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.
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.
1. Incubation of NADH-ubiquinone oxidoreductase (Complex I) with chymotrypsin caused loss of rotenone-sensitive ubiquinone-1 reduction and an increase in rotenone-insensitive ubiquinone reduction. 2. Within the same time-course, NADH-K(3)Fe(CN)(6) oxidoreductase activity was unaffected. 3. Mixing of chymotrypsin-treated Complex I with Complex III did not give rise to NADH-cytochrome c oxidoreductase activity. 4. Gel electrophoresis in the presence of sodium dodecyl sulphate revealed selective degradation of several constituent polypeptides by chymotrypsin. 5. With higher chymotrypsin concentrations and longer incubation times, a decrease in NADH-K(3)Fe(CN)(6) oxidoreductase was observed. The kinetics of this decrease correlated with solubilization of the low-molecular-weight type-II NADH dehydrogenase (subunit mol.wts. 53000 and 27000) and with degradation of a polypeptide of mol.wt. 30000. 6. Phospholipid-depleted Complex I was more rapidly degraded by chymotrypsin. Specifically, a subunit of mol.wt. 75000, resistant to chymotrypsin in untreated Complex I, was degraded in phospholipid-depleted Complex I. In addition, the 30000-mol.wt. polypeptide was also more rapidly digested, correlating with an increased rate of transformation to type II NADH dehydrogenase.  相似文献   

3.
Mitochondria isolated from potato (Solanum tuberosum L.) tuber were investigated for the presence of a nicotinamide nucleotide transhydrogenase activity. Submitochondrial particles derived from these mitochondria by sonication catalyzed a reduction of NAD+ or 3-acetylpyridine-NAD+ by NADPH, which showed a maximum of about 50 to 150 nanomoles/minute·milligram protein at pH 5 to 6. The Km values for 3-acetylpyridine-NAD+ and NADPH were about 24 and 55 micromolar, respectively. Intact mitochondria showed a negligible activity in the absence of detergents. However, in the presence of detergents the specific activity approached about 30% of that seen with submitochondrial particles. The potato mitochondria transhydrogenase activity was sensitive to trypsin and phenylarsine oxide, both agents that are known to inhibit the mammalian transhydrogenase. Antibodies raised against rat liver transhydrogenase crossreacted with two peptides in potato tuber mitochondrial membranes with a molecular mass of 100 to 115 kilodaltons. The observed transhydrogenase activities may be due to an unspecific activity of dehydrogenases and/or to a genuine transhydrogenase. The activity contributions by NADH dehydrogenases and transhydrogenase to the total transhydrogenase activity were investigated by determining their relative sensitivities to trypsin. It is concluded that, at high or neutral pH, the observed transhydrogenase activity in potato tuber submitochondrial particles is due to the presence of a genuine and specific high molecular weight transhydrogenase. At low pH an unspecific reaction of an NADH dehydrogenase with NADPH contributes to the total trans-hydrogenase activity.  相似文献   

4.
The interaction between pure transhydrogenase and ATPase (Complex V) from beef heart mitochondria was investigated with transhydrogenase-ATPase vesicles in which the two proteins were co-reconstituted by dialysis or dilution procedures. In addition to phosphatidylcholine and phosphatidylethanolamine, reconstitution required phosphatidylserine and lysophosphatidylcholine. Transhydrogenase-ATPase vesicles catalyzed a 20-30-fold stimulation of the reduction of NADP+ or thio-NADP+ by NADH and a 70-fold shift of the apparent equilibrium expressed as the nicotinamide nucleotide ratio [NADPH][NAD+]/[NADP+][NADH]. In both of these respects, the transhydrogenase-ATPase vesicles were severalfold more efficient than beef heart submitochondrial particles. By measuring the ATP-driven transhydrogenase and the oligomycin-sensitive ATPase activities simultaneously and under the same conditions at low ATP concentrations, i.e. below 15 microM, the ATP-driven transhydrogenase/oligomycin-sensitive ATPase activity ratio was found to be about 3. This value is consistent with the stoichiometries of three protons translocated per ATP hydrolyzed and one proton translocated per NADPH formed and with a mechanism where the two enzymes interact through a delocalized proton-motive force.  相似文献   

5.
DNA is damaged in vivo by the Fenton reaction mediated by Fe2+ and cellular reductants such as NADH, which reduce Fe3+ to Fe2+ and allow the recycling of iron. To study the response of Escherichia coli to such cycling, the activities of several enzymes involved in nicotinamide nucleotide metabolism were measured following an H2O2 challenge. NADPH-dependent peroxidase, NADH/NADP+ transhydrogenase, and glucose-6-phosphate dehydrogenase were most strongly induced, increasing 2.5-3-fold. In addition, the cellular ratios of NADPH to NADH increased 6- or 92-fold 15 min after exposure to 0.5 or 5 mm H2O2, respectively. In vitro, NADH was oxidized by Fe3+ up to 16-fold faster than NADPH, despite their identical reduction potentials. To understand this rate difference, the interactions of Fe3+ and Ga3+ with NAD(P)H were examined by 1H, 13C, and 31P NMR spectroscopy. Association with NADH occurred primarily with adenine at N7 and the amino group, but for NADPH, strong metal interactions also occurred at the 2'-phosphate group. Interaction of M3+ (Fe3+ or Ga3+) with the adenine ring would bring it into close proximity to the redox-active nicotinamide ring in the folded form of NAD(P)H, but interaction of M3+ with the 2'-phosphate group would avoid this close contact. In addition, as determined by absorbance spectroscopy, the energy of the charge-transfer species was significantly higher for the Fe3+.NADPH complex than for the Fe3+.NADH complex. We therefore suggest that upon exposure to H2O2 the NADH pool is depleted, and NADPH, which is less reactive with Fe3+, functions as the major nicotinamide nucleotide reductant.  相似文献   

6.
Pyridine nucleotide transhydrogenase activities of a highly purified soluble NADH dehydrogenase and particulate NADH-ubiquinone reductase (Complex I) differ in their pH optima (5.0 and 6.0, respectively) and in their sensitivity to inhibition by Mg2+ and ATP. The oxidation of NADPH with ferricyanide as acceptor is very similar in both preparations with a pH optimum around 5.0. It is concluded that Complex I possesses two types of transhydrogenase activity, whereas only one has been found in the soluble dehydrogenase.  相似文献   

7.
The effects of hypophysectomy and subsequent administration of bovine growth hormone (0.1 IU/100 g body wt) and l-thyroxine (5 μg/100 g body wt) on respiration, energization-dependent fluorescence of 1-anilino-8-naphthalene sulfonate, NADH dehydrogenase, energy-independent nicotinamide nucleotide transhydrogenase, and succinate dehydrogenase activities were investigated in submitochondrial particles of rat liver. Hormones were injected daily for 7 days. Hypophysectomy decreased the respiratory rate with NADH or succinate and the activities of the three enzymes. Administration of growth hormone increased the respiration but showed selectivity toward NADH. Thyroxine increased the respiration more than growth hormone did with both substrates. Growth hormone increased the activities of NADH dehydrogenase and transhydrogenase whereas thyroxine increased the activity of only succinate dehydrogenase. After growth hormone treatment transhydrogenase activity was increased to about three times that of controls which may have significance in some processes mediated either directly or permissively by growth hormone. When both hormones were injected together, there was a significant decrease in the thyroxine-dependent rise in respiration on succinate as well as the growth hormone-dependent rise in enzyme activities. Fluorescence yield of 1-anilino-8-naphthalene sulfonate in unenergized submitochondrial particles remained unchanged independent of the hormonal status. Energization with succinate or NADH increased the fluorescence yield by about 2–20 times. Several parameters of energizationdependent fluorescence were decreased after hypophysectomy. In restoring these parameters, growth hormone and thyroxine showed specificity toward the energization substrate NADH and succinate, respectively. From the present results we conclude that (a) growth hormone and thyroxine regulate mitochondrial activity by affecting different segments of the respiratory chain, namely Complex I and Complex II, respectively, and (b) growth hormone and thyroxine exert moderating effects on one another.  相似文献   

8.
1. Oxidation of NADPH by various acceptors catalyzed by submitochondrial particles and a partially purified NADH dehydrogenase from beef heart was investigated. Submitochondrial particles devoid of nicotinamide nucleotide transhydrogenase activity catalyze an oxidation of NADPH by oxygen. The partially purified NADH dehydrogenase prepared from these particles catalyzes an oxidation of NADPH by acetylpyridine-NAD. In both cases the rates of oxidation are about two orders of magnitude lower than those obtained with NADH as electron donor. 2. The kinetic characteristics of the NADPH oxidase reaction and reduction of acetylpyridine-NAD by NADPH are similar with regard to pH dependences and affinities for NADPH, indicating that both reactions involve the same binding site for NADPH. The binding of NADPH to this site appears to be rate limiting for the overall reactions. 3. At redox equilibrium NADPH and NADH reduce FMN and iron-sulphur center 1 of NADH dehydrogenase to the same extents. The rate of reduction of FMN by NADPH is at least two orders of magnitude lower than with NADH. 4. It is concluded that NADPH is a substrate of NADH dehydrogenase and that the nicotinamide nucleotide is oxidized by submitochondrial particles via the NADH--binding site of the enzyme.  相似文献   

9.
NADH treatment of complex I at pH 7–8 results in the appearance of electron paramagnetic resonance (epr) signals at x band due to reduced ironsulfur centers 1, 2, 3 and 4, while NADPH treatment gives rise to the appearance of signals due to centers 2 and 3. Similar results are obtained with complex I preparations in which transhydrogenase activity from NADPH to NAD has been >95% inhibited by treatment of the complex with trypsin. At pH 6.5 and in the presence of rotenone, addition of NADPH to complex I or transhydrogenase-inhibited complex I results in partial reduction of iron-sulfur center 1 as well. These and other experiments with reduced 3-acetylpyridine adenine dinucleotide and NADPH + NAD as substrates have suggested that the differences in the reduction of complex I iron-sulfur centers by the above nucleotides are essentially quantitative and related to (a) the dehydrogenation rate of the nucleotides, and (b) autoxidation of complex I components under the epr experimental conditions.  相似文献   

10.
Ancylostoma ceylanicum, the hookworm parasite of cat, dog and man, was found to contain NADH and/or NADPH oxidase as well as fumarate reductase activities. Both the enzyme systems were predominantly located in the membranes of mitochondrial-rich preparations. The membranes also exhibited the presence of a reduced pyridine nucleotide transhydrogenase activity which transferred hydrogen from NADPH to NAD. Amongst respiratory inhibitors, rotenone (Site I inhibitor) markedly depressed both NADH oxidase and fumarate reductase while others, namely antimycin-A, KCN and azide, had a lesser effect.  相似文献   

11.
NAD (P) H-dependent reduction of nicotinamide N-oxide was investigated with rabbit liver preparations. Microsomes, microsomal NADPH-cytochrome c reductase or cytosolic aldehyde oxidase alone exhibited no nicotinamide N-oxide reductase activity in the presence of NADPH or NADH. However, when the microsomal preparations were combined with the cytosolic enzyme, a significant N-oxide reductase activity was observed in the presence of the reduced pyridine nucleotide. The activity was enhanced by FAD or methyl viologen. Cytosol alone supplemented with NADPH or NADH exhibited only a slight, but when combined with microsomes, a significant N-oxide reductase activity. Based on these facts, we propose a new electron transfer system consisting of NADPH-cytochrome c reductase and aldehyde oxidase, which exhibits nicotinamide N-oxide reductase activity in the presence of the reduced pyridine nucleotide.  相似文献   

12.
The mitochondrial nicotinamide nucleotide transhydrogenase catalyzes hydride ion transfer between NAD(H) and NADP(H) in a reaction that is coupled to proton translocation across the inner mitochondrial membrane. The enzyme (1043 residues) is composed of an N-terminal hydrophilic segment (approximately 400 residues long) which binds NAD(H), a C-terminal hydrophilic segment (approximately 200 residues long) which binds NADP(H), and a central hydrophobic segment (approximately 400 residues long) which appears to form about 14 membrane-intercalating clusters of approximately 20 residues each. Substrate modulation of transhydrogenase conformation appears to be intimately associated with its mechanism of proton translocation. Using trypsin as a probe of enzyme conformation change, we have shown that NADPH (and to a much lesser extent NADP) binding alters transhydrogenase conformation, resulting in increased susceptibility of several bonds to tryptic hydrolysis. NADH and NAD had little or no effect, and the NADPH concentration for half-maximal enhancement of trypsin sensitivity of transhydrogenase activity (35 microM) was close to the Km of the enzyme for NADPH. The NADPH-promoted trypsin cleavage sites were located 200-400 residues distant from the NADP(H) binding domain near the C-terminus. For example, NADPH binding greatly increased the trypsin sensitivity of the K410-T411 bond, which is separated from the NADP(H) binding domain by the 400-residue-long membrane-intercalating segment. It also enhanced the tryptic cleavage of the R602-L603 bond, which is located within the central hydrophobic segment. These results, which suggest a protein conformation change as a result of NADPH binding, have been discussed in relation to the mechanism of proton translocation by the transhydrogenase.  相似文献   

13.
We studied the physiological effect of the interconversion between the NAD(H) and NADP(H) coenzyme systems in recombinant Saccharomyces cerevisiae expressing the membrane-bound transhydrogenase from Escherichia coli. Our objective was to determine if the membrane-bound transhydrogenase could work in reoxidation of NADH to NAD+ in S. cerevisiae and thereby reduce glycerol formation during anaerobic fermentation. Membranes isolated from the recombinant strains exhibited reduction of 3-acetylpyridine-NAD+ by NADPH and by NADH in the presence of NADP+, which demonstrated that an active enzyme was present. Unlike the situation in E. coli, however, most of the transhydrogenase activity was not present in the yeast plasma membrane; rather, the enzyme appeared to remain localized in the membrane of the endoplasmic reticulum. During anaerobic glucose fermentation we observed an increase in the formation of 2-oxoglutarate, glycerol, and acetic acid in a strain expressing a high level of transhydrogenase, which indicated that increased NADPH consumption and NADH production occurred. The intracellular concentrations of NADH, NAD+, NADPH, and NADP+ were measured in cells expressing transhydrogenase. The reduction of the NADPH pool indicated that the transhydrogenase transferred reducing equivalents from NADPH to NAD+.  相似文献   

14.
Kinetic measurements indicate that the energy-independent transhydrogenation of 3-acetylpyridine-NAD+ by NADPH in membranes of Escherichia coli follows a rapid equilibrium random bireactant mechanism. Each substrate, although reacting preferentially with its own binding site, is able to interact with the binding site of the other substrate to cause inhibition of enzyme activity. 5'-AMP (and ADP) and 2'-AMP interact with the NAD+- and NADP+-binding sites, respectively. Phenylglyoxal and 2,3-butanedione in borate buffer inhibit transhydrogenase activity presumably by reacting with arginyl residues. Protection against inhibition by 2,3-butanedione is afforded by NADP+, NAD+, and high concentrations of NADPH and NADH. Low concentrations of NADPH and NADH increase the rate of inhibition by 2,3-butanedione. Similar effects are observed for the inactivation of the transhydrogenase by tryptic digestion in the presence of these coenzymes. It is concluded that there are at least two conformations of the active site of the transhydrogenase which differ in the extent to which arginyl residues are accessible to exogenous agents such as trypsin and 2,3-butanedione. One conformation is induced by low concentrations of NADH and NADPH. Under these conditions the coenzymes could be reacting at the active site or at an allosteric site. The stimulation of transhydrogenase activity by low concentrations of the NADH is consistent with the latter possibility.  相似文献   

15.
Purified nicotinamide-nucleotide transhydrogenase from beef heart mitochondria was co-reconstituted with bacteriorhodopsin to from transhydrogenase-bacteriorhodopsin vesicles that catalyze a 20-fold light-dependent and uncoupler-sensitive stimulation of the reduction of NADP+ and NADP+ analogs by NADH and a 50-fold shift of the nicotinamide nucleotide ratio. In the presence of light, the transhydrogenase-bacteriorhodopsin vesicles catalyzed a pronounced light intensity-dependent inward proton pumping as indicated by a pH shift of the medium. As indicated by pH shifts, proton pumping by the bacteriorhodopsin essentially paralleled the light-driven transhydrogenase. Addition of valinomycin increased the pH shift twice with a concomitant 50% inhibition of the light-driven transhydrogenase, whereas nigericin inhibited the pH shift completely and the light-driven transhydrogenase partially. Taken together, these results suggest that transhydrogenase and bacteriorhodopsin interact through a delocalized proton-motive force. Possible partial reactions of transhydrogenase were investigated with transhydrogenase-bacteriorhodopsin vesicles energized by light. Reduction of oxidized 3-acetylpyridine adenine dinucleotide by NADH, previously claimed to represent partial reactions, was found to require NADPH. Similarly, reduction of thio-NADP+ by NADPH required NADH. It is concluded that these reactions do not represent partial reactions.  相似文献   

16.
Modification of mitochondrial nicotinamide nucleotide transhydrogenase (NADPH: NAD+ oxidoreductase, EC 1.6.1.1) with 4-chloro-7-nitrobenzo-2-oxa-1,3-diazole (NBD-Cl), followed by measurement of the absorption or fluorescence of the transhydrogenase-NBD adducts, resulted in a biphasic labelling of approx. 4-6 sulfhydryls, presumably cysteine residues. Of these 1-2 (27%) were fast-reacting and 3-4 (73%) slow-reacting sulfhydryls. In the presence of substrates, e.g., NADPH, the labelling was monophasic and all sulfhydryls were fast-reacting, suggesting that the modified sulfhydryls are predominantly localized peripheral to the NAD(P)(H)-binding sites. The rates of modification allowed the calculation of the rate constants for each phase of the labelling. Both in the absence and in the presence of a substrate, e.g., NADPH, the extent of labelling essentially parallelled the inhibition of transhydrogenase activity. Attempts to reactivate transhydrogenase by reduction of labelled sulfhydryls were not successful. Photo-induced transfer of the NBD adduct in partially inhibited transhydrogenase, from the sulfhydryls to reactive NH2 groups of amino-acid residue(s), identified as lysine residue(s), was parallelled by an inhibition of the residual transhydrogenase activity. It is suggested that a lysine localized close to the fast-reacting NBD-Cl-reactive sulfhydryl groups is essential for activity.  相似文献   

17.
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   

18.
The NADH and NADPH ferricyanide reductase activities present in mitochondrial NADH-CoQ reductase preparations have been studied utilizing two photoaffinity pyridine nucleotide analogues: arylazido--alanyl NAD+ (A3-O-{3-[N-(4-azido-2-nitrophenyl)amino]propionyl}NAD+) and arylazido--alanyl NADP+ (N3-O-{3-[N-(4-azido-3-nitrophenyl)amino]-propionyl}NADP+). For the NADH-K3Fe(CN)6 reductase activity, arylazido--alanyl NAD+ was found to be, in the dark, a competitive inhibitor with respect to both NADH and K3Fe(CN)6 withK i,app values of 9.7 and 15.5 µM, respectively. In comparison the NADP+ analogue exhibited weak noncompetitive inhibitor activity for this reaction against both substrates. Upon photoirradiation arylazido--alanyl NAD+ inhibited NADH-K3Fe(CN)6 reductase up to 70% in the presence of a 25-fold molar excess of analogue over the enzyme concentration. This photodependent inhibition could be prevented by the presence, during irradiation, of the natural substrate NADH. In contrast complex kinetic results were obtained with studies of the effects of the pyridine nucleotide analogues of NADPH-K3Fe(CN)6 reductase activity in the dark. Photoirradiation of either analogue in the presence of the enzyme complex resulted in an activation of NADPH-dependent activity. The possibility that the NADPH-K3Fe(CN)6 reductase activity of complex I represents a summation of the combined ferricyanide reductase activity of the NADPH-NAD+ transhydrogenase and NADH oxidoreductase is suggested.  相似文献   

19.
Pyridine dinucleotide transhydrogenase of the Rhodospirillum rubrum chromatophore membrane was readily resolved by a washing procedure into two inactive components, a soluble transhydrogenase factor protein and an insoluble membrane-bound factor. Transhydrogenation was reconstituted on reassociation of these components. The capacity of the membrane factor to reconstitute enzymatic activity was lost after proteolysis of soluble transhydrogenase factor-depleted membranes with trypsin. NADP+ or NADPH, but neither NAD+ nor NADH, stimulated by several fold the rate of trypsin-dependent inactivation of the membrane factor. Substantial protection of the membrane factor from proteolytic inactivation was observed in the presence of Mg2+ ions, an inhibitor of transhydrogenation, or when the soluble transhydrogenase factor was bound to the membrane. Coincident with the loss of enzymatic reconstitutive capacity of the membrane factor was a loss in the ability of the membranes to bind the soluble transhydrogenase factor in a stable complex. The membrane component was inactivated by preincubating soluble transhydrogenase factor-depleted membranes at temperatures above 45 degrees. NADP+, NADPH, or Mg2+ ions, but neither NAD+ nor NADH, protected against inactivation. These studies indicate that (a) the binding of NADP+ or NADPH to the membrane factor promotes a conformational alteration in the protein such that its themostability and susceptibility to proteolysis are increased, and (b) the inhibitory Mg2+ ion-binding site resides in the membrane component.  相似文献   

20.
1. NADH-ubiquinone-1 and NADH-menadione reductase activities of Complex I were inhibited by diphenyleneiodonium (apparent Ki 23 and 30 nmol/mg of protein respectively). Reduction of K3Fe(CN)6 and juglone was relatively unaffected. 2. Iodoniumdiphenyl and derivatives were much less effective inhibitors. Compounds with similar ring structures to diphenyleneiodonium, in particular dibenzofuran, were inhibitors of NADH-ubiquinone-1 oxidoreductase. 3. Diphenylene[125I]iodonium specifically labelled a polypeptide of mol.wt. 23500. Maximum incorporation was 1 mol/mol of Complex-I flavin or 1 mol/mol of the 23500-mol.wt. polypeptide. 4. The label associated with this polypeptide was of limited stability, especially at lower pH. 5. Complete inhibition of ubiquinone reduction was achieved when 1 mol of inhibitor was incorporated/mol of Complex-I flavin, but the relationship between inhibition and labelling was not linear. 6. No evidence for covalent interaction between diphenyleneiodonium and the phospholipids of Complex I was obtained. 7. Rotenone increased the apparent affinity of diphenyleneiodonium for the 23500-mol.wt. polypeptide without affecting the maximum incorporation. 8. The 23500-mol.wt. polypeptide was not solubilized by chaotropic agents. Prior treatment of Complex I with chaotropic agents or sodium dodecyl sulphate prevented incorporation of diphenyleneiodonium into this polypeptide.  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