain the growth rate. In addition, insufficient capacities to reoxidize NAD(P)H causes inhibition of growth, whereas increased NAD(P)H reoxidation by electron acceptors such as pyruvate increases the growth rate.  相似文献   

20.
Metabolic activity in eukaryotic cells is known to naturally oscillate. We have recently observed a 20-s period NAD(P)H oscillation in neutrophils and other polarized cells. Here we show that when polarized human neutrophils are exposed to interferon-gamma or to ultra-low-frequency electric fields with periods double that of the NAD(P)H oscillation, the amplitude of the NAD(P)H oscillations increases. Furthermore, increases in NAD(P)H amplitude, whether mediated by interferon-gamma or by an oscillating electric field, signals increased production of reactive oxygen metabolites. Hence, amplitude modulation of NAD(P)H oscillations suggests a novel signaling mechanism in polarized cells.  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