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.
Menadione in the presence of oxyhemoglobin will accelerate the formation of methomoglobin and result in the generation of superoxide anion. Menadione appears to oxidize slowly ferrohemoglobin to ferrihemoglobin, while forming menadione semiquinone in the process. Menadione semiquinone is known to react with molecular oxygen to yield superoxide anion. The superoxide anion appears to be the source of hydrogen peroxide which accounts for most of the observed methemoglobin formation when hemoglobin is reacted with menadione.  相似文献   

2.
1. Menadione was found to react with both the haem groups and the beta-93 thiol groups of haemoglobin. 2. It oxidized the haem groups of oxyhaemoglobin, giving mainly methaemoglobin and a smaller amount of haemichrome. The reaction rate was decrease in the presence of catalase and markedly accelerated in the presence of superoxide dismutase. It is proposed that the overall reaction involves the initial reversible formation of methaemoglobin and the semiquinone, and that the effect of superoxide dismutase is to prevent the reverse reaction, by removing superoxide and hene O2-. E.s.r. evidence for the information of the semi-quinone and its reactions is presented. 3. The reaction of menadione with the beta-93 thiol groups of haemoglobin appeared to be similar to that with other thiols, forming the 3-thioether derivative of menadione, but it was also accompanied by reduction of methaemoglobin. This reduction was prevented by superoxide dismutase, but appeared to be caused by the semiquinone radical, which was produced as an intermediate. 4. Reduced glutathione functioned only to a limited extent as a scavenger of the menadione semiquinone. Its main reaction was directly with menadione to form the thioether. Ascorbate was a more efficient scavenger, and accelerated the oxidation of oxyhaemoglobin by menadione. 5. The significance of these findings in relation to menadione-induced erythrocyte haemolysis is discussed.  相似文献   

3.
Quinones are believed to be toxic by a mechanism involving redox cycling and oxidative stress. In this study, we have used 2,3-dimethoxy-1,4-naphthoquinone (2,3-diOMe-1,4-NQ), which redox cycles to the same degree as menadione, but does not react with free thiol groups, to distinguish between the importance of redox cycling and arylation of free thiol groups in the causation of toxicity to isolated hepatocytes. Menadione was significantly more toxic to isolated hepatocytes than 2,3-diOMe-1,4-NQ. Both menadione and 2,3-diOMe-1,4-NQ caused an extensive GSH depletion accompanied by GSSG formation, preceding loss of viability. Both compounds stimulated a similar increase in oxygen uptake in isolated hepatocytes and NADPH oxidation in microsomes suggesting they both redox cycle to similar extents. Further evidence for the redox cycling in intact hepatocytes was the detection of the semiquinone anion radicals with electron spin resonance spectroscopy. In addition we have, using the spin trap DMPO (5,5-dimethyl-1-pyrroline N-oxide), demonstrated for the first time the formation of superoxide anion radicals by intact hepatocytes. These radicals result from oxidation of the semiquinone by oxygen and further prove that both these quinones redox cycle in intact hepatocytes. We conclude that while oxidative processes may cause toxicity, the arylation of intracellular thiols or nucleophiles also contributes significantly to the cytotoxicity of compounds such as menadione.  相似文献   

4.
Our previous studies showed that menadione causes endothelial dysfunction which results in decreased relaxation and increased contraction of blood vessels. This investigation examined the role of two possible mechanisms (oxidative stress and arylation) in menadione-induced endothelial dysfunction. Menadione increased superoxide anion generation in aortic rings in a dose-dependent manner. Superoxide dismutase (SOD), reversed the inhibitory effects of menadione on vascular relaxation. The relaxation induced by the NO donor, sodium nitroprusside, was inhibited by menadione pretreatment in a dose-dependent manner. Endothelial nitric oxide synthase activity (eNOS) was suppressed by menadione. Menadione resulted in a dose-dependent reduction of cGMP levels accumulated by acetylcholine. This reduction of cGMP levels was blocked by SOD treatment, suggesting that superoxide anion generated by menadione could play a role in the inhibition of the nitric oxide pathway. Evidence supporting a possible role for arylation in impaired vascular relaxation was suggested by the observation that benzoquinone, which does not induce oxidative stress in aortic rings, inhibited acetylcholine-induced vascular relaxation to the same extent as menadione. Collectively, these results suggest that menadione can cause endothelial dysfunction in blood vessels by the inhibition of the nitric oxide pathway via superoxide anion generation and that arylation activity may also be another important mechanism.  相似文献   

5.
Menadione is selectively toxic to erythrocytes. Although GSH is considered a primary target of menadione, intraerythrocyte thiolic alterations consequent to menadione exposure are only partially known. In this study alterations of GSH and protein thiols (PSH) and their relationship with methemoglobin formation were investigated in human and rat red blood cells (RBC) exposed to menadione. In both erythrocyte types, menadione caused a marked increase in methemoglobin associated with GSH depletion and increased oxygen consumption. However, in human RBC, GSH formed a conjugate with menadione, whereas, in rat RBC it was converted to GSSG, concomitantly with a loss of protein thiols (corresponding to menadione arylation), and an increase in glutathione-protein mixed disulfides (GS-SP). Such differences were related to the presence of highly reactive cysteines, which characterize rat hemoglobin (cys beta125). In spite of the greater thiol oxidation in rat than in human RBC, methemoglobin formation and the rate of oxygen consumption elicited by menadione in both species were rather similar. Moreover, in repeated experiments under N2 or CO-blocked heme, it was found that menadione conjugation (arylation) in both species was not dependent on the presence of oxygen or the status of heme. Therefore, we assumed that GSH (human RBC) and protein (rat RBC) arylation was equally responsible for increased oxygen consumption and Hb oxidation. Moreover, thiol oxidation of rat RBC was strictly related to methemoglobin formation.  相似文献   

6.
Youn H  Kang SO 《FEBS letters》2000,472(1):57-61
Lipoamide dehydrogenase from Streptomyces seoulensis could facilitate menadione-mediated cytochrome c reduction, which was mostly inhibited by superoxide dismutase, indicating the obvious involvement of superoxide radical anion. In this reaction, the production of superoxide radical anion occurred via a menadione semiquinone radical anion. When exposed to menadione, lipoamide dehydrogenase-overexpressing cells showed a much lower survival rate with a concomitant decrease of intracellular protein thiol than the wild-type strain. These results suggest that lipoamide dehydrogenase is a facilitating agent in the redox cycling of quinone compounds in vivo as well as in vitro and could inevitably increase the potential toxicity of the compounds.  相似文献   

7.
Macrophage cell cultures were treated with menadione, zymosan, or phorbol myristate acetate (PMA), and changes in productions of superoxide anion and hydroperoxide, and in glutathione oxidation and S-thiolation of cystatin-beta (formation of a mixed disulfide of cystatin-beta and glutathione) were examined. All three compounds promoted production of superoxide anion and hydroperoxide, but only menadione caused extensive oxidation of glutathione. Menadione caused S-thiolation of cystatin-beta in a dose-dependent fashion, but the other two compounds did not. Removal of menadione promptly reduced the oxidation of glutathione and S-thiolation of cystatin-beta induced by menadione. Inhibition of catalase by aminotriazol caused slight increase in the GSSG content in both menadione- and zymosan-treated cells, but not in S-thiolation of cystatin-beta in zymosan-treated cells. None of the three compounds influenced appreciably the activity of glutathione peroxidase, glutathione reductase, or superoxide dismutase in cultured cells. These results indicate that S-thiolation of cystatin-beta occurs in cells in response to oxidative challenge by menadione but not by zymosan or by the tumor promoter PMA. Dethiolation of cystatin-beta by purified thiol transferase and protein disulfide isomerase in the presence of different concentrations of GSH was examined in vitro. Both enzymes catalyzed dethiolation of cystatin-beta at a much lower level of GSH than that required for the non-enzymatic reaction, suggesting the importance of enzymatic catalysis of S-thiolation and dethiolation of cystatin-beta in cells.  相似文献   

8.
Univalent oxidation-reduction reactions coupled with the menadione (MK)/menadione semiquinone (MK' -) system were investigated by using microsomal Ravin enzymes. NADPH-cytochrome P-450 reduc-tase gave a dynamic equilibrium of oxidation-reduction of cytochrome b, in the presence of menadione (MK), the level of which depended on the concentration of O2 and superoxide dismutase. The data suggest that the superoxide and menadione radicals are involved as an active intermediate in this system. The overall reaction at steady state appears to be composed of four main reactions, eqs. 2-5. and eqs. 2 and 4 are in equilibrium  相似文献   

9.
Menadione is a synthetic derivative of the natural vitamins K with antiinflammatory activity among its potentially significant clinical properties. We have found this agent to stimulate the production of superoxide anion (O2-) in human polymorphonuclear leukocytes (PMN) and dimethylsulfoxide-differentiated HL-60 cells in a time-, cell number-, and drug concentration-dependent manner. Conversely, menadione attenuates both O2- production and lysozyme release in cells stimulated by phorbol myristate acetate (PMA), fMet-Leu-Phe, or Ca2+ ionophore. 4-Acetamido-4'-isothiocyano-2-2'-disulfonic acid stilbene and 4,4'-diisothiocyano-2-2'disulfonic acid stilbene, agents which inhibit transmembrane O2-) flux, do not alter menadione's effects on superoxide dismutase (SOD) inhibitable cytochrome c reduction in resting or PMA-stimulated PMN. Likewise, quinone reductase inhibitors, warfarin and dicumarol, known to attenuate vitamin K-dependent responses and enhance quinone-mediated oxidative stress, have no effect upon menadione-stimulated O2- production. Furthermore, menadione-induced suppression of stimulus-mediated lysozyme release is not reversed by cotreatment with oxygen metabolite scavenging enzymes SOD and catalase. Nevertheless, under conditions of restricted oxygen supply, the suppressive effect of menadione on stimulant-induced lysozyme release is greatly diminished. Thus, although pharmacological manipulation suggests otherwise, there appears to exist at least a component of the inhibitory activity of menadione that is oxygen dependent, and may be oxidative stress-related.  相似文献   

10.
Wheat chloroplasts photochemically reduced molecular oxygen, as a Hill oxidant in the Mehler reaction, to superoxide anion which then oxidized added 1,2-dihydroxybenzene-3,5-disulfonate to its semiquinone, a comparatively stable free radical at pH 7. The last mentioned reaction was rapid in aqueous solution, but the rate of formation of 1,2-dihydroxybenzene-3,5-disulfonate semiquinone by the chloroplast system was calculated as T1 of 0.6 s. The Mehler reaction, or more specifically the univalent reduction of oxygen by Photosystem I, was rate-limiting so that the 1,2-dihydroxybenzene-3,5-disulfonate seniquinone was a useful spin probe for superoxide anion production at room temperature. The ESR signal of 1,2-dihydroxybenzene-3,5-disulfonate semiquinone was proportional to its steady state concentration and decayed in the dark with a T1/2 of 5-6 s. This oxygen-dependent signal was enhanced by mediation of chloroplastic oxygen reduction through methyl viologen. The superoxide anion scavengers ascorbate and L-epinephrine competitively obscured 1,2-dihydroxybenzene-3,5-disulfonate semiquinone formation, butadded superoxide dismutase was not as effective in this role. Partial inhibition by superoxide dismutase was achieved only by preincubation of Photosystem I enriched particles with ten times the endogenous concentration of superoxide dismutase. This and the persistence of a small amount of a 1,2-dihydroxybenzene-3,5-disulfonate (Tiron) oxidizing species in the dark supports the concept of Tiron accessibility but not the superoxide dismutase accessibility of superoxide anion bound in its formative enzyme complex. Benzoquinone and naphthoquinone disulfonate also reacted with superoxide anion, and supported both the Hill reaction and the Mehler reaction as final oxidants of both water and superoxide anion.  相似文献   

11.
The formation of methemoglobin from oxyhemoglobin in a solution containing photoreduced riboflavin and oxygen was inhibited by superoxide dismutase. The rate of the reaction was pH-dependent in the range of 6.8 to 7.8, increasing as the pH was reduced. Inhibition by superoxide dismutase was enhanced as the EDTA concentration increased and was dependent on enzymatic activity. Under conditions in which superoxide dismutase inhibition was incomplete, catalase inhibited the reaction but mannitol had no effect. The data support the mediation of methemoglobin formation by superoxide. The hypothesis is offered that superoxide anion reduced the heme-bound oxygen in oxygemoglobin by one electron, permitting the subsequent dissociation of ferrihemoglobin and peroxide. The ability of superoxide dismutase to inhibit the formation of methemoglobin may represent one of its functions in the mature erythrocyte.  相似文献   

12.
R.W. Miller  F.D.H. Macdowall 《BBA》1975,387(1):176-187
Wheat chloroplasts photochemically reduced molecular oxygen, as a Hill oxidant in the Mehler reaction, to superoxide anion which then oxidized added 1,2-dihydroxybenzene-3,5-disulfonate to its semiquinone, a comparatively stable free radical at pH 7. The last mentioned reaction was rapid in aqueous solution, but the rate of formation of 1,2-dihydroxybenzene-3,5-disulfonate semiquinone by the chloroplast system was calculated as a T12 of 0.6 s. The Mehler reaction, or more specifically the univalent reduction of oxygen by Photosystem I, was rate-limiting so that the 1,2-dihydroxybenzene-3,5-disulfonate semiquinone was a useful spin probe for superoxide anion production at room temperature. The ESR signal of 1,2-dihydroxybenzene-3,5-disulfonate semiquinone was proportional to its steady state concentration and decayed in the dark with a T12 of 5–6 s. This oxygen-dependent signal was enhanced by mediation of chloroplastic oxygen reduction through methyl viologen. The superoxide anion scavengers ascorbate and l-epinephrine competitively obscured 1,2-dihydroxybenzene-3,5-disulfonate semiquinone formation, but added superoxide dismutase was not as effective in this role. Partial inhibition by superoxide dismutase was achieved only by preincubation of Photosystem I enriched particles with ten times the endogenous concentration of superoxide dismutase. This and the persistence of a small amount of a 1,2-dihydroxybenzene-3,5-disulfonate (Tiron) oxidizing species in the dark supports the concept of Tiron accessibility but not the superoxide dismutase accessibility of superoxide anion bound in its formative enzyme complex. Benzoquinone and naphthoquinone disulfonate also reacted with superoxide anion, and supported both the Hill reaction and the Mehler reaction as final oxidants of both water and superoxide anion.  相似文献   

13.
Human neutrophils incubated with phorbol myristate acetate oxidized hemoglobin within the intact erythrocyte by a mechanism dependent on cell-cell contact but independent of phagocytosis. Spectrophotometric examination of the erythrocyte lysates revealed that the major component formed was methemoglobin along with small amounts of a species with spectral characteristics similar to choleglobin. Methemoglobin formation was directly related to the neutrophil concentration and the time of incubation. The addition of superoxide dismutase or catalase modestly inhibited the formation of methemoglobin, while a combination of the enzymes provided the most dramatic protection. Methemoglobin of hydroxyl radical or hypochlorous acid scavengers. Apparently, either O2.- or H2O2 alone was capable of mediating methemoglobin formation in the intact erythrocyte. Maintenance of the intraerythrocytic hemoglobin in its oxygenated state or its derivatization to carbon monoxyhemoglobin markedly inhibited methemoglobin formation. Blockade of the anion channels in the intact erythrocyte with sulfonated stilbenes inhibited O2.- but not H2O2 from oxidizing intracellular hemoglobin. It appears that neutrophil-derived O2.- and H2O2 can cross the erythrocyte membrane through the anion channel or diffuse directly into the intracellular space and react with oxyhemoglobin or deoxyhemoglobin to form a mixture of hemoglobin oxidation products within the intact cell.  相似文献   

14.
Menadione (VK3), a quinone that undergoes redox cycles leading to the formation of superoxide radicals, was found to induce cell death in suspension culture of carrot cells. The effect of menadione was in a dose-dependent manner. 100-800 mumol/L menadione caused 10-33 percent cell death. When concentration of menadione reached 1 mmol/L, 100 percent of cell death was observed. DNA cleavage, a hallmark of apoptosis was further studied. DNA ladders were observed in cells treated with 600 and 800 mumol/L menadione but not with lower concentration treatments where only very low percentage of cell death was found. There was no DNA ladders in the cells treated with 1 mmol/L menadion indicating that necrosis may occur. In situ detection of nuclear DNA fragmentation by TUNEL reaction revealed fragmented nuclear DNA in cells treated with 100-800 mumol/L menadion but not in cells treated with 1 mmol/L menadione.  相似文献   

15.
Menadione promoted the production of active oxygen species (AOS) in both yeast cell suspension and the crude enzymes from the cells, but menadione sodium bisulfite (MSB) had little effect on the production of AOS in the cell suspension. MSB kept the stable increase in the electron transfer from intact yeast cells to anode compared to menadione, but the electron transfer promoted by MSB was inhibited in permeabilized yeast cell suspension. Menadione promoted oxidation of NAD(P)H much faster than MSB in permeabilized yeast cell suspension, suggesting the oxidative stress due to consumption of NAD(P)H. The proliferation of yeast cells was inhibited by menadione under aerobic conditions rather than anaerobic conditions, and the inhibitory effect was reduced by superoxide dismutase and catalase. The effect of MSB on the proliferation was much smaller than that of menadione. The above facts suggest that harmless MSB promotes the electron transfer from plasma membrane of yeast cells to anode. On the other hand, harmful menadione might promote the electron transfer from cytosol and plasma membrane to anode and dissolved oxygen.  相似文献   

16.
The prevention of injury from reactive oxygen species is critical for cellular resistance to many death stimuli. Resistance to death from the superoxide generator menadione in the hepatocyte cell line RALA255-10G is dependent on down-regulation of the c-Jun N-terminal kinase (JNK)/AP-1 signaling pathway by extracellular signal-regulated kinase 1/2 (ERK1/2). Because protein kinase C (PKC) regulates both oxidant stress and JNK signaling, the ability of PKC to modulate hepatocyte death from menadione through effects on AP-1 was examined. PKC inhibition with Ro-31-8425 or bisindolylmaleimide I sensitized this cell line to death from menadione. Menadione treatment led to activation of PKCmicro, or protein kinase D (PKD), but not PKCalpha/beta, PKCzeta/lambda, or PKCdelta/. Menadione induced phosphorylation of PKD at Ser-744/748, but not Ser-916, and translocation of PKD to the nucleus. PKC inhibition blocked menadione-induced phosphorylation of PKD, and expression of a constitutively active PKD prevented death from Ro-31-8425/menadione. PKC inhibition led to a sustained overactivation of JNK and c-Jun in response to menadione as determined by in vitro kinase assay and immunoblotting for the phosphorylated forms of both proteins. Cell death from PKC inhibition and menadione treatment resulted from c-Jun activation, since death was blocked by adenoviral expression of the c-Jun dominant negative TAM67. PKC and ERK1/2 independently down-regulated JNK/c-Jun, since inhibition of either kinase failed to affect activation of the other kinase, and simultaneous inhibition of both pathways caused additive JNK/c-Jun activation and cell death. Resistance to death from superoxide therefore requires both PKC/PKD and ERK1/2 activation in order to down-regulate proapoptotic JNK/c-Jun signaling.  相似文献   

17.
Superoxide-mediated release of iron from ferritin by some flavoenzymes   总被引:1,自引:0,他引:1  
NADH-lipoamide dehydrogenase mobilized iron from ferritin under aerobic conditions. Superoxide dismutase strongly inhibited this mobilization, indicating that the superoxide radical is generated by the enzymatic reaction and release iron from ferritin. Addition of lipoamide as an electron acceptor to NADH-lipoamide dehydrogenase increased the release of iron from ferritin and this release was partially inhibited by superoxide dismutase. Similarly, addition of menadione (2-methyl-1, 4-naphthoquinone) as an electron acceptor to xanthine-xanthine oxidase promoted the release of iron from ferritin and this release was strongly inhibited by superoxide dismutase. These results suggest that dihydrolipoamide and semiquinone of menadione can react with oxygen to form the superoxide radical that mediates release of iron from ferritin.  相似文献   

18.
Menadione (2-methyl-1,4-naphthoquinone) is a remarkably potent inhibitor of microsomal lipid peroxidation, effective at submicromolar concentrations. Its possible mechanism of action and the relationship between naphthoquinone structure and antioxidant activity were the topics of this investigation. In the microsomal lipid-peroxidizing system dependent on NADPH and ferric pyrophosphate, menadione, at concentrations of 50 microM or higher virtually eliminated the accumulation of malondialdehyde and lipid hydroperoxides. In the NADPH-independent, cumene hydroperoxide-dependent system, menadione was also an effective antioxidant, but only in the presence of reducing equivalents. These and other observations indicate that a reduced form of menadione, either the hydroquinone or semiquinone, is the active antioxidant, and suggest that it may trap hydroperoxy radicals, alkoxy radicals, or other free radicals involved in propagating lipid peroxidation. Moreover, these results show that electron diversion per se cannot account for the antioxidant effects of menadione. A comparison of the antioxidant activities of eight 1,4-naphthoquinones indicated that methyl substitution of C-2, lack of steric hindrance at C-3 or C-5, and (in the case of weak acids) a relatively high pKa are favorable structural features associated with strong antioxidant activity.  相似文献   

19.
The mechanism by which superoxide anion is generated by the interaction of phenylhydrazine with either oxy- or methemoglobin was investigated. Rather than superoxide anion generation resulting from an accelerated autooxidation of oxyhemoglobin, it was found that both oxy- and methemoglobin function as peroxidases toward phenylhydrazine with the resultant oxidation of this compound to phenyldiazine. Generation of phenyldiazine from the oxidation of phenylhydrazine by hemoglobin or by the hydrolysis and subsequent decarboxylation of methyl phenylazoformate (C6H5N=NCOOCH3) resulted in the production of superoxide anion. It is suggested that under certain conditions hemoglobin may function as a drug-metabolizing peroxidase.  相似文献   

20.
Interaction of menadione (2-methyl-1,4-naphthoquinone) with glutathione   总被引:2,自引:0,他引:2  
The interaction of menadione with reduced glutathione (GSH) led to a removal of menadione and formation of menadione-GSH conjugate and glutathione disulfide (GSSG). The changes in thiol level were essentially biphasic with an initial rapid decrease in GSH and appearance of GSSG (less than 1 min) followed by secondary less pronounced changes. The interaction of menadione and GSH caused an oxygen uptake and both superoxide anion radical and hydrogen peroxide were produced during the reaction, the amount dependent on the GSH/menadione ratio. Catalase did not protect against the initial decrease in GSH level but markedly inhibited the secondary changes while superoxide dismutase had little effect. These results suggest that the initial changes in thiol level are the result in part of a redox reaction between menadione and GSH as well as conjugate formation, whilst the secondary changes reflect conjugate formation and the activity of other oxidants such as hydrogen peroxide. The potential biological significance of this reaction was investigated using hepatocytes depleted of reduced pyridine nucleotides and thus not able to perform enzyme-catalyzed reduction of menadione. In these cells menadione induced GSSG formation at a rate similar to that observed in control cells. This suggests that quinone-induced oxidative challenge caused by the chemical interactions of a quinone and glutathione may have biological relevance.  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