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.
A NADH dehydrogenase was isolated from an inner membrane-enriched fraction of beetroot mitochondria (Beta vulgaris L.) by solubilization with sodium deoxycholate and purified using gel filtration and affinity chromatography. The NADH dehydrogenase preparation contained a minor ATPase contamination. Beetroot mitochondria were chosen as the isolation material for purifying the enzymes responsible for oxidizing matrix NADH due to the absence of the externally facing NADH dehydrogenase in the variety we have used. The purified NADH dehydrogenase complex catalyzed the reduction of various electron acceptors with NADH as the electron donor, was not sensitive to rotenone inhibition, and had a slow NADPH-ubiquinone 5 reductase activity. The isolated complex contained 14 major polypeptides. It was concluded that the dehydrogenase represented a form of the plant mitochondrial complex I and not the internally facing rotenone-insensitive NADH dehydrogenase found in plant mitochondria because of its complex structure, its cross-reactivity with antisera raised against bovine heart mitochondrial complex I, and the similarity of its kinetics and inhibitor responses to rotenone-sensitive NADH oxidation by beetroot submitochondrial particles.  相似文献   

2.
A sodium ion efflux, together with a proton influx and an inside-positive ΔΨ, was observed during NADH-respiration by Rhodothermus marinus membrane vesicles. Proton translocation was monitored by fluorescence spectroscopy and sodium ion transport by 23Na-NMR spectroscopy. Specific inhibitors of complex I (rotenone) and of the dioxygen reductase (KCN) inhibited the proton and the sodium ion transport, but the KCN effect was totally reverted by the addition of menaquinone analogues, indicating that both transports were catalyzed by complex I. We concluded that the coupling ion of the system is the proton and that neither the catalytic reaction nor the establishment of the delta-pH are dependent on sodium, but the presence of sodium increases proton transport. Moreover, studies of NADH oxidation at different sodium concentrations and of proton and sodium transport activities allowed us to propose a model for the mechanism of complex I in which the presence of two different energy coupling sites is suggested.  相似文献   

3.
Low concentrations of NAD+ inhibit the NADH: acceptor reductase reactions catalyzed by soluble NADH dehydrogenase from bovine heart mitochondria. The degree of incomplete inhibition of the enzyme depends on the nature and concentration of artificial electron acceptors and is manifested only at low concentrations of the latter. Marked inhibition was demonstrated for the 2.6-dichlorophenolindophenol-, ferricyanide- and O2-reductase reactions, being weakly pronounced during the measurement of the NADH: cytochrome c reductase activity. The inhibition of the above reactions by oxidized NAD+ isn't competitive towards NADH. A kinetic scheme is proposed, which postulates NADH: acceptor reductase reactions occurrence via two mechanisms, namely, a ping-pong mechanism and oxidation of the product-enzyme complex by the acceptor. It was shown that low concentrations of NAD+ also inhibit the NADH oxidase reaction catalyzed by complex I.  相似文献   

4.
Lauryl sulfate inhibits the Deltamu;(H)(+)-dependent reverse electron transfer reactions catalyzed by NADH:ubiquinone oxidoreductase (Complex I) in coupled bovine heart submitochondrial particles and in vesicles derived from Paracoccus denitrificans. The inhibitor affects neither NADH oxidase (coupled or uncoupled) nor NADH:ferricyanide reductase and succinate oxidase activities at the concentrations that selectively prevent the succinate-supported, rotenone-sensitive NAD(+) or ferricyanide reduction. Possible uncoupling effects of the inhibitor are ruled out: in contrast to oligomycin and gramicidin, which increases and decreases the rate of the reverse electron transfer, respectively, in parallel with their coupling and uncoupling effects, lauryl sulfate does not affect the respiratory control ratio. A mechanistic model for the unidirectional effect of lauryl sulfate on the Complex I catalyzed oxidoreduction is proposed.  相似文献   

5.
Conditions for the reversible dissociation of flavin mononucleotide (FMN) from the membrane-bound mitochondrial NADH:ubiquinone oxidoreductase (complex I) are described. The catalytic activities of the enzyme, i.e. rotenone-insensitive NADH:hexaammineruthenium III reductase and rotenone-sensitive NADH:quinone reductase decline when bovine heart submitochondrial particles are incubated with NADH in the presence of rotenone or cyanide at alkaline pH. FMN protects and fully restores the NADH-induced inactivation whereas riboflavin and flavin adenine dinucleotide do not. The data show that the reduction of complex I significantly weakens the binding of FMN to protein thus resulting in its dissociation when the concentration of holoenzyme is comparable with K(d ( approximately 10(-8)M at pH 10.0).  相似文献   

6.
The membrane fraction of Bacillus subtilis catalyzes the reduction of fumarate to succinate by NADH. The activity is inhibited by low concentrations of 2-(heptyl)-4-hydroxyquinoline-N-oxide (HOQNO), an inhibitor of succinate: quinone reductase. In sdh or aro mutant strains, which lack succinate dehydrogenase or menaquinone, respectively, the activity of fumarate reduction by NADH was missing. In resting cells fumarate reduction required glycerol or glucose as the electron donor, which presumably supply NADH for fumarate reduction. Thus in the bacteria, fumarate reduction by NADH is catalyzed by an electron transport chain consisting of NADH dehydrogenase (NADH:menaquinone reductase), menaquinone, and succinate dehydrogenase operating in the reverse direction (menaquinol:fumarate reductase). Poor anaerobic growth of B. subtilis was observed when fumarate was present. The fumarate reduction catalyzed by the bacteria in the presence of glycerol or glucose was not inhibited by the protonophore carbonyl cyanide m-chlorophenyl hydrazone (CCCP) or by membrane disruption, in contrast to succinate oxidation by O2. Fumarate reduction caused the uptake by the bacteria of the tetraphenyphosphonium cation (TPP+) which was released after fumarate had been consumed. TPP+ uptake was prevented by the presence of CCCP or HOQNO, but not by N,N'-dicyclohexylcarbodiimide, an inhibitor of ATP synthase. From the TPP+ uptake the electrochemical potential generated by fumarate reduction was calculated (Deltapsi = -132 mV) which was comparable to that generated by glucose oxidation with O2 (Deltapsi = -120 mV). The Deltapsi generated by fumarate reduction is suggested to stem from menaquinol:fumarate reductase functioning in a redox half-loop.  相似文献   

7.
The steady-state kinetics of the transhydrogenase reaction (the reduction of acetylpyridine adenine dinucleotide (APAD+) by NADH, DD transhydrogenase) catalyzed by bovine heart submitochondrial particles (SMP), purified Complex I, and by the soluble three-subunit NADH dehydrogenase (FP) were studied to assess a number of the Complex I-associated nucleotide-binding sites. Under the conditions where the proton-pumping transhydrogenase (EC 1.6.1.1) was not operating, the DD transhydrogenase activities of SMP and Complex I exhibited complex kinetic pattern: the double reciprocal plots of the velocities were not linear when the substrate concentrations were varied in a wide range. No binary complex (ping-pong) mechanism (as expected for a single substrate-binding site enzyme) was operating within any range of the variable substrates. ADP-ribose, a competitive inhibitor of NADH oxidase, was shown to compete more effectively with NADH (Ki = 40 microM) than with APAD+ (Ki = 150 microM) in the transhydrogenase reaction. FMN redox cycling-dependent, FP catalyzed DD transhydrogenase reaction was shown to proceed through a ternary complex mechanism. The results suggest that Complex I and the simplest catalytically competent fragment derived therefrom (FP) possess more than one nucleotide-binding sites operating in the transhydrogenase reaction.  相似文献   

8.
Additional characterization of complex I, rotenone-sensitive NADH:ubiquinone oxidoreductase, in the mitochondria of Trypanosoma brucei brucei has been obtained. Both proline:cytochrome c reductase and NADH:ubiquinone oxidoreductase of procyclic T. brucei were inhibited by the specific inhibitors of complex I rotenone, piericidin A, and capsaicin. These inhibitors had no effect on succinate: cytochrome c reductase activity. Antimycin A, a specific inhibitor of the cytochrome bc1 complex (ubiquinol:cytochrome c oxidoreductase), blocked almost completely cytochrome c reductase activity with either proline or succinate as electron donor, but had no inhibitory effect on NADH:ubiquinone oxidoreductase activity. The rotenone-sensitive NADH:ubiquinone oxidoreductase of procyclic T. brucei was partially purified by sucrose density centrifugation of mitochondria solubilized with dodecyl-beta-D-maltoside, with an approximately eightfold increase in specific activity compared to that of the mitochondrial membranes. Four polypeptides of the partially purified enzyme were identified as the homologous subunits of complex I (51 kDa, PSST, TYKY, and ND4) by immunoblotting with antibodies raised against subunits of Paracoccus denitrificans and against synthetic peptides predicted from putative complex I subunit genes encoded by mitochondrial and nuclear T. brucei DNA. Blue Native polyacrylamide gel electrophoresis of T. brucei mitochondrial membrane proteins followed by immunoblotting revealed the presence of a putative complex I with a molecular mass of 600 kDa, which contains a minimum of 11 polypeptides determined by second-dimensional Tricine-SDS/PAGE including the 51 kDa, PSST and TYKY subunits.  相似文献   

9.
Tightly coupled bovine heart submitochondrial particles treated to activate complex I and to block ubiquinol oxidation were capable of rapid uncoupler-sensitive inside-directed proton translocation when a limited amount of NADH was oxidized by the exogenous ubiquinone homologue Q1. External alkalization, internal acidification and NADH oxidation were followed by the rapidly responding (t1/2 < or = 1 s) spectrophotometric technique. Quantitation of the initial rates of NADH oxidation and external H+ decrease resulted in a stoichiometric ratio of 4 H+ vectorially translocated per 1 NADH oxidized at pH 8.0. ADP-ribose, a competitive inhibitor of the NADH binding site decreased the rates of proton translocation and NADH oxidation without affecting -->H+/2e- stoichiometry. Rotenone, piericidin and thermal deactivation of complex I completely prevented NADH-induced proton translocation in the NADH-endogenous ubiquinone reductase reaction. NADH-exogenous Q1 reductase activity was only partially prevented by rotenone. The residual rotenone- (or piericidin-) insensitive NADH-exogenous Q1 reductase activity was found to be coupled with vectorial uncoupler-sensitive proton translocation showing the same -->H+/2e- stoichiometry of 4. It is concluded that the transfer of two electrons from NADH to the Q1-reactive intermediate located before the rotenone-sensitive step is coupled with translocation of 4 H+.  相似文献   

10.
Enoyl-acyl carrier protein (ACP) reductases are critical for bacterial type II fatty acid biosynthesis and thus are attractive targets for developing novel antibiotics. We determined the crystal structure of enoyl-ACP reductase (FabK) from Streptococcus pneumoniae at 1.7 A resolution. There was one dimer per asymmetric unit. Each subunit formed a triose phosphate isomerase (TIM) barrel structure, and flavin mononucleotide (FMN) was bound as a cofactor in the active site. The overall structure was similar to the enoyl-ACP reductase (ER) of fungal fatty acid synthase and to 2-nitropropane dioxygenase (2-ND) from Pseudomonas aeruginosa, although there were some differences among these structures. We determined the crystal structure of FabK in complex with a phenylimidazole derivative inhibitor to envision the binding site interactions. The crystal structure reveals that the inhibitor binds to a hydrophobic pocket in the active site of FabK, and this is accompanied by induced-fit movements of two loop regions. The thiazole ring and part of the ureido moiety of the inhibitor are involved in a face-to-face pi-pi stacking interaction with the isoalloxazine ring of FMN. The side-chain conformation of the proposed catalytic residue, His144, changes upon complex formation. Lineweaver-Burk plots indicate that the inhibitor binds competitively with respect to NADH, and uncompetitively with respect to crotonoyl coenzyme A. We propose that the primary basis of the inhibitory activity is competition with NADH for binding to FabK, which is the first step of the two-step ping-pong catalytic mechanism.  相似文献   

11.
Verticipyrone has recently been isolated from the culture broth of Verticillium sp. and shown to inhibit NADH fumarate reductase, as well as NADH oxidoreductase (complex I) of the mitochondrial electron transport chain. In order to assess the structural elements in verticipyrone essential for complex I inhibitor, 15 structural analogues were prepared and analyzed for their effects on mitochondrial NADH oxidoreductase and NADH oxidase activities. Also measured were the abilities of several of the analogues to inhibit respiration as judged by a shift to glycolysis, and to inhibit the growth of several mammalian cell lines. The nature of the pyrone ring was shown to be important to potency of inhibition, as was the length and nature of substituents in the side chain of the analogues.  相似文献   

12.
The method described for the isolation of mitochondrial complex I (NADH-ubiquinone reductase) from bovine hearts could not be applied to donkey hearts as unacceptably large losses in enzyme activity occurred. This method was modified for the isolation of complex I using donkey hearts and two complexes were obtained: complex IA which was physiologically inactive and complex IB which was physiologically active as it catalyzed the reaction from NADH to ubiquinone. Both complexes had relatively low enzyme activity with artificial electron acceptors, except with potassium ferricyanide, and had more or less the same amount of acid-labile sulfur and nonheme iron although the polypeptide composition differed to a great extent.  相似文献   

13.
In mitochondria of Neurospora crassa grown in the presence of chloramphenicol a small form of NADH:ubiquinone reductase is made in place of the normal electron-transfer-complex I. This smaller enzyme has a molecular mass of approximately 350 kDa and consists of (at least) 13 different subunits which are all synthesized in the cytoplasm. The complex I which is normally found in Neurospora has a molecular mass of approximately 700 kDa and consists of around 30 different subunits, of which at least six are made in the mitochondria. Immunoblotting and peptide mapping suggest that the subunits of the small enzyme are homologous to subunits of the large enzyme, one subunit might even be identical. The small and the large NADH:ubiquinone reductases have the same high-affinity binding site for NADH but the two enzymes differ in the affinity and inhibitor sensitivity of the ubiquinone-binding site. The possibility is discussed that the small NADH:ubiquinone reductase is primitive isoform of complex I.  相似文献   

14.
Vanadate-dependent NAD(P)H oxidation, catalyzed by rat liver microsomes and microsomal NADPH-cytochrome P450 reductase (P450 reductase) and NADH-cytochrome b5 reductase (b5 reductase), was investigated. These enzymes and intact microsomes catalyzed NAD(P)H oxidation in the presence of either ortho- or polyvanadate. Antibody to P450 reductase inhibited orthovanadate-dependent NADPH oxidation catalyzed by either purified P450 reductase or rat liver microsomes and had no effect on the rates of NADH oxidation catalyzed by b5 reductase. NADPH-cytochrome P450 reductase catalyzed orthovanadate-dependent NADPH oxidation five times faster than NADH-cytochrome b5 reductase catalyzed NADH oxidation. Orthovanadate-dependent oxidation of either NADPH or NADH, catalyzed by purified reductases or rat liver microsomes, occurred in an anaerobic system, which indicated that superoxide is not an obligate intermediate in this process. Superoxide dismutase (SOD) inhibited orthovanadate, but not polyvanadate-mediated, enzyme-dependent NAD(P)H oxidation. SOD also inhibited when pyridine nucleotide oxidation was conducted anaerobically, suggesting that SOD inhibits vanadate-dependent NAD(P)H oxidation by a mechanism independent of scavenging of O2-.  相似文献   

15.
The reaction catalyzed by 2-aminobenzoyl-coenzyme-A monooxygenase/reductase from a denitrifying Pseudomonas sp. has been investigated. 2-Aminobenzoyl-CoA and 2-amino[carboxy-14C]benzoyl-CoA were synthesized enzymatically using 2-aminobenzoyl-CoA synthetase from the same organism. The product was purified by chromatography and characterized by ultraviolet/visible and 1H-NMR spectroscopy. The conversion of 2-aminobenzoyl-CoA catalyzed by the monooxygenase/reductase requires NADH and oxygen, and yields at least two different products depending on the relative concentration of NADH. At [NADH] less than Km (40 microM), i.e. [NADH]/[2-aminobenzoyl-CoA] approximately 0.02-0.05, the main product is probably a hydroxylated derivative of 2-aminobenzoyl-CoA, which is characterized by an absorbance maximum around 375 nm. When [NADH]/[2-aminobenzoyl-CoA] approximately 2-5, the predominant product is a non-aromatic coenzyme A thioester (lambda max approximately 320 nm). The stoichiometry in this case is 2.1-2.4 mol NADH oxidized (mol oxygen consumed)-1 (mol 2-aminobenzoyl-CoA metabolized)-1. The product is extremely unstable in the acidic pH range and undergoes decarboxylation in a few minutes at pH less than 5. Some degree of stabilisation is obtained upon reduction with sodium borohydride, probably resulting in a further reduced non-aromatic coenzyme-A thioester.  相似文献   

16.
NADH-coenzyme Q reductase from bovine heart mitochondria (complex I) was incorporated into phospholipid vesicles by the cholate dialysis procedure. Mixtures of purified phosphatidylcholine and phosphatidylethanolamine were required. Oxidation of NADH by coenzyme Q1 catalyzed by the reconstituted vesicles was coupled to proton translocation, directed inward, with an H+/2e ratio greater than 1.4. Similar experiments measuring proton translocation in submitochondrial particles gave an H+/2e ratio of 1.8. The proton translocation in both systems was not seen in the presence of uncoupling agents and was in addition to the net proton uptake from the reduction of coenzyme Q1 by NADH. Electron transfer in the reconstituted vesicles also caused the uptake of the permeant anion tetraphenylboron. The rate of electron transfer by the reconstituted vesicles was stimulated about 3-fold by uncouplers or by valinomycin plus nigericin and K+ ions. The results indicate that energy coupling can be observed with isolated NADH-coenzyme Q reductase if the enzyme complex is properly incorporated into a phospholipid vesicle.  相似文献   

17.
The NADH:ubiquinone, but not the NADH:ferricyanide, reductase activity of mitochondrial complex I (NADH:ubiquinone oxidoreductase) is inhibited by incubation of the enzyme at pH 6.0 and 0 degree C with ethoxyformic anhydride (EFA), and the inhibition is partially reversed by subsequent incubation of EFA-treated complex I with hydroxylamine. These results and spectral changes of EFA-treated complex I in the u.v. region are consistent with modification of essential histidyl or tyrosyl residues between the primary NADH dehydrogenase and the site of ubiquinone reduction. Treatment of complex I with EFA in the presence of high concentrations of Seconal or Demerol did not protect against EFA inactivation, suggesting that the site of EFA modification may not be the same as the inhibiton sites of Seconal and Demerol. However, the presence of NADH during incubation of complex I with EFA greatly enhanced the inhibition rate, indicating that the reduced conformation of complex I is more susceptible to attack by EFA.  相似文献   

18.
The catalytic properties of the rotenone-sensitive NADH:ubiquinone reductase (Complex I) in bovine heart submitochondrial particles and in inside-out vesicles derived from Paracoccus denitrificans and Rhodobacter capsulatus were compared. The prokaryotic enzymes catalyze the NADH oxidase and NADH:quinone reductase reactions with similar kinetic parameters as those for the mammalian Complex I, except for lower apparent affinities for the substrates--nucleotides. Unidirectional competitive inhibition of NADH oxidation by ADP-ribose, previously discovered for submitochondrial particles, was also evident for tightly coupled P. denitrificans vesicles, thus suggesting that a second, NAD(+)-specific site is present in the simpler prokaryotic enzyme. The inhibitor sensitivity of the forward and reverse electron transfer reactions was compared. In P. denitrificans and Bos taurus vesicles different sensitivities to rotenone and Triton X-100 for the forward and reverse electron transfer reactions were found. In bovine heart preparations, both reactions showed the same sensitivity to piericidin, and the inhibition was titrated as a straight line. In P. denitrificans, the forward and reverse reactions show different sensitivity to piericidin and the titrations of both activities were curvilinear with apparent I(50) (expressed as mole of inhibitor per mole of enzyme) independent of the enzyme concentration. This behavior is explained by a model involving two different sites rapidly interacting with piericidin within the hydrophobic phase.  相似文献   

19.
Scheide D  Huber R  Friedrich T 《FEBS letters》2002,512(1-3):80-84
The proton-pumping NADH:ubiquinone oxidoreductase, also called complex I, is the first energy-transducing complex of many respiratory chains. Homologues of complex I are present in the three domains of life. Here, we report the properties of complex I in membranes of the hyperthermophilic bacterium Aquifex aeolicus. The complex reacted with NADH but not with NADPH and F(420)H(2) as electron donors. Short-chain analogues of ubiquinone like decyl-ubiquinone and ubiquinone-2 were suitable electron acceptors. The affinities towards NADH and ubiquinone-2 were comparable to the ones obtained with the Escherichia coli complex I. The reaction was inhibited by piericidin A at the same concentration as in E. coli. The complex showed an unusual pH optimum at pH 9 and a maximal rate at 80 degrees C. We found no evidence for the presence of an alternative, single subunit NADH dehydrogenase in A. aeolicus membranes. The NADH:ferricyanide reductase activity of detergent extracts of A. aeolicus membranes sedimented as a protein with a molecular mass of approximately 550 kDa. From the data we concluded that A. aeolicus contains a NADH:ubiquinone oxidoreductase resembling complex I of mesophilic bacteria.  相似文献   

20.
Respiratory chains of bacteria and mitochondria contain closely related forms of the proton-pumping NADH:ubiquinone oxidoreductase, or complex I. The bacterial complex I consists of 14 subunits, whereas the mitochondrial complex contains some 25 extra subunits in addition to the homologues of the bacterial subunits. One of these extra subunits with a molecular mass of 40 kDa belongs to a heterogeneous family of reductases/isomerases with a conserved nucleotide binding site. We deleted this subunit in Neurospora crassa by gene disruption. In the mutant nuo 40, a complex I lacking the 40 kDa subunit is assembled. The mutant complex I does not contain tightly bound NADPH present in wild-type complex I. This NADPH cofactor is not connected to the respiratory electron pathway of complex I. The mutant complex has normal NADH dehydrogenase activity and contains the redox groups known for wild-type complex I, one flavin mononucleotide and four iron-sulfur clusters detectable by electron paramagnetic resonance spectroscopy. In the mutant complex these groups are all readily reduced by NADH. However, the mutant complex is not capable of reducing ubiquinone. A recently described redox group identified in wild-type complex I by UV-visible spectroscopy is not detectable in the mutant complex. We propose that the reductase/isomerase subunit with its NADPH cofactor takes part in the biosynthesis of this new redox group.  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