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.
1. Kinetic experiments suggested the possible existence of at least two different NAD(+)-dependent aldehyde dehydrogenases in rat liver. Distribution studies showed that one enzyme, designated enzyme I, was exclusively localized in the mitochondria and that another enzyme, designated enzyme II, was localized in both the mitochondria and the microsomal fraction. 2. A NADP(+)-dependent enzyme was also found in the mitochondria and the microsomal fraction and it is suggested that this enzyme is identical with enzyme II. 3. The K(m) for acetaldehyde was apparently less than 10mum for enzyme I and 0.9-1.7mm for enzyme II. The K(m) for NAD(+) was similar for both enzymes (20-30mum). The K(m) for NADP(+) was 2-3mm and for acetaldehyde 0.5-0.7mm for the NADP(+)-dependent activity. 4. The NAD(+)-dependent enzymes show pH optima between 9 and 10. The highest activity was found in pyrophosphate buffer for both enzymes. In phosphate buffer there was a striking difference in activity between the two enzymes. Compared with the activity in pyrophosphate buffer, the activity of enzyme II was uninfluenced, whereas the activity of enzyme I was very low. 5. The results are compared with those of earlier investigations on the distribution of aldehyde dehydrogenase and with the results from purified enzymes from different sources.  相似文献   

2.
Both the external oxidation of NADH and NADPH in intact potato (Solanum tuberosum L. cv. Bintje) tuber mitochondria and the rotenone-insensitive internal oxidation of NADPH by inside-out submitochondrial particles were dependent on Ca2+. The stimulation was not due to increased permeability of the inner mitochondrial membrane. Neither the membrane potential nor the latencies of NAD(+)-dependent and NADP(+)-dependent malate dehydrogenases were affected by the addition of Ca2+. The pH dependence and kinetics of Ca(2+)-dependent NADPH oxidation by inside-out submitochondrial particles were studied using three different electron acceptors: O2, duroquinone and ferricyanide. Ca2+ increased the activity with all acceptors with a maximum at neutral pH and an additional minor peak at pH 5.8 with O2 and duroquinone. Without Ca2+, the activity was maximal around pH 6. The Km for NADPH was decreased fourfold with ferricyanide and duroquinone, and twofold with O2 as acceptor, upon addition of Ca2+. The Vmax was not changed with ferricyanide as acceptor, but increased twofold with both duroquinone and O2. Half-maximal stimulation of the NADPH oxidation was found at 3 microM free Ca2+ with both O2 and duroquinone as acceptors. This is the first report of a membrane-bound enzyme inside the inner mitochondrial membrane which is directly dependent on micromolar concentrations of Ca2+. Mersalyl and dicumarol, two potent inhibitors of the external NADH dehydrogenase in plant mitochondria, were found to inhibit internal rotenone-insensitive NAD(P)H oxidation, at the same concentrations and in manners very similar to their effects on the external NAD(P)H oxidation.  相似文献   

3.
The opening of the mitochondrial permeability transition pore (PTP) has been suggested to play a key role in various forms of cell death, but direct evidence in intact tissues is still lacking. We found that in the rat heart, 92% of NAD(+) glycohydrolase activity is associated with mitochondria. This activity was not modified by the addition of Triton X-100, although it was abolished by mild treatment with the protease Nagarse, a condition that did not affect the energy-linked properties of mitochondria. The addition of Ca(2+) to isolated rat heart mitochondria resulted in a profound decrease in their NAD(+) content, which followed mitochondrial swelling. Cyclosporin A(CsA), a PTP inhibitor, completely prevented NAD(+) depletion but had no effect on the glycohydrolase activity. Thus, in isolated mitochondria PTP opening makes NAD(+) available for its enzymatic hydrolysis. Perfused rat hearts subjected to global ischemia for 30 min displayed a 30% decrease in tissue NAD(+) content, which was not modified by extending the duration of ischemia. Reperfusion resulted in a more severe reduction of both total and mitochondrial contents of NAD(+), which could be measured in the coronary effluent together with lactate dehydrogenase. The addition of 0.2 microm CsA or of its analogue MeVal-4-Cs (which does not inhibit calcineurin) maintained higher NAD(+) contents, especially in mitochondria, and significantly protected the heart from reperfusion damage, as shown by the reduction in lactate dehydrogenase release. Thus, upon reperfusion after prolonged ischemia, PTP opening in the heart can be documented as a CsA-sensitive release of NAD(+), which is then partly degraded by glycohydrolase and partly released when sarcolemmal integrity is compromised. These results demonstrate that PTP opening is a causative event in reperfusion damage of the heart.  相似文献   

4.
The effects of oxidative stress on DNA damage and associated reactions, increased polyadenosine diphosphate-ribose polymerase (PARP) activity and decreased nicotinamide adenine dinucleotide (NAD) and adenosine triphosphate (ATP) contents, have been tested in primary cultures of porcine aortic endothelial cells. The cells were treated with 50-500 microM H2O2 for 20 min or 100 microM paraquat for 3 days or were exposed to 95% O2 for 2 and 5 days. The administration of 250-500 microM H2O2 resulted in a marked increase in PARP activity and a profound depletion of ATP and NAD. Although hyperoxia had no effect on PARP activity and reduced only slightly the ATP and NAD stores, it markedly reduced the ability of endothelial cells to increase PARP activity upon exposure to DNase. Paraquat had a similar effect. Human dermal fibroblasts were also exposed to 50-500 microM H2O2 for 20 min or 95% O2 for 5 days. Their response to H2O2 differed from that of endothelial cells by their ability to maintain the ATP content at a normal level. Fibroblasts were also insensitive to the effect of hyperoxia. These results suggest that the oxidant-related DNA damage is a function of the type of oxidative stress used and may be cell-specific.  相似文献   

5.
Control of the rate of cardiac cell division by oxygen occurs most probably by altering the redox state of a control substance, e.g. NAD(+)right harpoon over left harpoonNADH. NAD(+) (and not NADH) forms poly(ADP-ribose), an inhibitor of DNA synthesis, in a reaction catalysed by poly(ADP-ribose) polymerase. Lower partial pressure of oxygen, which increases the rate of division, would shift NAD(+)-->NADH, decrease poly(ADP-ribose) synthesis, and increase DNA synthesis. Chick-embryo heart cells grown in culture in 20% O(2) (in which they divide more slowly than in 5% O(2)) did exhibit greater poly(ADP-ribose) polymerase activity (+83%, P<0.001) than when grown in 5% O(2). Reaction product was identified as poly(ADP-ribose) by its insensitivity to deoxyribonuclease, ribonuclease, NAD glycohydrolase, Pronase, trypsin and micrococcal nuclease, and by its complete digestion with snake-venom phosphodiesterase to phosphoribosyl-AMP and AMP. Isolation of these digestion products by Dowex 1 (formate form) column chromatography and paper chromatography allowed calculation of average poly(ADP-ribose) chain length, which was 15-26% greater in 20% than in 5% O(2). Thus in 20% O(2) the increase in poly(ADP-ribose) formation results from chain elongation. Formation of new chains also occurs, probably to an even greater degree than chain elongation. Additionally, poly(ADP-ribose) polymerase has very different K(m) and V(max.) values and pH optima in 20% and 5% O(2). These data suggest that poly(ADP-ribose) metabolism participates in the regulation of heart-cell division by O(2), probably by several different mechanisms.  相似文献   

6.
Sugars and sugar phosphates enable vanadate to catalyze the oxidation of NADH. Superoxide dismutase inhibits this oxidation. Incubation of sugars with vanadate, prior to addition of NADH, accelerates this oxidation of subsequently added NADH and eliminates the lag phase otherwise noted. Incubation of sugars with vanadate also results in the reduction of vanadate to vanadyl, with appearance of a blue-green color probably associated with a vanadyl-vanadate complex. It appears that sugars reduce vanadate to vanadyl which, in turn, reduces O2 to O2- and that vanadate plus O2- then catalyzes the oxidation of NAD(P)H by a free radical chain reaction. Such oxidation of NAD(P)H may account for several of the biological effects of vanadate.  相似文献   

7.
1. The NAD(P) transhydrogenase activity of the soluble fraction of sonicated rat liver mitochondrial preparations was greater than the NAD-linked isocitrate dehydrogenase activity, and the NAD-linked and NADP-linked isocitrate dehydrogenase activities were not additive. The NAD-linked isocitrate dehydrogenase activity was destroyed by an endogenous autolytic system or by added nucleotide pyrophosphatase, and was restored by a catalytic amount of NADP. 2. We concluded that the isocitrate dehydrogenase of rat liver mitochondria was exclusively NADP-specific, and that the oxoglutarate/isocitrate couple could therefore be used unequivocally as redox reactant for NADP in experiments designed to operate only the NAD(P) transhydrogenase (or loop 0) segment of the respiratory chain in intact mitochondria. 3. During oxidation of isocitrate by acetoacetate in intact, anaerobic, mitochondria via the rhein-sensitive, but rotenone- and arsenite-insensitive, NAD(P) transhydrogenase, measurements of the rates of carbonyl cyanide p-trifluoromethoxyphenylhydrazone-sensitive and carbonyl cyanide p-trifluoromethoxyphenylhydrazone-insensitive pH change in the presence of various oxoglutarate/isocitrate concentration ratios gave an -->H(+)/2e(-) quotient of 1.94+/-0.12 for outward proton translocation by the NAD(P) transhydrogenase. 4. Measurements with a K(+)-sensitive electrode confirmed that the electrogenicity of the NAD(P) transhydrogenase reaction corresponded to the translocation of one positive charge per acid equivalent. 5. Sluggish reversal of the NAD(P) transhydrogenase reaction resulted in a significant inward proton translocation. 6. The possibility that isocitrate might normally be oxidized via loop 0 at a redox potential of -450mV, or even more negative, is discussed, and implies that a P/O quotient of 4 for isocitrate oxidation might be expected.  相似文献   

8.
Electron transfer activities and steady state reduction levels of Fe-S centers of NADH-Q oxidoreductase were measured in mitochondria, submitochondrial particles (ETPH), and complex I after treatment with various reagents. p-Chloromercuribenzenesulfonate destroyed the signal from center N-4 (gx = 1.88) in ETPH but not in mitochondria, showing that N-4 is accessible only from the matrix side of the inner membrane. N-Bromosuccinimide also destroyed the signal from N-4 but without inhibiting rotenone-sensitive electron transfer to quinone, suggesting a branched pathway for electron transfer. Diethylpyrocarbonate caused oxidation of N-3 and N-4 in the steady state without changing N-1, suggesting N-1 is before N-3 and N-4. Difluorodinitrobenzene and dicyclohexylcarbodiimide inhibited oxidation of all Fe-S centers and tetranitromethane inhibited reduction of all Fe-S centers. Titrations of the rate of superoxide (O2-) generation in rotenone-treated submitochondrial particles were similar with the ratio [NADH]/[NAD] and that of 3-acetyl pyridine adenine nucleotide in spite of different midpoint potentials of the two couples. On reaction with inhibitors the inhibition of O2- formation was similar to that of ferricyanide reductase rather than quinone reductase. The rate of O2- formation during ATP-driven reverse electron transfer was 16% of the rate observed with NADH. The presence of NAD increased the rate to 83%. The results suggest that bound, reduced nucleotide, probably E-NAD., is the main source of O2- in NADH dehydrogenase. The effect of ATP on the reduction levels of Fe-S centers in well-coupled ETPH was measured by equilibrating with either NADH/NAD or succinate/fumarate redox couples. With NADH/NAD none of the Fe-S centers showed ATP induced changes, but with succinate/fumarate all centers showed ATP-driven reduction with or without NAD present. The effect on N-2 was smaller than that on N-1, N-3, and N-4. These observations indicate that the major coupling interaction is between N-2 and the low potential centers, N-1, N-3, and N-4. Possible schemes of coupling in this segment are discussed.  相似文献   

9.
Soleus and extensor digitorum longus (EDL) mitochondria and sarcotubular system were examined in sedentary and trained (treadmill for 12 wk) male rats that were treated with fluoxymesterone or methandrostanolone (2 mg/kg, 5 days/wk, for 8 wk). Neither physical exercise nor anabolic/androgenic steroid administration resulted in a significant change in muscle wet weight. Treatment with the anabolizing androgens increased succinate dehydrogenase activity in fast-twitch muscle mitochondria; this effect was not enhanced by training and was not observed in soleus mitochondria. On the other hand, the content of the slow-twitch muscle in sarcotubular fraction was increased in sedentary rats by fluoxymesterone or methandrostanolone treatment, whereas no significant changes were found in EDL. The training program affected adenosinetriphosphatase (ATPase) activities in the sarcotubular fraction; Mg2(+)-ATPase was increased in both soleus and EDL, but Ca2(+)-ATPase was decreased only in soleus. However, in sedentary animals only the Mg2(+)-dependent activity of EDL was increased by anabolizing androgen treatment, and this change was not potentiated by additional training. The present data indicate that anabolic/androgenic steroids can affect mitochondrial and sarcotubular enzymes in skeletal muscle. The effects are muscle-type specific.  相似文献   

10.
Prasad N  Topping RS  Zhou D  Decker SJ 《Biochemistry》2000,39(23):6929-6935
Phosphoinositide-dependent kinase (PDK1) regulates a number of pathways involved in responses to stress and in growth factor signaling; however, little is known concerning the mechanisms governing the activity of PDK1. In this report, we find that oxidative stress (H(2)O(2)) and vanadate induce tyrosine phosphorylation of PDK1. These effects of H(2)O(2) and vanadate were found in 293T cells and CH310T1/2 cells expressing exogenous PDK1 and in A20 lymphoma cells expressing endogenous PDK1. Exogenously expressed PDK1 was also tyrosine-phosphorylated in response to NGF treatment of 293T expressing TrkA. H(2)O(2) induced a more rapid tyrosine phosphorylation of PDK1 relative to vanadate, and only vanadate-induced tyrosine phosphorylation of PDK1 was sensitive to pretreatment of cells with wortmannin. In vitro, PDK1 could be tyrosine-phosphorylated by both the c-Src and Abl tyrosine kinases. Both H(2)O(2) and vanadate treatments increased the activity of PDK1 when the serum/glucocorticoid regulated kinase (SGK) was used as substrate. Vanadate treatment appeared to bypass the requirement for phosphatidylinositol 3,4,5-trisphosphate when Akt was used as substrate for PDK1. Tyrosine phosphorylation of PDK1 by the Abl tyrosine kinase also increased the activity of PDK1 toward SGK and Akt. These data suggest a novel mechanism through which PDK1 activity may be regulated.  相似文献   

11.
It has been reported that p53 acetylation, which promotes cellular senescence, can be regulated by the NAD(+)-dependent deacetylase SIRT1, the human homolog of yeast Sir2, a protein that modulates lifespan. To clarify the role of SIRT1 in cellular senescence induced by oxidative stress, we treated normal human diploid fibroblast TIG-3 cells with H(2)O(2) and examined DNA cleavage, depletion of intracellular NAD(+), expression of p21, SIRT1, and acetylated p53, cell cycle arrest, and senescence-associated beta-galactosidase (SA-beta-gal) activity. DNA cleavage was observed immediately in TIG-3 cells treated with H(2)O(2), though no cell death was observed. NAD(+) levels in TIG-3 cells treated with H(2)O(2) were also decreased significantly. Pre-incubation with the poly (ADP-ribose) polymerase (PARP) inhibitor resulted in preservation of intracellular NAD(+) levels. The amount of acetylated p53 was increased in TIG-3 cells at 4h after H(2)O(2) treatment, while there was little to no decrease in SIRT1 protein expression. The expression level of p21 was increased at 12h and continued to increase for up to 24h. Additionally, exposure of TIG-3 cells to H(2)O(2) induced cell cycle arrest at 24h and increased SA-beta-gal activity at 48h. This pathway likely plays an important role in the acceleration of cellular senescence by oxidative stress.  相似文献   

12.
The mitochondria-rich epithelial cells of the renal medullary thick ascending limb (mTAL) reabsorb nearly 25% of filtered sodium (Na(+)) and are a major source of cellular reactive oxygen species. Although we have shown that delivery of Na(+) to the mTAL of rats increases superoxide (O(2)(·-)) production in mTAL, little is known about H(2)O(2) production, given the lack of robust and selective fluorescent indicators for determining changes within the whole cell, specifically in the mitochondria. The present study determined the effect of increased tubular flow and Na(+) delivery to mTAL on the production of mitochondrial H(2)O(2) in mTAL. H(2)O(2) responses were determined in isolated, perfused mTAL of Sprague-Dawley rats using a novel mitochondrial selective fluorescent H(2)O(2) indicator, mitochondria peroxy yellow 1, and a novel, highly sensitive and stable cytosolic-localized H(2)O(2) indicator, peroxyfluor-6 acetoxymethyl ester. The results showed that mitochondrial H(2)O(2) and cellular fluorescent signals increased progressively over a period of 30 min following increased tubular perfusion (5-20 nl/min), reaching levels of statistical significance at ~10-12 min. Responses were inhibited with rotenone or antimycin A (inhibitors of the electron-transport chain), polyethylene glycol-catalase and by reducing Na(+) transport with furosemide or ouabain. Inhibition of membrane NADPH-oxidase with apocynin had no effect on mitochondrial H(2)O(2) production. Cytoplasmic H(2)O(2) (peroxyfluor-6 acetoxymethyl ester) increased in parallel with mitochondrial H(2)O(2) (mitochondria peroxy yellow 1) and was partially attenuated (~65%) by rotenone and completely inhibited by apocynin. The present data provide clear evidence that H(2)O(2) is produced in the mitochondria in response to increased flow and delivery of Na(+) to the mTAL, and that whole cell H(2)O(2) levels are triggered by the mitochondrial reactive oxygen species production. The mitochondrial production of H(2)O(2) may represent an important target for development of more effective antioxidant therapies.  相似文献   

13.
1. The activities of NMN adenylyltransferase and of NAD(+) kinase have been measured in the livers of adrenalectomized or alloxan-diabetic rats and in the livers of rats treated with glucagon, pituitary growth hormone or thyroxine. 2. The activities of these enzymes have been compared with the effects of the same treatments on the nicotinamide nucleotide concentrations in the liver. 3. Alloxandiabetes (+37%) and thyroxine (+27%) both increased the activity of NMN adenylyltransferase. The other treatments were without effect on this enzyme. 4. Only thyroxine increased the activity of NAD(+) kinase significantly (+26%) although both adrenalectomy and glucagon tended to increase its activity. 5. The activity of NAD(+) glycohydrolase was measured in the livers of diabetic rats, and in the livers of rats treated with either growth hormone or thyroxine. Of these treatments, only growth hormone altered the enzyme activity (+26%, calculated on a total hepatic activity basis). 6. Female rats had a greater hepatic NAD(+)-kinase activity than males but there was no sex difference with respect to NMN adenylyltransferase. 7. The lack of correlation between the maximum potential activity of these three enzymes and the known changes of the nicotinamide nucleotides in each of the hormone conditions is discussed.  相似文献   

14.
Endogenous fluorescence was used to measure the extent of reduction of mitochondrial NAD in individual, isolated rat cardiac myocytes. NAD reduction was determined from emitted fluorescence at 415 and 470 nm during brief epi-illumination at 365 nm. NAD reduction of resting myocytes, superfused with medium equilibrated with 95% O2/5% CO2, was 27 +/- 3% (SE) (n = 78), comparable to that in beating whole heart. Increasing intracellular Ca2+ did not significantly change NAD reduction. NAD reduction decreased reversibly to 11 +/- 1% (n = 78) in contracting myocytes electrically paced at 5 Hz for 10 min. Oxygen uptake was stimulated fivefold. There was minimal change in sarcoplasmic pH measured by fluorescence of carboxy-seminaphthorhodafluor-1. However, NAD reduction increased reversibly in response to electrically paced contractions when: (a) myoglobin was inactivated with sodium nitrite (37 +/- 7%; n = 48); or (b) cells were more densely layered and gassed with 20% O2/5% CO2 (48 +/- 3%; n = 30). We conclude that (a) the ratio NADH/NAD is decreased in well-oxygenated cells with increased work; (b) steady-state NAD reduction is increased with increased work when oxygen delivery is limited; and (c) functional myoglobin ensures an oxygen supply to the mitochondria of working cells.  相似文献   

15.
It was shown that the increase in the activities of transhydrogenase and NAD(+)-dependent isocitrate dehydrogenase after incubation of mitochondria with cAMP is due to the stimulating effect of cAMP on mitochondria, but not to the increased stability of mitochondria to the incubation procedure. Treatment of mitochondria with trypsin prevents the action of cAMP on the both enzymes. The integrity of the inner mitochondrial membrane is necessary for the manifestation of cAMP effect. Pretreatment of mitochondria with the local anesthetic, lidocaine, prevents the activation of NAD(P)(+)-transhydrogenase and NAD(+)-dependent isocitrate dehydrogenase during subsequent incubation of mitochondria with cAMP. It is concluded that the role of the inner mitochondrial membrane consists in the reception of the cAMP signal for the internal compartment of mitochondria, i.e. for mitoplasts. Peripheral protein(s) on the external side of the inner mitochondrial membrane seems to play a role in cAMP reception.  相似文献   

16.
Fusicoccin (FC) is a well known toxin acting as a 14-3-3 protein-mediated activator of the plasma membrane H(+)-ATPase and the biochemical and physiological changes induced in the cell by this toxin have, up to now, been ascribed to the increased rate of proton extrusion by this pump leading to external acidification and cell hyperpolarization. In a recent work (Malerba M et al. 2003, Physiologia Plantarum, 119: 480-488) it was shown that, besides the previously well studied changes, FC induces a large stimulation of H(2)O(2) production, an activation of alternative respiration and a leakage of cytochrome c from mitochondria. In this article further studies on the relation between the H(2)O(2) overproduction and medium acidification are reported. The increase in the rate of H(2)O(2) accumulation is particularly evident when high concentrations of the toxin ensure a rapid acidification of the medium, but it is not obtained when the time-course of acidification is reproduced by external acid additions. The FC-dependent H(2)O(2) overproduction is strongly inhibited by inhibitors of the H(+)-ATPase activity, such as vanadate and erythrosin B, and it does not occur when the activation of the H(+)-ATPase is prevented by phenylarsine oxide (PAO), an inhibitor of the activating interaction between the enzyme and its regulative 14-3-3 protein. Interestingly, all these inhibitors only partially prevent the leakage of cytochrome c from the mitochondria. A kinetic analysis of FC-dependent changes of 14-3-3s shows that the initial increase in the plasma membrane level of these proteins, presumably due to translocation of free cytosolic forms, is followed by a remarkable increase in the level of the 14-3-3 proteins located in the cytosol. This latter change is not prevented by inhibitors of the activity or activation of the H(+)-ATPase. These results suggest that, besides the H(+)-ATPase activation, FC can induce other cell changes possibly mediated by changes of the regulative 14-3-3 proteins.  相似文献   

17.
H2O2 and vanadate are known insulinomimetic agents. Together they induce insulin's bioeffects with a potency which exceeds that seen with insulin, vanadate, or H2O2 alone. Employing Western blotting with anti-P-Tyr antibodies, we have identified in Fao cells at least four proteins (pp180, 150, 114, and 100) whose P-Tyr content is rapidly increased upon treatment of the cells with 3 mM H2O2. Tyrosine phosphorylation of these and additional proteins was markedly potentiated (6-10-fold) when 100 microM sodium orthovanadate was added together with H2O2. The effects of H2O2 and vanadate on protein tyrosine phosphorylation were rapid and specific. The enhanced tyrosine phosphorylation was accompanied by a concomitant inhibition of a cytosolic protein tyrosine phosphatase activity. The latter was inhibited by 50% in 3 mM H2O2-treated cells. The inhibitory effect was augmented in the combined presence of H2O2 and vanadate. Half- and maximal effects of vanadate were obtained at 15 microM and 1 mM, respectively. Vanadate (1 mM) alone, added to the cells, had only a trivial effect on protein tyrosine phosphatase activity. A 45-s challenge with insulin (10(-7) M) of cells pretreated with H2O2 largely mimicked the potentiating effects of vanadate on protein tyrosine phosphorylation but not on protein tyrosine phosphatase activity. Our results suggest the involvement of multiple tyrosine-phosphorylation proteins in mediating the biological effects of H2O2/vanadate. Their enhanced phosphorylation can be attributed at least in part, to the inhibitory effects exerted by H2O2 alone, or in combination with vanadate, on protein tyrosine phosphatase activity. The similarity between proteins phosphorylated in Fao cells in response to H2O2/vanadate or H2O2/insulin, suggests that either treatment stimulates protein tyrosine kinases having similar substrate specificities. The insulin receptor kinase is a likely candidate as its activity is markedly enhanced either by insulin (plus H2O2) or by H2O2/vanadate.  相似文献   

18.
Regulation of mitochondrial functions in vivo by catecholamines was examined indirectly by depleting the catecholamines stores by reserpine treatments of the experimental animals. Reserpine treatment resulted in decreased respiratory activity in liver and brain mitochondria with the two NAD+-linked substrates: glutamate and pyruvate + malate with succinate ATP synthesis rate decreased in liver mitochondria only. With ascorbate + TMPD system, the ADP/O ratio and ADP phosphorylation rate decreased in brain mitochondria. For the heart mitochondria, state 3 respiration rates decreased for all substrates. In the liver mitochondria basal ATPase activity decreased by 51%, but in the presence of Mg2+ and/or DNP increased significantly. In the brain and heart mitochondria ATPase activities were unchanged. The energy of activation in high temperature range increased liver mitochondrial ATPase while in brain mitochondria reserpine treatment resulted in abolishment in phase transition. Total phospholipid (TPL) content of the brain mitochondria increased by 22%. For the heart mitochondria TPL content decreased by 19% and CHL content decreased by 34%. Tissue specific differential effects were observed for the mitochondrial phospholipid composition. Liver mitochondrial membranes were more fluidized in the reserpine-treated group. The epinephrine and norepinephrine contents in the adrenals decreased by 68 and 77% after reserpine treatment.  相似文献   

19.
Respiring mitochondria produce H(2)O(2) continuously. When production exceeds scavenging, H(2)O(2) emission occurs, endangering cell functions. The mitochondrial peroxidase peroxiredoxin-3 reduces H(2)O(2) to water using reducing equivalents from NADPH supplied by thioredoxin-2 (Trx2) and, ultimately, thioredoxin reductase-2 (TrxR2). Here, the contribution of this mitochondrial thioredoxin system to the control of H(2)O(2) emission was studied in isolated mitochondria and cardiomyocytes from mouse or guinea pig heart. Energization of mitochondria by the addition of glutamate/malate resulted in a 10-fold decrease in the ratio of oxidized to reduced Trx2. This shift in redox state was accompanied by an increase in NAD(P)H and was dependent on TrxR2 activity. Inhibition of TrxR2 in isolated mitochondria by auranofin resulted in increased H(2)O(2) emission, an effect that was seen under both forward and reverse electron transport. This effect was independent of changes in NAD(P)H or membrane potential. The effects of auranofin were reproduced in cardiomyocytes; superoxide and H(2)O(2) levels increased, but similarly, there was no effect on NAD(P)H or membrane potential. These data show that energization of mitochondria increases the antioxidant potential of the TrxR2/Trx2 system and that inhibition of TrxR2 results in increased H(2)O(2) emission through a mechanism that is independent of changes in other redox couples.  相似文献   

20.
The three enzymes required for the production and utilization of l-(+)-3-hydroxybutyrate were sought in various tissues of the rat. All tissues examined contained substantial amounts of (No. 1) l-(+)-3-hydroxybutyryl CoA dehydrogenase (EC 1.1.1.35). The specific activity of (No. 2) l-(+)-3-hydroxybutyryl CoA deacylase (EC 3.1.2) was highest in liver (3.8 mU/mg in mitochondrial matrix (1 U = 1 μmol/min). Brain, heart, and skeletal muscle contained < 20% of this activity. The chromatography of liver mitochondrial “matrix” preparations on DEAE-cellulose resolved the deacylase into two peaks. Peak I hydrolyzed 2- or 3- carbon acylCoA esters more efficiently than l-(+)-3-hydroxybutyrate CoA, while Peak II activity was highest using l-(+)-3-hydroxybutyryl CoA. The Km(app) for Peak II deacylase with l-(+)-3-hydroxybutyryl CoA was 19 μm. Acyl CoA synthetase (EC 6.2.1.2) (No. 3) was assayed with sorbate (sorboyl CoA ligase) or l-(+)-3-hydroxybutyrate (l-(+)-3-hydroxybutyryl CoA ligase). The highest specific activity for l-(+)-3-hydroxybutyryl CoA ligase was associated with brain mitochondria (8.3 mU/mg). In the “matrix” fraction of rat liver mitochondria the activities of these two acyl CoA synthetases were distinguished chromatographically and by their stability at various pH values. Heart and skeletal muscle mitochondria contained <10% of the liver activities of both ligases. These data implicate the liver as a site of l-(+)-3-hydroxybutyrate production.  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