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 15 毫秒
1.
3-Nitropropionic acid (3-NPA), an inhibitor of succinate dehydrogenase (SDH) at complex II of the mitochondrial electron transport chain induces cellular energy deficit and oxidative stress-related neurotoxicity. In the present study, we identified the site of reactive oxygen species production in mitochondria. 3-NPA increased O2- generation in mitochondria respiring on the complex I substrates pyruvate+malate, an effect fully inhibited by rotenone. Antimycin A increased O2- production in the presence of complex I and/or II substrates. Addition of 3-NPA markedly increased antimycin A-induced O2- production by mitochondria incubated with complex I substrates, but 3-NPA inhibited O2- formation driven with the complex II substrate succinate. At 0.6 microM, myxothiazol inhibits complex III, but only partially decreases complex I activity, and allowed 3-NPA-induced O2- formation; however, at 40 microM myxothiazol (which completely inhibits both complexes I and III) eliminated O2- production from mitochondria respiring via complex I substrates. These results indicate that in the presence of 3-NPA, mitochondria generate O2- from a site between the ubiquinol pool and the 3-NPA block in the respiratory complex II.  相似文献   

2.
S-nitrosation of mitochondrial proteins has been proposed to contribute to the pathophysiological interactions of nitric oxide (NO) and its derivatives with mitochondria but has not been shown directly. Furthermore, little is known about the mechanism of formation or the fate of these putative S-nitrosothiols. Here we have determined whether mitochondrial membrane protein thiols can be S-nitrosated on exposure to free NO from 3,3-bis(aminoethyl)-1-hydroxy-2-oxo-1-triazene (DETA-NONOate) by interaction with S-nitrosoglutathione or S-nitroso-N-acetylpenicillamine (SNAP) and by the NO derivative peroxynitrite. S-Nitrosation of protein thiols was measured directly by chemiluminescence detection. S-Nitrosoglutathione and S-nitroso-N-acetylpenicillamine led to extensive protein thiol oxidation, with about 30% of the modified protein thiols persistently S-nitrosated. In contrast, there was no protein thiol oxidation or S-nitrosation on exposure to 3,3-bis (aminoethyl)-1-hydroxy-2-oxo-1-triazene. Peroxynitrite extensively oxidized protein thiols but produced negligible amounts of S-nitrosothiols. Therefore, mitochondrial membrane protein thiols are S-nitrosated by preformed S-nitrosothiols but not by NO or by peroxynitrite. These S-nitrosated protein thiols were readily reduced by glutathione, so S-nitrosation will only persist when the mitochondrial glutathione pool is oxidized. Respiratory chain complex I was S-nitrosated by S-nitrosothiols, consistent with it being an important target for S-nitrosation during nitrosative stress. The S-nitrosation of complex I correlated with a significant loss of activity that was reversed by thiol reductants. S-Nitrosation was also associated with increased superoxide production from complex I. These findings point to a significant role for complex I S-nitrosation and consequent dysfunction during nitrosative stress in disorders such as Parkinson disease and sepsis.  相似文献   

3.
We have determined the underlying sites of H(2)O(2) generation by isolated rat brain mitochondria and how these can shift depending on the presence of respiratory substrates, electron transport chain modulators and exposure to stressors. H(2)O(2) production was determined using the fluorogenic Amplex red and peroxidase system. H(2)O(2) production was higher when succinate was used as a respiratory substrate than with another FAD-dependent substrate, alpha-glycerophosphate, or with the NAD-dependent substrates, glutamate/malate. Depolarization by the uncoupler p-trifluoromethoxyphenylhydrazone decreased H(2)O(2) production stimulated by all respiratory substrates. H(2)O(2) production supported by succinate during reverse transfer of electrons was decreased by inhibitors of complex I (rotenone and diphenyleneiodonium) whereas in glutamate/malate-oxidizing mitochondria diphenyleneiodonium decreased while rotenone increased H(2)O(2) generation. The complex III inhibitors antimycin and myxothiazol decreased succinate-induced H(2)O(2) production but stimulated H(2)O(2) production in glutamate/malate-oxidizing mitochondria. Antimycin and myxothiazol also increased H(2)O(2) production in mitochondria using alpha-glycerophosphate as a respiratory substrate. In substrate/inhibitor experiments maximal stimulation of H(2)O(2) production by complex I was observed with the alpha-glycerophosphate/antimycin combination. In addition, three forms of in vitro mitochondrial stress were studied: Ca(2+) overload, cold storage for more than 24 h and cytochrome c depletion. In each case we observed (i) a decrease in succinate-supported H(2)O(2) production by complex I and an increase in succinate-supported H(2)O(2) production by complex III, (ii) increased glutamate/malate-induced H(2)O(2) generation by complex I and (iii) increased alpha-glycerophosphate-supported H(2)O(2) generation by complex III. Our results suggest that all three forms of mitochondrial stress resulted in similar shifts in the localization of sites of H(2)O(2) generation and that, in both normal and stressed states, the level and location of H(2)O(2) production depend on the predominant energetic substrate.  相似文献   

4.
Mitochondria generate reactive oxygen species (ROS) dependent on substrate conditions, O(2) concentration, redox state, and activity of the mitochondrial complexes. It is well known that the FADH(2)-linked substrate succinate induces reverse electron flow to complex I of the electron transport chain and that this process generates superoxide (O(2)(*-)); these effects are blocked by the complex I blocker rotenone. We demonstrated recently that succinate + rotenone-dependent H(2)O(2) production in isolated mitochondria increased mildly on activation of the putative big mitochondrial Ca(2+)-sensitive K(+) channel (mtBK(Ca)) by low concentrations of 1,3-dihydro-1-[2-hydroxy-5-(trifluoromethyl)phenyl]-5-(trifluoromethyl)-2H-benzimidazol-2-one (NS-1619). In the present study we examined effects of NS-1619 on mitochondrial O(2) consumption, membrane potential (DeltaPsi(m)), H(2)O(2) release rates, and redox state in isolated guinea pig heart mitochondria respiring on succinate but without rotenone. NS-1619 (30 microM) increased state 2 and state 4 respiration by 26 +/- 4% and 14 +/- 4%, respectively; this increase was abolished by the BK(Ca) channel blocker paxilline (5 microM). Paxilline alone had no effect on respiration. NS-1619 did not alter DeltaPsi(m) or redox state but decreased H(2)O(2) production by 73% vs. control; this effect was incompletely inhibited by paxilline. We conclude that under substrate conditions that allow reverse electron flow, matrix K(+) influx through mtBK(Ca) channels reduces mitochondrial H(2)O(2) production by accelerating forward electron flow. Our prior study showed that NS-1619 induced an increase in H(2)O(2) production with blocked reverse electron flow. The present results suggest that NS-1619-induced matrix K(+) influx increases forward electron flow despite the high reverse electron flow, and emphasize the importance of substrate conditions on interpretation of effects on mitochondrial bioenergetics.  相似文献   

5.
Characteristics of reactive oxygen species (ROS) production in isolated guinea-pig brain mitochondria respiring on alpha-glycerophosphate (alpha-GP) were investigated and compared with those supported by succinate. Mitochondria established a membrane potential (DeltaPsi(m)) and released H(2)O(2) in parallel with an increase in NAD(P)H fluorescence in the presence of alpha-GP (5-40 mm). H(2)O(2) formation and the increase in NAD(P)H level were inhibited by rotenone, ADP or FCCP, respectively, being consistent with a reverse electron transfer (RET). The residual H(2)O(2) formation in the presence of FCCP was stimulated by myxothiazol in mitochondria supported by alpha-GP, but not by succinate. ROS under these conditions are most likely to be derived from alpha-GP-dehydrogenase. In addition, huge ROS formation could be provoked by antimycin in alpha-GP-supported mitochondria, which was prevented by myxothiazol, pointing to the generation of ROS at the quinol-oxidizing center (Q(o)) site of complex III. FCCP further stimulated the production of ROS to the highest rate that we observed in this study. We suggest that the metabolism of alpha-GP leads to ROS generation primarily by complex I in RET, and in addition a significant ROS formation could be ascribed to alpha-GP-dehydrogenase in mammalian brain mitochondria. ROS generation by alpha-GP at complex III is evident only when this complex is inhibited by antimycin.  相似文献   

6.
We measured production of reactive oxygen species by intact mitochondria from rat skeletal muscle, heart, and liver under various experimental conditions. By using different substrates and inhibitors, we determined the sites of production (which complexes in the electron transport chain produced superoxide). By measuring hydrogen peroxide production in the absence and presence of exogenous superoxide dismutase, we established the topology of superoxide production (on which side of the mitochondrial inner membrane superoxide was produced). Mitochondria did not release measurable amounts of superoxide or hydrogen peroxide when respiring on complex I or complex II substrates. Mitochondria from skeletal muscle or heart generated significant amounts of superoxide from complex I when respiring on palmitoyl carnitine. They produced superoxide at considerable rates in the presence of various inhibitors of the electron transport chain. Complex I (and perhaps the fatty acid oxidation electron transfer flavoprotein and its oxidoreductase) released superoxide on the matrix side of the inner membrane, whereas center o of complex III released superoxide on the cytoplasmic side. These results do not support the idea that mitochondria produce considerable amounts of reactive oxygen species under physiological conditions. Our upper estimate of the proportion of electron flow giving rise to hydrogen peroxide with palmitoyl carnitine as substrate (0.15%) is more than an order of magnitude lower than commonly cited values. We observed no difference in the rate of hydrogen peroxide production between rat and pigeon heart mitochondria respiring on complex I substrates. However, when complex I was fully reduced using rotenone, rat mitochondria released significantly more hydrogen peroxide than pigeon mitochondria. This difference was solely due to an elevated concentration of complex I in rat compared with pigeon heart mitochondria.  相似文献   

7.
The mitochondrial respiratory chain is a major source of reactive oxygen species (ROS) under pathological conditions including myocardial ischemia and reperfusion. Limitation of electron transport by the inhibitor rotenone immediately before ischemia decreases the production of ROS in cardiac myocytes and reduces damage to mitochondria. We asked if ROS generation by intact mitochondria during the oxidation of complex I substrates (glutamate, pyruvate/malate) occurred from complex I or III. ROS production by mitochondria of Sprague-Dawley rat hearts and corresponding submitochondrial particles was studied. ROS were measured as H2O2 using the amplex red assay. In mitochondria oxidizing complex I substrates, rotenone inhibition did not increase H2O2. Oxidation of complex I or II substrates in the presence of antimycin A markedly increased H2O2. Rotenone prevented antimycin A-induced H2O2 production in mitochondria with complex I substrates but not with complex II substrates. Catalase scavenged H2O2. In contrast to intact mitochondria, blockade of complex I with rotenone markedly increased H2O2 production from submitochondrial particles oxidizing the complex I substrate NADH. ROS are produced from complex I by the NADH dehydrogenase located in the matrix side of the inner membrane and are dissipated in mitochondria by matrix antioxidant defense. However, in submitochondrial particles devoid of antioxidant defense ROS from complex I are available for detection. In mitochondria, complex III is the principal site for ROS generation during the oxidation of complex I substrates, and rotenone protects by limiting electron flow into complex III.  相似文献   

8.
We have investigated the consequences of permeability transition pore (PTP) opening on the rate of production of reactive oxygen species in isolated rat liver mitochondria. We found that PTP opening fully inhibited H(2)O(2) production when mitochondria were energized both with complex I or II substrates. Because PTP opening led to mitochondrial pyridine nucleotide depletion, H(2)O(2) production was measured again in the presence of various amounts of NADH. PTP opening-induced H(2)O(2) production began when NADH concentration was higher than 50 microm and reached a maximum at over 300 microm. At such concentrations of NADH, the maximal H(2)O(2) production was 4-fold higher than that observed when mitochondria were permeabilized with the channel-forming antibiotic alamethicin, indicating that the PTP opening-induced H(2)O(2) production was not due to antioxidant depletion. Moreover, PTP opening decreased rotenone-sensitive NADH ubiquinone reductase activity, whereas it did not affect the NADH FeCN reductase activity. We conclude that PTP opening induces a specific conformational change of complex I that (i) dramatically increases H(2)O(2) production so long as electrons are provided to complex I, and (ii) inhibits the physiological pathway of electrons inside complex I. These data allowed the identification of a novel consequence of permeability transition that may partly account for the mechanism by which PTP opening induces cell death.  相似文献   

9.
Despite the considerable interest in superoxide as a potential cause of pathology, the mechanisms of its deleterious production by mitochondria remain poorly understood. Previous studies in purified mitochondria have found that the highest rates of superoxide production are observed with succinate-driven reverse-electron transfer through complex I, although the physiological importance of this pathway is disputed because it necessitates high concentrations of succinate and is thought not to occur when NAD is in the reduced state. However, very few studies have examined the rates of superoxide production with mitochondria respiring on both NADH-linked (e.g. glutamate) and complex II-linked substrates. In the present study, we find that the rates of superoxide production (measured indirectly as H2O2) with glutamate+succinate (approximately 1100 pmol of H2O2 x min(-1) x mg(-1)) were unexpectedly much higher than with succinate (approximately 400 pmol of H2O2 x min(-1) x mg(-1)) or glutamate (approximately 80 pmol of H2O2 x min(-1) x mg(-1)) alone. Superoxide production with glutamate+succinate remained high even at low substrate concentrations (<1 mM), was decreased by rotenone and was completely eliminated by FCCP (carbonyl cyanide p-trifluoromethoxyphenylhydrazone), indicating that it must in large part originate from reverse-electron transfer through complex I. Similar results were obtained when glutamate was replaced with pyruvate, alpha-ketoglutarate or palmitoyl carnitine. In contrast, superoxide production was consistently lowered by the addition of malate (malate+succinate approximately 30 pmol of H2O2 x min(-1) x mg(-1)). We propose that the inhibitory action of malate on superoxide production can be explained by oxaloacetate inhibition of complex II. In summary, the present results indicate that reverse-electron transfer-mediated superoxide production can occur under physiologically realistic substrate conditions and suggest that oxaloacetate inhibition of complex II may be an adaptive mechanism to minimize this.  相似文献   

10.
Mitochondrial membrane fatty acid composition has been proposed to play a role in determining mitochondrial proton leak rate. The purpose of this study was to determine if feeding rats diets with different fatty acid sources produces changes in liver proton leak and H(2)O(2) production. Six-month-old male FBNF(1) rats were fed diets with a primary fat source of either corn or fish oil for a 6-month period. As expected, diet manipulations produced substantial differences in mitochondrial fatty acid composition. These changes were most striking for 20:4n6 and 22:6n3. However, proton leak and phosphorylation kinetics as well as lipid and protein oxidative damage were not different (P > 0.10) between fish and corn oil groups. Metabolic control analysis, however, did show that control of both substrate oxidation and phosphorylation was shifted away from substrate oxidation reactions to increased control by phosphorylation reactions in fish versus corn oil groups. Increased mitochondrial H(2)O(2) production was observed in corn versus fish oil-fed rats when mitochondria were respiring on succinate alone or on either succinate or pyruvate/malate in the presence of antimycin A. These results show that mitochondrial H(2)O(2) production and the regulation of oxidative phosphorylation are altered in liver mitochondria from rats consuming diets with either fish or corn oil as the primary lipid source.  相似文献   

11.
Santiago AP  Chaves EA  Oliveira MF  Galina A 《Biochimie》2008,90(10):1566-1577
Mitochondrial hexokinase (mt-HK) and creatine kinase (mt-CK) activities have been recently proposed to reduce the rate of mitochondrial ROS generation through an ADP re-cycling mechanism. Here, we determined the role of mt-HK and mt-CK activities in regulate mitochondrial ROS generation in rat brain, kidney, heart and liver, relating them to the levels of classical antioxidant enzymes. The activities of both kinases were significantly higher in the brain than in other tissues, whereas the activities of catalase (CAT), glutathione peroxidase (GPx) and glutathione reductase (GR) were higher in both liver and kidney mitochondria. In contrast, manganese superoxide dismutase (Mn-SOD) activity was not significantly different among these tissues. Activation of mitochondrial kinases by addition of their substrates increased the ADP re-cycling and thus the respiration by enhancing the oxidative phosphorylation. Succinate induced hydrogen peroxide (H(2)O(2)) generation was higher in brain than in kidney and heart mitochondria, and the lowest in liver mitochondria. Mitochondrial membrane potential (DeltaPsi(m)) and H(2)O(2) production, decreased with additions of 2-DOG or Cr to respiring brain and kidney mitochondria but not to liver. The inhibition of H(2)O(2) production by 2-DOG and Cr correspond to almost 100% in rat brain and about 70% in kidney mitochondria. Together our data suggest that mitochondrial kinases activities are potent preventive antioxidant mechanism in mitochondria with low peroxidase activities, complementing the classical antioxidant enzymes against oxidative stress.  相似文献   

12.
Reactive oxygen species (ROS) generation in mitochondria as a side product of electron and proton transport through the inner membrane is important for normal cell operation as well as development of pathology. Matrix and cytosol alkalization stabilizes semiquinone radical, a potential superoxide producer, and we hypothesized that proton deficiency under the excess of electron donors enhances reactive oxygen species generation. We tested this hypothesis by measuring pH dependence of reactive oxygen species released by mitochondria. The experiments were performed in the media with pH varying from 6 to 8 in the presence of complex II substrate succinate or under more physiological conditions with complex I substrates glutamate and malate. Matrix pH was manipulated by inorganic phosphate, nigericine, and low concentrations of uncoupler or valinomycin. We found that high pH strongly increased the rate of free radical generation in all of the conditions studied, even when DeltapH=0 in the presence of nigericin. In the absence of inorganic phosphate, when the matrix was the most alkaline, pH shift in the medium above 7 induced permeability transition accompanied by the decrease of ROS production. ROS production increase induced by the alkalization of medium was observed with intact respiring mitochondria as well as in the presence of complex I inhibitor rotenone, which enhanced reactive oxygen species release. The phenomena revealed in this report are important for understanding mechanisms governing mitochondrial production of reactive oxygen species, in particular that related with uncoupling proteins.  相似文献   

13.
We studied the effects of adrenaline administration and depletion (induced by reserpine) on rat liver oxidative metabolism. We showed that adrenaline increases, and reserpine decreases aerobic capacity (inferred by cytochrome oxidase activity) in tissue modifying the hepatic content of mitochondrial proteins without changing mitochondrial aerobic capacity. The changes in tissue cytochrome oxidase activity, which agreed with the expression levels of factors involved in mitochondrial biogenesis, such as PGC-1, NRF-1, and NRF-2, were associated with similar changes in tissue and mitochondrial State 3 respiration. Adrenaline and reserpine induced extensive lipid and protein oxidative damage in tissue and mitochondria. The increase in H2O2 release by respiring mitochondria and the decrease in the activities of the antioxidant enzymes glutathione peroxidase and reductase contributed to the reserpine effect on oxidative damage. The adrenaline effect is more difficult to explain, since the hormone increased the antioxidant enzyme activities but, in respiring mitochondria, increased ROS release rate in the presence of succinate and decreased it in the presence of pyruvate/malate. These opposite changes were due to the increased content of the autoxidizable electron carrier located at complex III and decreased content of that located at complex I. Our data suggest that adrenaline can be involved in the mitochondrial population adaptation which verify in conditions in which an increased body energy expenditure verify such as cold exposure.  相似文献   

14.
Characterization of superoxide-producing sites in isolated brain mitochondria   总被引:17,自引:0,他引:17  
Mitochondrial respiratory chain complexes I and III have been shown to produce superoxide but the exact contribution and localization of individual sites have remained unclear. We approached this question investigating the effects of oxygen, substrates, inhibitors, and of the NAD+/NADH redox couple on H2O2 and superoxide production of isolated mitochondria from rat and human brain. Although rat brain mitochondria in the presence of glutamate+malate alone do generate only small amounts of H2O2 (0.04 +/- 0.02 nmol H2O2/min/mg), a substantial production is observed after the addition of the complex I inhibitor rotenone (0.68 +/- 0.25 nmol H2O2/min/mg) or in the presence of the respiratory substrate succinate alone (0.80 +/- 0.27 nmol H2O2/min/mg). The maximal rate of H2O2 generation by respiratory chain complex III observed in the presence of antimycin A was considerably lower (0.14 +/- 0.07 nmol H2O2/min/mg). Similar observations were made for mitochondria isolated from human parahippocampal gyrus. This is an indication that most of the superoxide radicals are produced at complex I and that high rates of production of reactive oxygen species are features of respiratory chain-inhibited mitochondria and of reversed electron flow, respectively. We determined the redox potential of the superoxide production site at complex I to be equal to -295 mV. This and the sensitivity to inhibitors suggest that the site of superoxide generation at complex I is most likely the flavine mononucleotide moiety. Because short-term incubation of rat brain mitochondria with H2O2 induced increased H2O2 production at this site we propose that reactive oxygen species can activate a self-accelerating vicious cycle causing mitochondrial damage and neuronal cell death.  相似文献   

15.
Release of H(2)O(2) in response to Ca(2+) loads (1-100 microM) was investigated using Amplex red fluorescent assay in isolated guinea-pig brain mitochondria respiring on glutamate plus malate or succinate. In mitochondria challenged with Ca(2+) (10 microM), in the absence of adenine nucleotides and inhibitors of the respiratory chain, the rate of H(2)O(2) release, taken as an indication of H(2)O(2) production, was decreased by 21.8+/-1.6% in the presence of NADH-linked substrates and by 86.5+/-1.8% with succinate. Parallel with this, a Ca(2+)-induced loss in NAD(P)H fluorescence, sustained depolarization, decrease in fluorescent light scattering signal and in calcein fluorescence were detected indicating an increased permeability and swelling of mitochondria, which were prevented by ADP (2 mM). In the presence of ADP H(2)O(2) release from mitochondria was decreased, but Ca(2+) no longer influenced the generation of H(2)O(2). We suggest that the decreased H(2)O(2) generation induced by Ca(2+) is related to depolarization and NAD(P)H loss resulting from a non-specific permeability increase of the mitochondrial inner membrane.  相似文献   

16.
Thymocytes undergo negative and positive selection during development in the thymus. During this selection process, the majority of thymocytes are eliminated by apoptosis through signaling via TCR or die by neglect, possibly mediated through glucocorticoids. In this study, we report that thymocytes require molecular oxygen to undergo apoptosis induced by dexamethasone (DEX), a synthetic glucocorticoid, and treatment with N-acetyl-L-cysteine (NAC), a thiol antioxidant, inhibits thymocyte apoptosis in vivo as well as ex vivo. We detected elevated intracellular levels of hydrogen peroxide (H(2)O(2)) during DEX-induced apoptosis, which is reduced by NAC treatment, indicating that the elevated levels of intracellular H(2)O(2) are proapoptotic. We also show that loss of mitochondrial membrane potential, cytochrome c release, as well as caspase-3 activation induced by DEX are attenuated by NAC treatment. We identified the production site for H(2)O(2) as the ubiquinone cycle at complex III of mitochondria by using various inhibitors of the mitochondrial electron transport chain, and we show that the cell death events mediated by mitochondria are also significantly reduced when the inhibitors were used. Through inhibition of the proteasome, we also show that the production of H(2)O(2) and the cell death events mediated by mitochondria are regulated by proteosomal activities in DEX-induced thymocyte apoptosis. We conclude that in DEX-treated thymocytes, the increased production of H(2)O(2) originates from mitochondria and is proapoptotic for cell death mediated by mitochondria. We also conclude that all the apoptotic events mediated by mitochondria are regulated by proteasomes.  相似文献   

17.
Increased production of reactive oxygen species (ROS) by mitochondria is involved in oxidative damage to the organelle and in committing cells to apoptosis or senescence, but the mechanisms of this increase are unknown. Here we show that ROS production by mitochondrial complex I increases in response to oxidation of the mitochondrial glutathione pool. This correlates with thiols on the 51- and 75-kDa subunits of complex I forming mixed disulfides with glutathione. Glutathionylation of complex I increases superoxide production by the complex, and when the mixed disulfides are reduced, superoxide production returns to basal levels. Within intact mitochondria oxidation of the glutathione pool to glutathione disulfide also leads to glutathionylation of complex I, which correlates with increased superoxide formation. In this case, most of this superoxide is converted to hydrogen peroxide, which can then diffuse into the cytoplasm. This mechanism of reversible mitochondrial ROS production suggests how mitochondria might regulate redox signaling and shows how oxidation of the mitochondrial glutathione pool could contribute to the pathological changes that occur to mitochondria during oxidative stress.  相似文献   

18.
Cardiac ischemia decreases complex III activity, cytochrome c content, and respiration through cytochrome oxidase in subsarcolemmal mitochondria (SSM) and interfibrillar mitochondria (IFM). The reversible blockade of electron transport with amobarbital during ischemia protects mitochondrial respiration and decreases myocardial injury during reperfusion. These findings support that mitochondrial damage occurs during ischemia and contributes to myocardial injury during reperfusion. The current study addressed whether ischemic damage to the electron transport chain (ETC) increased the net production of reactive oxygen species (ROS) from mitochondria. SSM and IFM were isolated from 6-mo-old Fisher 344 rat hearts following 25 min global ischemia or following 40 min of perfusion alone as controls. H(2)O(2) release from SSM and IFM was measured using the amplex red assay. With glutamate as a complex I substrate, the net production of H(2)O(2) was increased by 178 +/- 14% and 179 +/- 17% in SSM and IFM (n = 9), respectively, following ischemia compared with controls (n = 8). With succinate as substrate in the presence of rotenone, H(2)O(2) increased by 272 +/- 22% and 171 +/- 21% in SSM and IFM, respectively, after ischemia. Inhibitors of electron transport were used to assess maximal ROS production. Inhibition of complex I with rotenone increased H(2)O(2) production by 179 +/- 24% and 155 +/- 14% in SSM and IFM, respectively, following ischemia. Ischemia also increased the antimycin A-stimulated production of H(2)O(2) from complex III. Thus ischemic damage to the ETC increased both the capacity and the net production of H(2)O(2) from complex I and complex III and sets the stage for an increase in ROS production during reperfusion as a mechanism of cardiac injury.  相似文献   

19.
Interruption of electron flow at the quinone-reducing center (Q(i)) of complex III of the mitochondrial respiratory chain results in superoxide production. Unstable semiquinone bound in quinol-oxidizing center (Q(o)) of complex III is thought to be the sole source of electrons for oxygen reduction; however, the unambiguous evidence is lacking. We investigated the effects of complex III inhibitors antimycin, myxothiazol, and stigmatellin on generation of H(2)O(2) in rat heart and brain mitochondria. In the absence of antimycin A, myxothiazol stimulated H(2)O(2) production by mitochondria oxidizing malate, succinate, or alpha-glycerophosphate. Stigmatellin inhibited H(2)O(2) production induced by myxothiazol. Myxothiazol-induced H(2)O(2) production was dependent on the succinate/fumarate ratio but in a manner different from H(2)O(2) generation induced by antimycin A. We conclude that myxothiazol-induced H(2)O(2) originates from a site located in the complex III Q(o) center but different from the site of H(2)O(2) production inducible by antimycin A.  相似文献   

20.
The production of NO by heart mitochondria was 0.7-1.1 nmol NO/min.mg protein, an activity similar to the ones observed in mitochondrial membranes from other organs. Heart mtNOS seems to contribute with about 56% of the total cellular NO production. The immunological nature of the mtNOS isoform of cardiac tissue remains unclear; in our laboratory, heart mtNOS reacted with an anti-iNOS anti-body. Heart mtNOS expression and activity are regulated by physiological and pharmacological effectors. The state 4/state 3 transition regulates heart mtNOS activity and NO release in intact respiring mitochondria: NO production rates in state 3 were 40% lower than in state 4. Heart mtNOS expression was selectively regulated by O(2) availability in hypobaric conditions and the activity was 20-60% higher in hypoxic rats than in control animals, depending on age. In contrast, NADH-cytochrome c reductase and cytochrome oxidase activities were not affected by hypoxia. The activity of rat heart mtNOS decreased 20% on aging from 12 to 72 weeks of age. On the pharmacological side, mitochondrial NO production was increased after enalapril treatment (the inhibitor of the angiotensin converting enzyme) with modification of heart mtNOS functional activity in the regulation of mitochondrial O(2) uptake and H(2)O(2) production. Thus, heart mtNOS is a highly regulated mitochondrial enzyme, which in turn, plays a regulatory role through mitochondrial NO steady state levels that modulate O(2) uptake and O(2)(-) and H(2)O(2) production rates. Nitric oxide and H(2)O(2) constitute signals for metabolic control that are involved in the regulation of cellular processes, such as proliferation and apoptosis.  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