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. Increasing concentrations of 3-(3,4-dichlorophenyl)-1,1-dimethylurea (DCMU), a mild respiratory-chain inhibitor [Halestrap (1987) Biochim. Biophys. Acta 927, 280-290], caused progressive inhibition of glucose production from lactate + pyruvate by hepatocytes from starved rats incubated in the presence or absence of oleate and gluconeogenic hormones. 2. No significant changes in tissue ATP content were observed, but there were concomitant decreases in ketone-body output and cytochrome c reduction and increases in NADH fluorescence and the ratios of [lactate]/[pyruvate] and [beta-hydroxybutyrate]/[acetoacetate]. 3. The inhibition by DCMU of palmitoylcarnitine oxidation by isolated liver mitochondria was used to calculate a flux control coefficient of the respiratory chain towards gluconeogenesis. In the presence of 1 mM-oleate, the calculated values were 0.61, 0.39 and 0.25 in the absence of hormone and in the presence of glucagon or phenylephrine respectively, consistent with activation of the respiratory chain in situ as previously suggested [Quinlan & Halestrap (1986) Biochem. J. 236, 789-800]. 4. Cytoplasmic oxaloacetate concentrations were shown to decrease under these conditions, implying inhibition of pyruvate carboxylase. 5. Inhibition of gluconeogenesis from fructose and dihydroxyacetone was also observed with DCMU and was accompanied by an increased output of lactate + pyruvate, suggesting that activation of pyruvate kinase was occurring. With the latter substrate, measurements of tissue ADP and ATP contents showed that DCMU caused a small fall in [ATP]/[ADP] ratio. 6. Two inhibitors of fatty acid oxidation, pent-4-enoate and 2-tetradecylglycidate, were shown to abolish and to decrease respectively the effects of hormones, but not valinomycin, on gluconeogenesis from lactate + pyruvate, without changing tissue ATP content. 7. It is concluded that the hormonal increase in mitochondrial matrix volume stimulates fatty acid oxidation and respiratory-chain activity, allowing stimulation of pyruvate carboxylation and thus gluconeogenesis to occur without major changes in [ATP]/[ADP] or [NADH]/[NAD+] ratios. 8. The high flux control coefficient of the respiratory chain towards gluconeogenesis may account for the hypoglycaemic effect of mild respiratory-chain inhibitors.  相似文献   

2.
Methods have been developed to measure the lysophospholipid content and matrix volume of liver cell mitochondria in situ in order to test the hypothesis that these parameters may be important in the hormonal control of mitochondrial function [Armston, Halestrap & Scott (1982) Biochim. Biophys. Acta 681, 429-439]. No change in the labelling of mitochondrial lysophospholipids with [32P]Pi was detected after treatment of liver cells with glucagon, phenylephrine or vasopressin. Incorporation of [32P]Pi into mitochondrial phosphatidylinositol was enhanced by phenylephrine and vasopressin. Mitochondrial volumes were measured using rapid disruption of cells by sonication into 3H2O and [14C]sucrose or without cell disruption using 3H2O and [14C]mannitol. In control cells the two methods gave values of 1.09 and 0.40 microliters/mg of mitochondrial protein respectively, which represent 19 and 7% respectively of the total cell volume measured with 3H2O and inulin [14C]carboxylic acid. Both methods showed that glucagon, phenylephrine and 1 nm-valinomycin produced significant increases (13% and 26% using sucrose and mannitol respectively) in mitochondrial volume. The increase was coincident with the stimulation of gluconeogenesis from L-lactate and pyruvate and of mitochondrial respiratory chain activity. The effects of glucagon and phenylephrine were additive on both mitochondrial volume and respiratory chain activity, but not on gluconeogenesis. Liver cells exposed to gluconeogenic hormones or low concentrations of valinomycin showed a decrease in light scattering at 520 nM correlating with the change in mitochondrial volume but without a change in whole-cell volume. The time course and hormone sensitivity of this response were similar to those for the hormonal stimulation of gluconeogenesis. The light-scattering response to glucagon, phenylephrine and vasopressin, but not to valinomycin, were greatly reduced or abolished in Ca2+-free media.  相似文献   

3.
Phenylephrine, vasopressin and glucagon each increased the amount of active (dephospho) pyruvate dehydrogenase (PDHa) in isolated rat hepatocytes. Treatment with 4 beta-phorbol 12-myristate 13-acetate (PMA) opposed the increase in PDHa caused by both phenylephrine and glucagon, but had no effect on the response to vasopressin: PMA alone had no effect on PDHa. As PMA is known to prevent the phenylephrine-induced increase in cytoplasmic free Ca2+ concentration ([Ca2+]c) and to diminish the increase [Ca2+]c caused by glucagon, while having no effect on the ability of vasopressin to increase [Ca2+]c, these data are consistent with the notion that in intact cells an increase in [Ca2+]c results in an increase in the mitochondrial free Ca2+ concentration, which in turn leads to the activation of PDH. In the presence of 2.5 mM-Ca2+, glucagon caused an increase in NAD(P)H fluorescence in hepatocytes. This increase is taken to reflect an enhanced activity of mitochondrial dehydrogenases. PMA alone had no effect on NAD(P)H fluorescence; it did, however, compromise the increase produced by glucagon. When the extracellular free [Ca2+] was decreased to 0.2 microM, glucagon could still increase NAD(P)H fluorescence. Vasopressin also increased fluorescence under these conditions; however, if vasopressin was added after glucagon, no further increase in fluorescence was observed. Treatment of the cells with PMA resulted in a smaller increase in NAD(P)H fluorescence on addition of glucagon: the subsequent addition of vasopressin now caused a further increase in fluorescence. Changes in [Ca2+]c corresponding to the changes in NAD(P)H fluorescence were observed, again supporting the idea that [Ca2+]c indirectly regulates intramitochondrial dehydrogenase activity in intact cells. PMA alone had no effect on pyruvate kinase activity, and the phorbol ester did not prevent the inactivation caused by glucagon. The latter emphasizes the different mechanisms by which the hormone influences mitochondrial and cytoplasmic metabolism.  相似文献   

4.
The ESR spectra of the free radicals formed by the autoxidation of serotonin, 5-hydroxyindole, and 5-hydroxytryptophan in 1 N NaOH are presented. The analysis of the hyperfine splitting constants in H2O and D2O characterize these free radicals as semiquinone-imines, the one-electron oxidation product of the corresponding indole. At alkaline pH, autoxidation of these compounds ultimately leads to solid precipitate and unresolved ESR spectra characteristic of polymeric material. The reduction of cytochrome c at pH 7.4 by a wide variety of indoles correlates with the amplitude of the ESR signal in 1 N NaOH, as do other processes thought to be related to 5-hydroxyindole free radical formation. Relative to the rate of cytochrome c reduction, neither serotonin nor the serotonin free radical appears to react with oxygen to form superoxide. In the presence of NAD(P)H, the serotonin radical most probably oxidizes NAD(P)H to form the NAD(P). radical. The NAD(P). radical then reacts with oxygen to form superoxide, which ultimately reduces cytochrome c.  相似文献   

5.
The interaction of heme nonapeptide (a proteolytic product of cytochrome c) with purified NADH:cytochrome b5 (EC 1.6.2.2) and NADPH:cytochrome P-450 (EC 1.6.2.4) reductases was investigated. In the presence of heme nonapeptide, NADH or NADPH were enzymatically oxidized to NAD+ and NADP+, respectively. NAD(P)H consumption was coupled to oxygen uptake in both enzyme reactions. In the presence of carbon monoxide the spectrum of a carboxyheme complex was observed during NAD(P)H oxidation, indicating the existence of a transient ferroheme peptide. NAD(P)H oxidation could be partially inhibited by cyanide, superoxide dismutase and catalase. Superoxide and peroxide ions (generated by enzymic xanthine oxidation) only oxidized NAD(P)H in the presence of heme nonapeptide. Oxidation of NAD(P)H was more rapid with O2- than O2-2. We suggest that a ferroheme-O2 and various heme-oxy radical complexes (mainly ferroheme-O-2 complex) play a crucial role in NAD(P)H oxidation.  相似文献   

6.
The sensitivity of glucose production from L-lactate by isolated liver cells from starved rats to inhibition by alpha-cyano-4-hydroxycinnamate was studied. A small percentage of the maximal rate of gluconeogenesis was insensitive to inhibition by alpha-cyano-4-hydroxycinnamate, and evidence is presented to show that this is due to pyruvate entry into the mitochondria as alanine. After subtraction of this rate, Dixon plots of the reciprocal of the rate of gluconeogenesis against inhibitor concentration were linear both in the absence and presence of glucagon, phenylephrine or valinomycin, each of which stimulated gluconeogenesis by 30-50%. Pyruvate kinase activity was decreased by glucagon, but not by phenylephrine or valinomycin. Inhibition of gluconeogenesis by quinolinate (inhibitor of phosphoenolpyruvate carboxykinase) or monochloroacetate (probably inhibiting pyruvate carboxylation) caused a significant deviation from linearity of the Dixon plot obtained with alpha-cyano-4-hydroxycinnamate. Amytal, however, inhibited gluconeogenesis without affecting the linearity of this plot. These data, coupled with a computer simulation study, suggest that pyruvate transport may control gluconeogenesis from L-lactate and that hormones may stimulate this process through an effect on the respiratory chain. An additional role for pyruvate kinase and pyruvate carboxylase is quite compatible with the data presented.  相似文献   

7.
Direct reaction of NAD(P)H with oxidants like singlet oxygen ((1)O(2)) has not yet been demonstrated in biological systems. We therefore chose different rhodamine derivatives (tetramethylrhodamine methyl ester, TMRM; 2',4',5',7'-tetrabromorhodamine 123 bromide; and rhodamine 123; Rho 123) to selectively generate singlet oxygen within the NAD(P)H-rich mitochondrial matrix of cultured hepatocytes. In a cell-free system, photoactivation of all of these dyes led to the formation of (1)O(2), which readily oxidized NAD(P)H to NAD(P)(+). In hepatocytes loaded with the various dyes only TMRM and Rho 123 proved suited to generating (1)O(2) within the mitochondrial matrix space. Photoactivation of the intracellular dyes (TMRM for 5-10 s, Rho 123 for 60 s) led to a significant (29.6 +/- 8.2 and 30.2 +/- 5.2%) and rapid decrease in mitochondrial NAD(P)H fluorescence followed by a slow increase. Prolonged photoactivation (> or =15 s) of TMRM-loaded cells resulted in even stronger NAD(P)H oxidation, the rapid onset of mitochondrial permeability transition, and apoptotic cell death. These results demonstrate that NAD(P)H is the primary target for (1)O(2) in hepatocyte mitochondria. Thus NAD(P)H may operate directly as an intracellular antioxidant, as long as it is regenerated. At cell-injurious concentrations of the oxidant, however, NAD(P)H depletion may be the event that triggers cell death.  相似文献   

8.
When O2 was injected into an anaerobic suspension of valinomycin-treated rat liver mitochondria inhibited with rotenone, antimycin, and myxothiazol, a small amount of O2 (0.23-0.33 ng-atom of O/mg of protein) was reduced extremely rapidly (within the 2 s time-resolution of the oxygen electrode). The subsequent steady-state rate of flow of electrons to oxygen was very low [less than 3 nequiv. X s-1 X (g of mitochondrial protein)-1]. In the presence of valinomycin there was a rapid ejection of protons synchronous with the rapid phase of O2 consumption corresponding to 0.38-0.61 nequiv. of H+ X (mg of mitochondrial protein)-1. When valinomycin was replaced by carbonyl cyanide p-trifluoromethoxyphenylhydrazone (FCCP) there was a rapid alkalification of the medium corresponding to 0.20-0.42 nequiv. of H+ X (mg of mitochondrial protein)-1. When 2 mM-Fe(CN)6(4-) was present to re-reduce endogenous cytochrome c, O2 consumption was still biphasic but the second phase of O2 consumption was very much more rapid [600 nequiv. X s-1 X (g of protein)-1], and resulted in the virtually complete consumption of the O2 in the pulse within 4 s. With 60 microM-Ru(NH3)6(2+) as reductant, O2 consumption was even faster [1200 nequiv. X s-1 X (g of protein)-1]. In a medium containing 150 mM-choline chloride with Ru(NH3)6(2+) as reductant, the proton per reducing equivalent stoichiometry (delta H+O/e-) was +0.95 in the presence of valinomycin and -0.94 in the presence of FCCP. In choline chloride medium containing Ru(NH3)6(2+) and valinomycin, there was an uptake of K+ ions corresponding to 1.86 K+/e-. It is concluded that nearly 1 proton is translocated outwards through cytochrome oxidase per oxidizing equivalent injected in this medium. In low ionic strength sucrose-based medium, with Ru(NH3)6(2+) as reductant, delta H+O/e- was 1.05 in the presence of valinomycin, and -0.71 in the presence of FCCP. It is concluded that the translocation of protons is accompanied by net acid production in this medium.  相似文献   

9.
1. ADH activity of Euglena grown with 50 mM ethanol decreased, but MEOS activity increased with a corresponding increase in the total amount of cytochrome P-450. 2. Phenobarbital treatment increased the total amount of cytochrome P-450. 3. CO and KCN, cytochrome P-450 ligands, diminished acetaldehyde formed from ethanol oxidation by MEOS. 4. The amounts of NAD(P)H cytochrome c reductases and cytochrome b5 type, components of microsomal monooxygenase reaction, have been spectrophotometrically measured. 5. NAD(P)H cytochrome c reductases activities were induced by phenobarbital. 6. DMSO, an inhibitor of rabbit MEOS, inhibited O2 consumption (11-20%) by Euglena grown with an ethanol, but not a lactate medium. 7. These studies indicate the presence of cytochrome P-450-dependent MEOS in Euglena similar to that in the mammalian hepatic cell.  相似文献   

10.
Previously, we demonstrated that mitochondrial NAD(P)H is the primary target of singlet oxygen (1O(2)) generated by photoactivation of mitochondria-selective rhodamine derivatives. Hence, local NAD(P)H oxidation/fluorescence decrease may be used to reveal the site of intracellular 1O(2) generation. Therefore, in addition to the previously used tetramethylrhodamine methylester (TMRM), 2('),4('),5('),7(')-tetrabromorhodamine 123 bromide (TBRB) and rhodamine 123 (Rho 123), we tested here whether mitochondrial NAD(P)H of cultured hepatocytes is directly oxidized upon irradiation of different "mitochondrial" photosensitizers (Photofrin; protoporphyrin IX; Al(III) phthalocyanine chloride tetrasulfonic acid; meso-tetra(4-sulfonatophenyl)porphine dihydrochloride; Visudyne). In contrast to TMRM and Rho 123, which directly oxidized NAD(P)H upon irradiation, irradiation of intracellular TBRB and the photochemical drugs only indirectly affected mitochondrial NAD(P)H due to loss of mitochondrial integrity. In line with this result only TMRM and Rho 123 exclusively localized within the mitochondrial matrix. Due to these results it is doubtful whether real mitochondrial photosensitizers actually exist among the photochemical drugs applicable/used for photodynamic therapy.  相似文献   

11.
A study is presented of the characteristics of redox-linked proton translocation in the b-c1 complex isolated from beef-heart mitochondria and reconstituted into phospholipid vesicles. Measurements of the H+/e- stoichiometry, with three different methods, show that four protons are released from the vesicles per 2e- flowing from quinols to cytochrome c, two of these protons formally deriving from scalar oxidation of quinols by cytochrome c. This H+/e- stoicheiometry is independent of the initial redox state of the b-c1 complex (fully reduced or oxidized) and the rate of electron flow through the complex. It does not change in the pH range 6.0 - 7.2, but declines to 1.5 going with pH from 7.2 - 8.3. This decrease is accompanied by enhancement of the rate of electron flow in the coupled state. Collapse of delta psi effected by valinomycin addition to turning-over b-c1 vesicles resulted in substantial oxidation of cytochrome b-566 and comparable reduction of cytochrome c1, with little oxidation of cytochrome b-562. Nigericin alone had no effect on the steady-state redox levels of b and c cytochromes. Its addition in the presence of valinomycin caused oxidation of b cytochromes but no change in the redox state of cytochrome c1. Valinomycin alone caused a marked enhancement of the rate of electron flow through the complex. Nigericin alone was ineffective, but caused further stimulation of electron flow when added in the presence of valinomycin. The data presented are discussed in terms of two mechanisms: the Q cycle and a model based on combination of protonmotive catalysis by special bound quinone and proton conduction along pathways in the apoproteins.  相似文献   

12.
A study of the steady-state kinetics of NADH(NADPH)-cytochrome c reductase (FMN-containing) from ale yeast (M. S. Johnson and S. A. Kuby (1985) J. Biol. Chem. 260, 12341-12350) has led to a postulated three-substrate random-ordered hybrid mechanism, where NAD(P)H and FMN add randomly and very likely in a steady-state fashion, followed by an ordered addition of cytochrome c. Kinetic parameters have been derived from this mechanism. Arrhenius plots showed large differences between NADH and NADPH, as the substrate-reductant. Menadione accelerated cytochrome c reduction and also O2 uptake, but vitamin K1 and coenzyme Q10 were ineffective as electron mediators, possibly as a result of their insolubility. With NADPH as the substrate-reductant, the order of the rate of reduction of electron acceptors was ferricyanide greater than DCIP greater than cytochrome c greater than oxygen; with menadione, the specificity sequence was cytochrome c greater than ferricyanide greater than DCIP greater than oxygen. With NADH, the order was ferricyanide greater than cytochrome c greater than oxygen greater than DCIP, which changed to cytochrome c greater than ferricyanide greater than oxygen greater than DCIP on addition of menadione. Cytochrome b5 was also reduced in the absence of oxygen. No transhydrogenase activity was observed, but the reduced thionicotinamide analogs of NADH and NADPH acted as substrates. Superoxide dismutase inhibited cytochrome c reduction in air by 50%, but O2-. was not necessary for cytochrome c reduction, as evidenced by the increase in rate in the absence of O2. The product of the reaction with oxygen appeared to be H2O2.  相似文献   

13.
Cytoplasmic Ca2+ signals are transferred to the mitochondria and activate the Krebs cycle. We have compared the efficiency of this process for two Ca2+ mobilising agonists, PGF2alpha and ATP (acting at metabotropic P2 receptors) in rat luteal cells. [Ca2+]c, [Ca2+]m and mitochondrial NAD(P)H were monitored by means of microspectrofluorimetry and confocal microscopy. While both agonists caused similar elevations of [Ca2+]c, changes in NAD(P)H were larger in response to PGF2alpha than to ATP. PGF2alpha more effectively increased NAD(P)H level also in mouse luteal cells. PGF2alpha caused a faster rate of rise of NAD(P)H fluorescence than ATP when reoxidation was prevented with rotenone, suggesting a faster rate of NAD(P)+ reduction. The NAD(P)H response to both agonists was dependent on the mobilisation of stored Ca2+. We found no difference in the efficacy of transmission of the [Ca2+]c signal to mitochondria in response to PGF2alpha and ATP. Raising [Ca2+]c with ionomycin increased the NAD(P)H signal, which was further raised by PGF2alpha but not by ATP. These data suggest that PGF2alpha potentiates the Ca2+-induced stimulation of mitochondrial metabolism by a Ca2+-independent mechanism and shows that agonists may modulate mitochondrial function differentially through a novel process beyond the simple transfer of Ca2+ from ER to mitochondria.  相似文献   

14.
It has been reported that vanadate-stimulated oxidation of NAD(P)H by microsomal systems can proceed anaerobically, in contrast to the general notion that the oxidation proceeds exclusively by an O(2-)-dependent free radical chain mechanism. The current study indicates that microsomal systems are endowed with a vanadate-reductase property, involving a NAD(P)H-dependent electron transport cytochrome P450 system. Our ESR measurements demonstrated the formation of a vanadium(IV) species in a mixture containing vanadate, rat liver microsomes, and NAD(P)H. This vanadium(IV) species was identified as the vanadyl ion (VO2+) by comparison with the ESR spectrum of VOSO4. The initial rate of vanadium(IV) formation depends linearly on the concentration of microsomes. The Michaelis-Menten constants were found to be: km = 1.25 mM and Vmax = 0.066 mumol (min)-1 (mg microsomes)-1, respectively. Pretreatment of the microsomes with carbon monoxide or K3Fe(CN)6 reduced vanadium(IV) generation, suggesting that the NAD(P)H-dependent electron transport cytochrome P450 system plays a significant role in the microsomal reduction of vanadate. Measurements under argon or in the presence of superoxide dismutase caused only minor (less than 10%) reductions in vanadium(IV) generation. The VO2+ species was also detected in NAD(P)H oxidation by fructose plus vanadate, a reaction known to proceed via an O(2-)-mediated chain mechanism. However, the amount of vanadium(IV) generated by this reaction was an order of magnitude smaller than that by the microsomal system and was inhibitable by superoxide dismutase, affirming the conclusion that the microsomal/NAD(P)H system is endowed with the (O(2-)-independent) vanadium(V) reductase property.  相似文献   

15.
1. A study is presented of the effects of pH, transmembrane pH gradient and electrical potential on oxidoreductions of b and c cytochromes in ox heart mitochondria and 'inside-out' submitochondrial particles. 2. Kinetic analysis shows that, in mitochondria at neutral pH, there is a restraint on the aerobic oxidation of cytochrome b566 with respect to cytochrome b562. Valinomycin plus K+ accelerates cytochrome b566 oxidation and retards net oxidation of cytochrome b562. At alkaline pH the rate of cytochrome b566 oxidation approaches that of cytochrome b562 and the effects of valinomycin on b cytochromes are impaired. 3. At slightly acidic pH, oxygenation of antimycin-supplemented mitochondria causes rapid reduction of cytochrome b566 and small delayed reduction of cytochrome b562. Valinomycin or a pH increase in the medium promote reduction of cytochrome b562 and decrease net reduction of cytochrome b566. 4. Addition of valinomycin to mitochondria and submitochondrial particles in the respiring steady state causes, at pH values around neutrality, preferential oxidation of cytochrome b566 with respect to cytochrome b562. The differential effect of valinomycin on oxidation of cytochromes b566 and b562 is enhanced by substitution of 1H2O of the medium with 2H2O and tends to disappear as the pH of the medium is raised to alkaline values. 5. Nigericin addition in the aerobic steady state causes, both in mitochondria and submitochondrial particles, preferential oxidation of cytochrome b562 with respect to cytochrome b566. This is accompanied by c cytochrome oxidation in mitochondria but c cytochrome reduction in submitochondrial particles. 6. In mitochondria as well as in submitochondrial particles, the aerobic transmembrane potential (delta psi) does not change by raising the pH of the external medium from neutrality to alkalinity. The transmembrane pH gradient (delta pH) on the other hand, decrease slightly. 7. The results presented provide evidence that the delta psi component of the aerobic delta microH+ (the sum of the proton chemical and electrical activities) exerts a pH-dependent constraint on forward electron flow from cytochrome b566 to cytochrome b562. This effect is explained as a consequence of anisotropic location of cytochromes b566 and b562 in the membrane and the pH-dependence of the redox function of these cytochromes. Transmembrane delta pH, on the other hand, exerts control on electron flow from cytochrome b562 to c cytochromes.  相似文献   

16.
The O-dealkylation of 7-alkoxyresorufins to the highly fluorescent compound, resorufin (7-hydroxyphenoxazone), provides a rapid, sensitive, and convenient assay of certain forms of liver microsomal cytochrome P450. The results of this study indicate that NADPH-cytochrome P450 reductase catalyzes the reduction of resorufin (and the 7-alkoxyresorufins) to a colorless, nonfluorescent compound(s). The reduction of resorufin by NADPH-cytochrome P450 reductase was supported by NADPH but not NADH, and was not inhibited by dicumarol, which established that the reaction was not catalyzed by contaminating DT-diaphorase (NAD[P]H-quinone oxidoreductase). In addition to the rate of reduction, the extent of reduction of resorufin was dependent on the concentration of NADPH-cytochrome P450 reductase. The maintenance of steady-state levels of reduced resorufin required the continuous oxidation of NADPH, during which molecular O2 was consumed. When NADPH was completely consumed, the spectroscopic and fluorescent properties of resorufin were fully restored. These results indicate that the reduction of resorufin by NADPH-cytochrome P450 reductase initiates a redox cycling reaction. Stoichiometric measurements revealed of 1:1:1 relationship between the amount of NADPH and O2 consumed and the amount of H2O2 formed (measured fluorometrically). The amount of O2 consumed during the redox cycling of resorufin decreased approximately 50% in the presence of catalase, whereas the rate of O2 consumption decreased in the presence of superoxide dismutase. These results suggest that, during the reoxidation of reduced resorufin, O2 is converted to H2O2 via superoxide anion. Experiments with acetylated cytochrome c further implicated superoxide anion as an intermediate in the reduction of O2 to H2O2. However, the ability of reduced resorufin to reduce acetylated cytochrome c directly (i.e., without first reducing O2 to superoxide anion) precluded quantitative measurements of superoxide anion formation. Superoxide dismutase, but not catalase, increased the steady-state level of reduced resorufin and considerably delayed its reoxidation. This indicates that superoxide anion is not only capable of reoxidizing reduced resorufin, but is considerably more effective than molecular O2 in this regard. Overall, these results suggest that NADPH-cytochrome P450 reductase catalyzes the one-electron reduction of resorufin (probably to the corresponding semiquinoneimine radical) which can either undergo a second, one-electron reduction (presumably to the corresponding dihydroquinoneimine) or a one-electron oxidation by reducing molecular O2 to superoxide anion.(ABSTRACT TRUNCATED AT 400 WORDS)  相似文献   

17.
Enzymatic reduction of physiological Fe(III) complexes of the "labile iron pool" has not been studied so far. By use of spectrophotometric assays based on the oxidation of NAD(P)H and formation of [Fe(II) (1,10-phenanthroline)3]2+ as well as by utilizing electron paramagnetic resonance spectrometry, it was demonstrated that the NAD(P)H-dependent flavoenzyme lipoyl dehydrogenase (diaphorase, EC 1.8.1.4) effectively catalyzes the one-electron reduction of Fe(III) complexes of citrate, ATP, and ADP at the expense of the co-enzymes NAD(P)H. Deactivated or inhibited lipoyl dehydrogenase did not reduce the Fe(III) complexes. Likewise, in the absence of NAD(P)H or in the presence of NAD(P)+, Fe(III) reduction could not be detected. The fact that reduction also occurred in the absence of molecular oxygen as well as in the presence of superoxide dismutase proved that the Fe(III) reduction was directly linked to the enzymatic activity of lipoyl dehydrogenase and not mediated by O2. Kinetic studies revealed different affinities of lipoyl dehydrogenase for the reduction of the low molecular weight Fe(III) complexes in the relative order Fe(III)-citrate > Fe(III)-ATP > Fe(III)-ADP (half-maximal velocities at 346-485 microm). These Fe(III) complexes were enzymatically reduced also by other flavoenzymes, namely glutathione reductase (EC 1.6.4.2), cytochrome c reductase (EC 1.6.99.3), and cytochrome P450 reductase (EC 1.6.2.4) with somewhat lower efficacy. The present data suggest a (patho)physiological role for lipoyl dehydrogenase and other flavoenzymes in intracellular iron metabolism.  相似文献   

18.
The antioxidant functions of cytochrome c   总被引:6,自引:0,他引:6  
Low (C(1/2) = 1.5 x 10(-7) M) concentrations of horse cytochrome c strongly inhibit H(2)O(2) production by rat heart mitochondria under conditions of reverse electron transfer from succinate to NAD(+). The effect is abolished by binding of cytochrome c with liposomes and is not prevented by SOD. Yeast cytochrome c is much less effective than the horse protein whereas acetylated horse cytochrome c is without effect. H(2)O(2) formation stimulated by antimycin A is resistant to added cytochrome c. In inside-out submitochondrial vesicles, H(2)O(2) production is suppressed by all three cytochrome c samples tested, but at higher concentrations (C(1/2) is about 5 x 10(-7) M). In vesicles, SOD abolishes the cytochrome c inhibition. We conclude that extramitochondrial cytochrome c is competent in down-regulation of the Complex I H(2)O(2) production linked to the reverse electron transfer. Such an effect is absent in the inside-out submitochondrial vesicles where another antioxidant cytochrome c function can be observed, i.e. the oxidation of O(2-*) to O(2). A possible role of cytochrome c in the antioxidant defence is discussed.  相似文献   

19.
The determinants of reduction of the dye MTT (3-[4,5dimethylthiazol-2-yl]-2,5-diphenyltetrazolium bromide) in rat hepatocytes have been investigated. NADH, NADPH, and succinate were substrates for MTT reduction in rat liver homogenate, activity being greatest with NADH and least with succinate. Similar results were obtained with submitochondrial particles isolated from rat liver. NAD(P)Hdependent reduction of MTT was also detected in rat liver microsomes and cytosol. Rotenone, at a concentration that inhibited NAD(P)H-dependent MTT reduction in submitochondrial particles, did not inhibit MTT reduction in rat hepatocytes. Malonate, at a concentration that inhibited succinate-dependent MTT reduction in liver homogenate, did not inhibit MTT reduction in rat hepatocytes. Incubation of rat hepatocytes with ethanol or lactate (increase NADH levels), dicoumarol (inhibitor of DT-diaphorase), aminopyrine or hexobarbitone (substrates for the NADPH-requiring cytochrome P450-dependent microsomal monooxygenase) led to significant increases in the level of cellular MTT reduction. From these data, it is concluded that extramitochondrial NAD(P)H is the principal reductant for MTT reduction in rat hepatocytes, with mitochondrial dehydrogenase activity being only a minor contributor. It is also possible that cellular generation of superoxide (as might be expected on redox cycling of endogenous quinones following inhibition of DT diaphorase by dicoumarol) may be another source of MTT reduction. Caution should be exercised in ascribing an alteration in the level of cellular MTT reduction to a change in mitochondrial performance in the absence of corroborating evidence.  相似文献   

20.
We have recently reported that with a linear sucrose density gradient centrifugation two distinct types of membrane fragments, designated as X- and Y-fragments are obtained (Huang, C. H., Keyhani, E. and Lee, C. P. (1973) Biochim. Biophys. Acta 305, 455-473). Further characterization of these two membranes fragments is reported. (1) Potassium chloride at the concentration of 0.15 m extracts 7% and 30% of cytochrome c from the X- and Y-fragments, respectively. (2) When cytochrome c was added to the mitochondrial suspension prior to sonication, the cytochrome c content was increased by 6-8-fold in both X- and Y-fragments. Subsequently KC1 extraction resulted in loss of cytochrome c by 1/4 in the X- and by 2/3 in the Y-fragments. (3) With partially inhibitory concentrations of KCN, cytochrome c in either the X- or the KC1 extracted X-fragments showed uncoupler-sensitive, biphasic reduction kinetics upon the addition of NADH to the oligomycin-supplemented system. Under identical conditions rapid first order reduction kinetics were seen for cytochrome c in Y-fragments supplemented with either oligomycin or oligomycin + carbonyl cyanide p-trifluoromethoxyphenylhydrazone (FCCP). (4) When cytochrome c was added to the mitochondrial suspension after sonication, a significant amount of cytochrome c was bound to both X- and Y-fragments, but was readily removed with a high ionic strength medium. (5) Lubrol had little effect on the ATPase activity of the X- and the Y-fragments, suggesting a lack of membrane-buried ATPase. (6) Partial depletion of ATPase in X-fragments did not induce an increase in reactivity towards externally added cytochrome c. (7) Both the X- and the Y-fragments showed an energy-linked fluorescence enhancement of 8-anilinonaphthalene-1-sulfonate and an energy-linked fluorescence decrease of quinacrine. (8) In the presence of K-+ nigericin alone or in combination with valinomycin exhibited a stimulating effect on the rate of NADH oxidase of the oligomycinsupplemented X- and Y-fragments.  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