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 625 毫秒
1.
The respiratory control and rate of oxidation of exogenous NADH in vitro by liver mitochondria from vitamin E deficient rats were studied as a means of providing information concerning possible mitochondrial membrane alterations due to the deficiency.When mitochondria were aged at different temperatures for various periods of time, half-maximal inhibition of respiratory control occurred at lower temperatures and shorter aging periods in deficient mitochondria than in normal ones. Also, respiratory control was lost more rapidly in deficient mitochondria than in normal ones in the presence of either digitonin or low (hypotonic) concentrations of mannitol.Microsomes, both freshly prepared and boiled, dramatically lowered respiratory control and the effect was greater in the deficient mitochondria. Bovine serum albumin overcame the suppressed respiratory control, and exogenously added fatty acids mimiced the action of the microsomes.NADH oxidation by normal mitochondria proceeded slowly in isotonic media, while mitochondria of vitamin E deficient rats oxidized NADH much more rapidly. When mitochondria were subjected to ultrasonic disruption or incubated in hypotonic media, the rates of NADH oxidation by both types of mitochondria were similar.Respiratory decline associated with oxidation of β-hydroxybutyrate by the deficient mitochondria was decreased by including in the medium either a high concentration of NAD+, 0.5 mm oxalacetate, or 2 mm aspartate plus 1 mm α-ketoglutarate. This observation, plus the finding of similar activities of malate dehydrogenase and glutamic-oxalacetic transminase in normal and deficient livers, suggests that the action of each was due to an elevation of the mitochondrial NAD+/NADH ratio via a malate shuttle and cytoplasmic and mitochondrial glutamic-oxalacetate transaminase. It is postulated that the marked mitochondrial respiratory decline in the deficient rats is attributed to a limiting availability of NAD+ and a low ratio of NAD+ to NADH.  相似文献   

2.
Determinations were made for corn (Zea mays L., WF9-Tms × M14) mitochondria of the stoichiometric relationship between K+ transport and bond energy produced in respiration (K+/~ ratio). With inward pumping of potassium acetate activated by NADH oxidation, the initial rate of K+ transported into the sucrose inaccessible space varied between 0.58 and 0.97 K+/~, assuming 2 high energy bond equivalents per NADH oxidized. Only small amounts of H+ were ejected. Valinomycin did not alter the ratio.  相似文献   

3.
Respiration-linked oxidation of 3-hydroxybutyryl-CoA, crotonyl-CoA and saturated fatty acyl (C4, C8 and C14)-CoA esters was studied in different mitochondrial preparations. Oxidation of acyl-CoA esters was poor in intact mitochondria; however, it was significant, as well as, NAD+ and CoA-dependent in gently and in vigorously sonicated mitochondria. The respiration-linked oxidation of crotonyl-CoA and 3-hydroxybutyryl-CoA proceeded at much higher rates (over 700%) in gently disrupted mitochondria than in completely disrupted mitochondria. The redox dye-linked oxidation of crotonyl-CoA (with inhibited respiratory chain) was also higher in gently disrupted mitochondria (149%) than in disrupted ones. During the respiration-linked oxidation of 3-hydroxybutyryl-CoA the steady-state NADH concentrations in the reaction chamber were determined, and found to be 8 μM in gently sonicated and 15 μM in completely sonicated mitochondria in spite of the observation that the gently sonicated mitochondria oxidized the 3-hydroxybutyryl-CoA much faster than the completely sonicated mitochondria. The NAD+-dependence of 3-hydroxybutyryl-CoA oxidation showed that a much smaller NAD+ concentration was enough to half-saturate the reaction in gently disrupted mitochondria than in completely disrupted ones. Thus, these observations indicate the positive kinetic consequence of organization of β-oxidation enzyme in situ. Respiration-linked oxidation of bytyryl-, oxtanoyl- and palmitoyl-CoA was also studied and these CoA intermediates were oxidized at approx. 50% of the rate of crotonyl- and 3-hydroxybutyryl-CoA in the gently disrupted mitochondria. In vigorously disrupted mitochondria the oxidation rate of these saturated acyl-CoA intermediates was hardly detectable indicating that the connection between the acyl-CoA dehydrogenase and the respiratory chain had been disrupted.  相似文献   

4.
Mitochondria were isolated from fresh beetroots (Beta vulgaris L. cvs Rubria and Nina) by differential centrifugation followed by Percoll gradient centrifugation. These purified mitochondria oxidized external NADH, although relatively slowly (20-40 versus 100-120 nanomoles oxygen per minute times milligram protein for NADH and succinate oxidation, respectively), with respiratory control ratios of two to three and ADP/O ratios of 1.2 to 1.6. NADPH was also oxidized, but even more slowly and with little or no coupling. The optimum for both NADH and NADPH oxidation by fresh beetroot mitochondria was pH 6. The rate of external NADH oxidation by isolated mitochondria was enhanced threefold during storage of the intact tubers at 10°C for 12 weeks. The optimum of the induced NADH oxidation was approximately pH 6.8. Succinate and malate oxidation only increased by 30% during the same period and NADPH oxidation was constant. This is strong evidence that NADH and NADPH oxidation are catalyzed by different enzymes at least in beetroots. Activity staining of nondenaturing polyacrylamide gels with NADH and Nitro Blue Tetrazolium did not show differences in banding pattern between mitochondria isolated from fresh and stored beetroots. The induction is discussed in relation to physiological aging processes.  相似文献   

5.
Mitochondria isolated from pea leaves (Pisum sativum L. var Massey Gem) and purified on a linear sucrose density gradient were substantially free of contamination by Chl and peroxisomes. They showed high respiratory rates and good respiratory control and ADP/O ratios. Malate, glutamate, succinate, glycine, pyruvate, α-ketoglutarate, NADH, and NADPH were oxidized but little or no oxidation of citrate, isocitrate, or proline was detected. The oxidation of NADPH by the purified mitochondria did not occur via a transhydrogenase or phosphatase converting it to NADH. NADPH oxidation had an absolute requirement for added Ca2+, whereas NADH oxidation proceeded in its absence. In addition, oxidation of the two substrates showed different sensitivities to chelators and sulfhydryl reagents, and faster rates of O2 uptake were observed with both substrates than with either alone. This indicates that the NADPH dehydrogenase is distinct from the exogenous NADH dehydrogenase.  相似文献   

6.
External NADH and succinate were oxidized at similar rates by soybean (Glycine max) cotyledon and leaf mitochondria when the cytochrome chain was operating, but the rate of NADH oxidation via the alternative oxidase was only half that of succinate. However, measurements of the redox poise of the endogenous quinone pool and reduction of added quinones revealed that external NADH reduced them to the same, or greater, extent than did succinate. A kinetic analysis of the relationship between alternative oxidase activity and the redox state of ubiquinone indicated that the degree of ubiquinone reduction during external NADH oxidation was sufficient to fully engage the alternative oxidase. Measurements of NADH oxidation in the presence of succinate showed that the two substrates competed for cytochrome chain activity but not for alternative oxidase activity. Both reduced Q-1 and duroquinone were readily oxidized by the cytochrome oxidase pathway but only slowly by the alternative oxidase pathway in soybean mitochondria. In mitochondria isolated from the thermogenic spadix of Philodendron selloum, on the other hand, quinol oxidation via the alternative oxidase was relatively rapid; in these mitochondria, external NADH was also oxidized readily by the alternative oxidase. Antibodies raised against alternative oxidase proteins from Sauromatum guttatum cross-reacted with proteins of similar molecular size from soybean mitochondria, indicating similarities between the two alternative oxidases. However, it appears that the organization of the respiratory chain in soybean is different, and we suggest that some segregation of electron transport chain components may exist in mitochondria from nonthermogenic plant tissues.  相似文献   

7.
Mitochondria from skeletal muscle, heart and liver of strain 129/ReJ-dy dystrophic mice and their littermate controls were characterized with respect to their respiratory and phosphorylating activities. Skeletal muscle mitochondria from dystrophic mice showed significantly lower state 3 respiratory rates than controls with both pyruvate + malate and succinate as substrates (P < 0.01). ADP/O and Ca2+/O ratios were found to be normal. A decreased rate of NADH oxidation (0.01 <P < 0.05) by sonicated mitochondrial suspensions from dystrophic mice was also seen. High respiratory rates with ascorbate + phenazine methosulfate as substrates indicated that cytochrome oxidase was not rate limiting in the oxidation of either pyruvate + malate or succinate. Skeletal muscle mitochondria from dystrophic mice showed no deficiency in any of the cytochromes or coenzyme Q. Mg2+-stimulated ATPase activity was higher in dystrophic muscle mitochondria than in controls, but basal and oligomycin-insensitive activities were virtually identical to those of controls. A significant reduction in the intramitochondrial NAD+ content (0.01 <P < 0.02) was seen in dystrophic skeletal muscle as compared to controls. Heart mitochondria from dystrophic mice showed similar, though less extensive abnormalities while liver mitochondria were essentially normal. We concluded from these results that skeletal muscle mitochondria from strain 129 dystrophic mice possess impairments in substrate utilization which may result from (1) an abnormality in the transfer of electrons on the substrate side of coenzyme Q in the case of succinate oxidation; (2) a defect on the path of electron flow from NADH to cytochrome c, and (3) a deficiency of NAD+ in the case of NAD+-linked substrates.  相似文献   

8.
This study aims at characterizing NAD(P)H dehydrogenases on the inside and outside of the inner membrane of mitochondria of one phosphoenolpyruvate carboxykinase??crassulacean acid metabolism plant, Hoya carnosa. In crassulacean acid metabolism plants, NADH is produced by malate decarboxylation inside and outside mitochondria. The relative importance of mitochondrial alternative NADH dehydrogenases and their association was determined in intact??and alamethicin??permeabilized mitochondria of H. carnosa to discriminate between internal and external activities. The major findings in H. carnosa mitochondria are: (i) external NADPH oxidation is totally inhibited by DPI and totally dependent on Ca2+, (ii) external NADH oxidation is partially inhibited by DPI and mainly dependent on Ca2+, (iii) total NADH oxidation measured in permeabilized mitochondria is partially inhibited by rotenone and also by DPI, (iv) total NADPH oxidation measured in permeabilized mitochondria is partially dependent on Ca2+ and totally inhibited by DPI. The results suggest that complex I, external NAD(P)H dehydrogenases, and internal NAD(P)H dehydrogenases are all linked to the electron transport chain. Also, the total measurable NAD(P)H dehydrogenases activity was less than the total measurable complex I activity, and both of these enzymes could donate their electrons not only to the cytochrome pathway but also to the alternative pathway. The finding indicated that the H. carnosa mitochondrial electron transport chain is operating in a classical way, partitioning to both Complex I and alternative Alt. NAD(P)H dehydrogenases.  相似文献   

9.
Exogenous NADH oxidation by cauliflower (Brassica oleracea L.) bud mitochondria was sensitive to antimycin A and gave ADP/O ratios of 1.4 to 1.9. In intact mitochondria, NADH-cytochrome c reductase activity was only slightly inhibited by antimycin A. The antimycin-insensitive activity was associated with the outer membrane. Malate oxidation was sensitive to both rotenone and antimycin A and gave ADP/O values of 2.4 to 2.9. However in the presence of added NAD+, malate oxidation displayed similar properties to exogenous NADH oxidation. In both the presence and absence of added NAD+, malate oxidation was dependent on inorganic phosphate and inhibited by 2-n-butyl malonate.  相似文献   

10.
The effect of polyamines (putrescine, spermine, and spermidine) on the oxidation of exogenous NADH by Jerusalem artichoke (Helianthus tuberosus L. cv. OB1) mitochondria, have been studied. Addition of spermine and/or spermidine to a suspension of mitochondria in a low-cation medium (2 millimolar-K+) caused a decrease in the apparent Km and an increase in the apparent Vmax for the oxidation of exogenous NADH. These polycations released by screening effect the mitochondrially induced quenching of 9-aminoacridine fluorescence, their efficiency being dependent on the valency of the cation (C4+ > C3+). Conversely, putrescine only slightly affected both kinetic parameters of exogenous NADH oxidation and the number of fixed charges on the membranes. Spermine and spermidine, but not putrescine, decreased the apparent Km for Ca2+ from about 1 to about 0.2 micromolar, required to activate external NADH oxidation in a high-cation medium, containing physiological concentrations of Pi, Mg2+ and K+. The results are interpreted as evidence for a role of spermine and spermidine in the modulation of exogenous NADH oxidation by plant mitochondria in vivo.  相似文献   

11.
In aerobic microorganisms, the entry point of respiratory electron transfer is represented by the NADH:quinone oxidoreductase. The enzyme couples the oxidation of NADH with the reduction of quinone. In the type 1 NADH:quinone oxidoreductase (Ndh1), this reaction is accompanied by the translocation of cations, such as H+ or Na+. In Escherichia coli, cation translocation is accomplished by the subunit NuoL, thus generating membrane potential (Δψ). Some microorganisms achieve NADH oxidation by the alternative, nonelectrogenic type 2 NADH:quinone oxidoreductase (Ndh2), which is not cation translocating. Since these enzymes had not been described in Staphylococcus aureus, the goal of this study was to identify proteins operating in the NADH:quinone segment of its respiratory chain. We demonstrated that Ndh2 represents a NADH:quinone oxidoreductase in S. aureus. Additionally, we identified a hypothetical protein in S. aureus showing sequence similarity to the proton-translocating subunit NuoL of complex I in E. coli: the NuoL-like protein MpsA. Mutants with deletion of the nuoL-like gene mpsA and its corresponding operon, mpsABC (mps for membrane potential-generating system), exhibited a small-colony-variant-like phenotype and were severely affected in Δψ and oxygen consumption rates. The MpsABC proteins did not confer NADH oxidation activity. Using an Na+/H+ antiporter-deficient E. coli strain, we could show that MpsABC constitute a cation-translocating system capable of Na+ transport. Our study demonstrates that MpsABC represent an important functional system of the respiratory chain of S. aureus that acts as an electrogenic unit responsible for the generation of Δψ.  相似文献   

12.
In rat liver mitochondria, the macrocyclic polyether, dibenzo-18-crown-6 (polyether XXVIII) inhibits the oxidation of NAD-dependent substrates, as stimulated by ADP, uncouplers and valinomycin plus K+. It does not inhibit the oxidation of succinate. It is concluded that polyether XXVIII inhibits electron transfer in the NADH-CoQ span of the respiratory chain. This is a process that is reversed by menadione. Inhibition of oxidation of NAD-dependent substrates in K+-depleted mitochondria induced by the polyether is reversed by concentrations of K+ higher than 60 mM, and also by Li+, a cation that does not complex with polyether XXVIII. As assayed by swelling mitochondria, reversal of the inhibition of electron transfer is accompanied by influx of monovalent cations. Polyether XXVIII also inhibits in submitochondrial particles the aerobic oxidation of NADH, but not that of succinate; this inhibition is also reversed by K+ at high concentrations, and Li+. The data are consistent with the hypothesis that a monovalent cation is required for maximal rates of electron transport in the NADH-CoQ span of the respiratory chain.  相似文献   

13.
Michel Neuburger  Roland Douce 《BBA》1980,589(2):176-189
Mitochondria isolated from spinach leaves oxidized malate by both a NAD+-linked malic enzyme and malate dehydrogenase. In the presence of sodium arsenite the accumulation of oxaloacetate and pyruvate during malate oxidation was strongly dependent on the malate concentration, the pH in the reaction medium and the metabolic state condition.Bicarbonate, especially at alkaline pH, inhibited the decarboxylation of malate by the NAD+-linked malic enzyme in vitro and in vivo. Analysis of the reaction products showed that with 15 mM bicarbonate, spinach leaf mitochondria excreted almost exclusively oxaloacetate.The inhibition by oxaloacetate of malate oxidation by spinach leaf mitochondria was strongly dependent on malate concentration, the pH in the reaction medium and on the metabolic state condition.The data were interpreted as indicating that: (a) the concentration of oxaloacetate on both sides of the inner mitochondrial membrane governed the efflux and influx of oxaloacetate; (b) the NAD+/NADH ratio played an important role in regulating malate oxidation in plant mitochondria; (c) both enzymes (malate dehydrogenase and NAD+-linked malic enzyme) were competing at the level of the pyridine nucleotide pool, and (d) the NAD+-linked malic enzyme provided NADH for the reversal of the reaction catalyzed by the malate dehydrogenase.  相似文献   

14.
A procedure was described for preparing intact mitochondria from spinach (Spinacia oleracea L.) leaves. These mitochondria oxidized succinate, malate, pyruvate, α-ketoglutarate, and NADH with good respiratory control and ADP/O ratios comparable to those observed with mitochondria from other plant tissues. Glycine was oxidized by the preparations. This oxidation linked to the mitochondrial electron transport chain, was coupled to three phosphorylation sites and was sensitive to electron transport and phosphorylation inhibitors.  相似文献   

15.
Inside-out submitochondrial particles from both potato (Solanum tuberosum L. cv. Bintje) tubers and pea (Pisum sativum L. cv. Oregon) leaves possess three distinct dehydrogenase activities: Complex I catalyzes the rotenone-sensitive oxidation of deamino-NADH, NDin(NADPH) catalyzes the rotenone-insensitive and Ca2+-dependent oxidation of NADPH and NDin(NADH) catalyzes the rotenone-insensitive and Ca2+-independent oxidation of NADH. Diphenylene iodonium (DPI) inhibits complex I, NDin(NADPH) and NDin (NADH) activity with a Ki of 3.7, 0.17 and 63 µM, respectively, and the 400-fold difference in Ki between the two NDin made possible the use of DPI inhibition to estimate NDin (NADPH) contribution to malate oxidation by intact mitochondria. The oxidation of malate in the presence of rotenone by intact mitochondria from both species was inhibited by 5 µM DPI. The maximum decrease in rate was 10–20 nmol O2 mg?1 min?1. The reduction level of NAD(P) was manipulated by measuring malate oxidation in state 3 at pH 7.2 and 6.8 and in the presence and absence of an oxaloacetate-removing system. The inhibition by DPI was largest under conditions of high NAD(P) reduction. Control experiments showed that 125 µM DPI had no effect on the activities of malate dehydrogenase (with NADH or NADPH) or malic enzyme (with NAD+ or NADP+) in a matrix extract from either species. Malate dehydrogenase was unable to use NADP+ in the forward reaction. DPI at 125 µM did not have any effect on succinate oxidation by intact mitochondria of either species. We conclude that the inhibition caused by DPI in the presence of rotenone in plant mitochondria oxidizing malate is due to inhibition of NDin(NADPH) oxidizing NADPH. Thus, NADP turnover contributes to malate oxidation by plant mitochondria.  相似文献   

16.
Malate oxidation in plant mitochondria proceeds through the activities of two enzymes: a malate dehydrogenase and a NAD+-dependent malic enzyme. In cauliflower, mitochondria malate oxidation via malate dehydrogenase is rotenone- and cyanide-sensitive. Addition of exogenous NAD+ stimulates the oxidation of malate via malic enzyme and generates an electron flux that is both rotenone- and cyanide-insensitive. The same effects of exogenous NAD+ are also observed with highly cyanide-sensitive mitochondria from white potato tubers or with mitochondria from spinach leaves. Both enzymes are located in the matrix, but some experimental data also suggest that part of malate dehydrogenase activity is also present outside the matrix compartment (adsorbed cytosolic malate dehydrogenase?). It is concluded that malic enzyme and a specific pool of NAD+/NADH are connected to the cyanide-insensitive alternative pathway by a specific rotenone-insensitive NADH dehydrogenase located on the inner face of the inner membrane. Similarly, malate dehydrogenase and another specific pool of NAD+/NADH are connected to the cyanide- (and antimycin-) sensitive pathway by a rotenone-sensitive NADH dehydrogenase located on the inner face of the inner membrane. A general scheme of electron transport in plant mitochondria for the oxidation of malate and NADH can be given, assuming that different pools of ubiquinone act as a branch point between various dehydrogenases, the cyanide-sensitive cytochrome pathway and the cyanide-insensitive alternative pathway.  相似文献   

17.
Shibasaka M  Tsuji H 《Plant physiology》1988,86(4):1008-1012
Respiratory activities were compared among rice seedlings germinated in air for 6 days (aerobic seedlings), those germinated under water for 5 days (submerged seedlings), and those grown in air for 1 day after 5 days' submerged germination (air-adapted seedlings). The respiratory activity of the submerged seedlings increased rapidly on transfer to air and reached a plateau at 16 hours in air. Respiration of the submerged seedlings was as sensitive to cyanide as those of aerobic and air-adapted seedlings. 2,4-Dinitrophenol had no effect on the respiration of the submerged seedlings, but stimulated those of the other two types of seedlings. Mitochondria from three types of seedlings did not differ in the ADP/O ratio and the respiratory control ratio (RCR) when succinate was oxidized. However, mitochondria from submerged seedlings (submerged mitochondria) showed poor RCR of about unity when malate was oxidized. Both the rate of succinate oxidation and succinate dehydrogenase activity were low in submerged mitochondria, but increased during air adaptation. Although submerged mitochondria oxidized malate very slowly, this activity increased after exposure to air without any increase in malate dehydrogenase activity. When NAD+ was added to submerged mitochondria, oxidation of malate was restored to the level of the aerobic controls. Addition of NAD+ enhanced the state 3 rate in submerged mitochondria, and RCR recovered to nearly the same value as that of the aerobic controls. Similar effects of NAD+ on 2-oxoglutarate oxidation were observed. All these defects in submerged mitochondria were repaired during air adaptation. These results suggest that NAD+-linked substrate oxidation was low in submerged mitochondria because of NAD+ deficiency, and that the oxidation increased with an increasing level of NAD+ during air adaptation.  相似文献   

18.
Mitochondria from liver, kidney, brain, and skeletal muscle metabolized acetaldehyde. Acetaldehyde oxidation by liver and kidney mitochondria was maximal at low levels of acetaldehyde and was sensitive to rotenone, suggesting the involvement of a NAD+-dependent aldehyde dehydrogenase with a high affinity for acetaldehyde. Acetaldehyde oxidation was stimulated 50% by ADP, suggesting that, in state 4, reoxidation of NADH is rate limiting for acetaldehyde oxidation. In state 4, acetaldehyde oxidation was decreased by NAD+-dependent substrates, as well as by succinate and ascorbate. The inhibition by the latter two substrates was prevented by ADP, dinitrophenol, valinomycin, and gramicidin, but not by oligomycin. Since these compounds are linked to energy transduction and utilization, the data suggest that the inhibition is mediated via energy-dependent reversed electron transport. In state 3, all of these substrates caused considerably less inhibition of acetaldehyde oxidation, suggesting that the activity of aldehyde dehydrogenase, and not of NADH reoxidation, is probably rate limiting for acetaldehyde oxidation. The ionophores valinomycin and gramicidin stimulated acetaldehyde oxidation to a greater extent than ADP. These ionophores also stimulated acetaldehyde oxidation in the presence of ADP. Stimulation by valinomycin occurred in the presence of monovalent cations transported by this ionophore, e.g., K+, Rb+, Cs+. Stimulation by gramicidin also occurred in the presence of these cations, but did not occur with Na+ or Li+. Na+ prevents the stimulation of acetaldehyde oxidation, which occurs in the presence of gramicidin and K+. The stimulation by valinomycin and gramicidin was energy dependent and required the presence of a permeant anion. In the absence of an ionophore, potassium phosphate had no effect on acetaldehyde oxidation. These data suggest that the oxidation of acetaldehyde by rat liver and kidney mitochondria is influenced by the oxidation-reduction state of the mitochondria and by the cationic environment. With brain and muscle mitochondria, the rate of acetaldehyde oxidation increased two- to threefold as the concentration of acetaldehyde was raised from 0.167 to 0.50 mm. Acetaldehyde oxidation in these mitochondria was also sensitive; to rotenone, indicating dependence on NAD+. ADP, valinomycin, gramicidin, and succinate, compounds which either increased or decreased the rate of acetaldehyde oxidation by liver and kidney mitochondria, had no effect on acetaldehyde oxidation by muscle or brain mitochondria. In state 4, mitochondria from Becker-transplantable hepatocellular carcinoma HC-252 oxidized acetaldehyde at the same rate as liver mitochondria. However, in the presence of ADP, dinitrophenol, valinomycin and gramicidin, the rate of acetaldehyde oxidation by the tumor mitochondria was two to three times greater than that of liver mitochondria, suggesting the presence of a more active; acetaldehyde-oxidizing system in tumor than in liver mitochondria.  相似文献   

19.
In isolated hepatic mitochondria, the oxidation of NAD+-dependent substrates was decreased after chronic consumption of ethanol or by the addition of acetaldehyde in vitro. Reversed electron transport from succinate to NAD?, which requires transfer of electrons through the NADH dehydrogenase complex and energy transduction through coupling site 1, was depressed by ethanol feeding and by acetaldehyde in vitro, whereas NADH formation from glutamate, which is mediated directly by substrate oxidation and is not energy-dependent, was slightly increased. By contrast, reactions involving the terminal portion of the phosphorylation chain, e.g., ATP-32P exchange or dinitrophenolstimulated ATPase activity, were not affected. Adenine nucleotide translocase activity was not altered by chronic consumption of ethanol or the addition of acetaldehyde in vitro. These data suggest that the NADH-ubiquinone oxidoreductase complex of the respiratory chain, a segment which contains several iron-sulfur centers which participate in electron transport and energy transduction, may be impaired by chronic consumption of ethanol and is especially sensitive to inhibition by acetaldehyde in vitro. Neither energy coupling sites II or III, nor the terminal reactions of oxidative phosphorylation share this sensitivity. CO2 production from various labeled intermediates of the citric acid cycle was depressed after chronic consumption of ethanol and after the addition of acetaldehyde. Acetate had no effect on these reactions, indicating that the inhibition by acetaldehyde is not mediated via acetate. Impairment of the activities of the respiratory chain and the citric acid cycle, or both, may explain the decreases in oxygen uptake and CO2 production from citric acid cycle intermediates and fatty acids, as well as the increase in ketone body production, found in mitochondria from ethanolfed rats.  相似文献   

20.
Plant (and fungal) mitochondria contain multiple NAD(P)H dehydrogenases in the inner membrane all of which are connected to the respiratory chain via ubiquinone. On the outer surface, facing the intermembrane space and the cytoplasm, NADH and NADPH are oxidized by what is probably a single low-molecular-weight, nonproton-pumping, unspecific rotenone-insensitive NAD(P)H dehydrogenase. Exogenous NADH oxidation is completely dependent on the presence of free Ca2+ with aK 0.5 of about 1 µM. On the inner surface facing the matrix there are two dehydrogenases: (1) the proton-pumping rotenone-sensitive multisubunit Complex I with properties similar to those of Complex I in mammalian and fungal mitochondria. (2) a rotenone-insensitive NAD(P)H dehydrogenase with equal activity with NADH and NADPH and no proton-pumping activity. The NADPH-oxidizing activity of this enzyme is completely dependent on Ca2+ with aK 0.5 of 3 µM. The enzyme consists of a single subunit of 26 kDa and has a native size of 76 kDa, which means that it may form a trimer.  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