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 (2-methyl-1,4-naphthoquinone) was used as a model compound to test the hypothesis that thioether conjugates of quinones can be toxic to tissues associated with their elimination through a mechanism involving oxidative stress. Unlike menadione, the glutathione (2-methyl-3-(glutathion-S-yl)-1,4-naphthoquinone; MGNQ) and N-acetyl-L-cysteine (2-methyl-3-(N-acetylcysteine-S-yl)-1,4-naphthoquinone; M(NAC)NQ) thioether conjugates were not able to arylate protein thiols but were still able to redox cycle with cytochrome c reductase/NADH and rat kidney microsomes and mitochondria. Interestingly, menadione and M(NAC)NQ were equally toxic to isolated rat renal epithelial cells (IREC) while MGNQ was nontoxic. The toxicity of both menadione and M(NAC)NQ was preceded by a rapid depletion of soluble thiols and was associated with a depletion of soluble thiols and was associated with a depletion of protein thiols. Treatment of IREC with the glutathione reductase inhibitor, 1,3-bis(2-chloroethyl)-1-nitrosourea, potentiated the thiol depletion and toxicity observed with menadione and M(NAC)NQ indicating the involvement of oxidative stress in this model of renal cell toxicity. The lack of MGNQ toxicity can be attributed to an intramolecular cyclization reaction which destroys the quinone nucleus and therefore eliminates its ability to redox cycle. These findings have important implications with regard to our understanding of the toxic potential of quinone thioether conjugates and of quinone toxicity in general.  相似文献   

2.
DT-diaphorase catalysed the reduction of 1,4-naphthoquinones with hydroxy, methyl, methoxy and glutathionyl substituents at the expense of reducing equivalents from NADPH. The initial rates of quinone reduction did not correlate with either the half-wave reduction potential (E1/2) value (determined by h.p.l.c. with electrochemical detection against an Ag/AgCl reference electrode) or the partition coefficient of the quinones. After their reduction by DT-diaphorase the 1,4-naphthoquinone derivatives autoxidized at distinct rates, the extent of which was influenced by the nature of the substituents. Thus for the 1,4-naphthoquinone series the following order of rate of autoxidation was found: 5-hydroxy-1,4-naphthoquinone greater than 3-glutathionyl-1,4-naphthoquinone greater than 5-hydroxy-3-glutathionyl-1,4-naphthoquinone greater than 1,4-naphthoquinone greater than 2-hydroxy-1,4-naphthoquinone. For the 2-methyl-1,4-naphthoquinone (menadione) series the following order was observed: 5-hydroxy-2-methyl-1,4-naphthoquinone greater than 3-glutathionyl-5-hydroxy-2-methyl-1,4-naphthoquinone greater than 3-glutathionyl-2-methyl-1,4-naphthoquinone greater than 2-methyl-1,4-naphthoquinone greater than 3-hydroxy-2-methyl-1,4-naphthoquinone. The autoxidized naphthohydroquinone derivatives were re-reduced by DT-diaphorase, thus closing a cycle of enzymic reduction in equilibrium autoxidation. This was expressed as an excess of NADPH oxidized over the initial concentration of quinone present as well as H2O2 formation. These findings demonstrate that glutathionyl conjugates of 1,4-naphthoquinone and 2-methyl-1,4-naphthoquinone and those of their respective 5-hydroxy derivatives are able to act as substrates for DT-diaphorase and that they also autoxidize at rates higher than those for the unsubstituted parent compounds. These results are discussed in terms of the cellular role of DT-diaphorase in the reduction of hydroxy- or glutathionyl-substituted naphthoquinones as well as the further conjugation of these hydroquinones with glucuronide or sulphate within the cellular milieu, thereby facilitating their disposal from the cells.  相似文献   

3.
In order to clarify the role of oxidative processes in cytotoxicity we have studied the metabolism and toxicity of 2-methyl-1,4-naphthoquinone (menadione) and its 2,3 dimethyl (DMNQ) and 2,3 diethyl (DENQ) analogs in isolated rat hepatocytes. The two analogs, unlike menadione, cannot alkylate nucleophiles directly and were considerably less toxic than menadione. This decreased toxicity was consistent with the inability of DMNQ and DENQ to alkylate but we also found them to undergo lower rates of redox cycling in hepatocytes and a higher ratio of two electron as opposed to one electron reduction relative to menadione. Thus, facile analysis of the respective roles of alkylation and oxidation in cytotoxicity was not possible using these compounds. In hepatocytes pretreated with bischloroethyl-nitrosourea (BCNU) to inhibit glutathione reductase, all three naphthoquinones caused a potentiation of reduced glutathione (GSH) removal/oxidized glutathione (GSSG) generation and cytotoxicity relative to that observed in control cells. These data show that inhibition of hepatocyte glutathione reductase by BCNU results in enhanced naphthoquinone-induced oxidative challenge and subsequent cellular toxicity. That DMNQ and DENQ are cytotoxic, albeit at high concentrations, and that this cytotoxicity is potentiated by BCNU pretreatment suggest that oxidative processes alone can be a determinant of cytotoxicity.  相似文献   

4.
Selected biological effects of 1,4-naphthoquinone, menadione (2-methyl-1,4-naphthoquinone) and structurally related quinones from natural sources - the 5-hydroxy-naphthoquinones juglone, plumbagin and the 2-hydroxy-naphthoquinones lawsone and lapachol - were studied in human keratinocytes (HaCaT). 1,4-naphthoquinone and menadione as well as juglone and plumbagin were highly cytotoxic, strongly induced reactive oxygen species (ROS) formation and depleted cellular glutathione. Moreover, they induced oxidative DNA base damage and accumulation of DNA strand breaks, as demonstrated in an alkaline DNA unwinding assay. Neither lawsone nor lapachol (up to 100 μM) were active in any of these assays. Cytotoxic and oxidative action was paralleled by stimulation of stress signaling: all tested quinones except lawsone and lapachol strongly induced phosphorylation of the epidermal growth factor receptor (EGFR) and the related ErbB2 receptor tyrosine kinase. EGFR activation by plumbagin, juglone and menadione was attenuated by a superoxide dismutase mimetic, indicating that ROS-related mechanisms contribute to EGFR activation by these naphthoquinones.  相似文献   

5.
In the lung, chemical redox cycling generates highly toxic reactive oxygen species that can cause alveolar inflammation and damage to the epithelium, as well as fibrosis. In this study, we identified a cytosolic NADPH-dependent redox cycling activity in mouse lung epithelial cells as sepiapterin reductase (SPR), an enzyme important for the biosynthesis of tetrahydrobiopterin. Human SPR was cloned and characterized. In addition to reducing sepiapterin, SPR mediated chemical redox cycling of bipyridinium herbicides and various quinones; this activity was greatest for 1,2-naphthoquinone followed by 9,10-phenanthrenequinone, 1,4-naphthoquinone, menadione, and 2,3-dimethyl-1,4-naphthoquinone. Whereas redox cycling chemicals inhibited sepiapterin reduction, sepiapterin had no effect on redox cycling. Additionally, inhibitors such as dicoumarol, N-acetylserotonin, and indomethacin blocked sepiapterin reduction, with no effect on redox cycling. Non-redox cycling quinones, including benzoquinone and phenylquinone, were competitive inhibitors of sepiapterin reduction but noncompetitive redox cycling inhibitors. Site-directed mutagenesis of the SPR C-terminal substrate-binding site (D257H) completely inhibited sepiapterin reduction but had minimal effects on redox cycling. These data indicate that SPR-mediated reduction of sepiapterin and redox cycling occur by distinct mechanisms. The identification of SPR as a key enzyme mediating chemical redox cycling suggests that it may be important in generating cytotoxic reactive oxygen species in the lung. This activity, together with inhibition of sepiapterin reduction by redox-active chemicals and consequent deficiencies in tetrahydrobiopterin, may contribute to tissue injury.  相似文献   

6.
The non-cytotoxic concentration (20 microM) of menadione (2-methyl-1,4-naphthoquinone), after 1 h of incubation, leads to loss of the activity of rhodanese by 33%, 3-mercaptopyruvate sulfurtransferase by 20%, as well as the level of sulfane sulfur by about 23% and glutathione by 12%, in the culture of U373 cells, in comparison with the control culture. Reactive oxygen species generated by menadione oxidize sulfhydryl groups in active centers of the investigated enzymes, inhibiting them and saving cysteine for glutathione synthesis. A decreased sulfane sulfur level can be correlated with an oxidative stress.  相似文献   

7.

Menadione (2-methyl-1,4-naphthoquinone) is a synthetic derivative of vitamin K that allows rapid redox cycling in cells and thereby generates reactive oxygen species (ROS). To test for the consequences of a treatment of brain astrocytes with menadione, we incubated primary astrocyte cultures with this compound. Incubation with menadione in concentrations of up to 30 µM did not affect cell viability. In contrast, exposure of astrocytes to 100 µM menadione caused a time-dependent impairment of cellular metabolism and cell functions as demonstrated by impaired glycolytic lactate production and strong increases in the activity of extracellular lactate dehydrogenase and in the number of propidium iodide-positive cells within 4 h of incubation. In addition, already 5 min after exposure of astrocytes to menadione a concentration-dependent increase in the number of ROS-positive cells as well as a concentration-dependent and transient accumulation of cellular glutathione disulfide (GSSG) were observed. The rapid intracellular GSSG accumulation was followed by an export of GSSG that was prevented in the presence of MK571, an inhibitor of the multidrug resistance protein 1 (Mrp1). Menadione-induced glutathione (GSH) oxidation and ROS formation were found accelerated after glucose-deprivation, while the presence of dicoumarol, an inhibitor of the menadione-reducing enzyme NQO1, did not affect the menadione-dependent GSSG accumulation. Our study demonstrates that menadione rapidly depletes cultured astrocytes of GSH via ROS-induced oxidation to GSSG that is subsequently exported via Mrp1.

  相似文献   

8.
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   

9.
Oxidative stress studied in intact mammalian cells   总被引:1,自引:0,他引:1  
Exposure of isolated rat hepatocytes to toxic doses of menadione (2-methyl-1,4-naphthoquinone) results in enhanced formation of active oxygen species, depletion of cellular glutathione and protein thiols, and perturbation of intracellular calcium ion homeostasis. An increase in cytosolic Ca2+ concentration, resulting from inhibition of the plasma membrane Ca2+ translocase by menadione metabolism, appears to be critically involved in the development of cytotoxicity.  相似文献   

10.
Quinones are widely distributed in the environment, both as natural products and as pollutants. This paper reports that one of the simplest quinones, 2-methyl-1,4-naphthoquinone (menadione), effectively inhibited apoptosis in the presence of UVA. Menadione suppressed the apoptosis induced by serum depletion and cell detachment. This effect was significantly enhanced by UVA irradiation. An antioxidant, N-acetylcysteine, completely inhibited the antiapoptotic effects of both menadione itself and menadione plus UVA, and peroxidation of the cells after treatment was observed using a probe to detect the intracellular production of peroxides. By contrast, 2-hydroxy-1,4-naphtoquinone (lawsone) showed no antiapoptotic effect in the presence or absence of UVA. Lawsone is reported not to undergo the redox process that produces reactive oxygen species. These results indicated that intracellular peroxidation contributed to the antiapoptotic effects of both menadione itself and menadione plus UVA. Dysregulation of the apoptotic process is critical to carcinogenesis. The photosensitization of quinone compounds as it relates to the inhibition of apoptosis should be examined in the future.  相似文献   

11.
Rat liver epithelial cells were exposed to three quinones with different properties: menadione (2-methyl-1,4-naphthoquinone, vitamin K3), an alkylating as well as redox-cycling quinone, the strongly alkylating p-benzoquinone (BQ), and the non-arylating redox-cycler, 2,3-dimethoxy-1,4-naphthoquinone (DMNQ). All three quinones induced the activation of extracellular signal-regulated kinase (ERK) 1 and ERK 2 via the activation of epidermal growth factor receptor (EGFR) and MAPK/ERK kinases (MEK) 1/2. ERK activation resulted in phosphorylation at Ser-279 and Ser-282 of the gap junctional protein, connexin-43, known to result in the loss of gap junctional intercellular communication. Another EGFR-dependent pathway was stimulated, leading to the activation of the antiapoptotic kinase Akt via phosphoinositide 3-kinase. The activation of EGFR-dependent signaling by these quinones was by different mechanisms: (i) menadione, but not BQ or DMNQ, inhibited a protein-tyrosine phosphatase regulating the EGFR, as concluded from an EGFR dephosphorylation assay; (ii) although menadione-induced activation of ERK was unimpaired by pretreatment of cells with N-acetyl cysteine, activation by BQ and DMNQ was prevented; (iii) cellular glutathione (GSH) levels were strongly depleted by BQ. The mere depletion of GSH by application of diethyl maleate EGFR-dependently activated ERK and Akt, thus mimicking BQ effects. GSH levels were only moderately decreased by menadione and not affected by DMNQ. In summary, EGFR-dependent signaling was mediated by protein-tyrosine phosphatase inactivation (menadione), GSH depletion (BQ), and redox-cycling (DMNQ), funneling into the same signaling pathway.  相似文献   

12.
Formation of excited species such as singlet molecular oxygen during redox cycling (one-electron reduction-oxidation) was detected by low-level chemiluminescence emitted from perfused rat liver and isolated hepatocytes supplemented with the quinone, menadione (vitamin K3). Chemiluminescence was augmented when the two-electron reduction of the quinone catalyzed by NAD(P)H:quinone reductase was inhibited by dicoumarol, thus underlining the protective function of this enzyme also known as DT-diaphorase. Interference with NADPH supply by inhibition of energy-linked transhydrogenase by rhein or of mitochondrial electron transfer by antimycin A led to a depression in the level of photoemission. Unexpectedly, glutathione depletion of the liver led to a lowering of chemiluminescence elicited by menadione, whereas conversely the depletion of glutathione led to increased chemiluminescence levels when a hydroperoxide was added instead of the quinone. As the GSH conjugate of menadione, 2-methyl-3-glutathionyl-1,4-naphthoquinone, studied with microsomes, was shown also to be capable of redox cycling, we conclude that menadione-induced chemiluminescence of the perfused rat liver does not only arise from menadione itself but from the menadione-GSH conjugate as well. Therefore, the conjugation of the quinone with glutathione is not in itself of protective nature and does not abolish semiquinone formation. A biologically useful aspect of conjugate formation resides in the facilitation of biliary elimination from the liver. Nonenzymatic formation of the conjugate from menadione and GSH in vitro was found to be accompanied by the formation of aggressive oxygen species.  相似文献   

13.
《Free radical research》2013,47(4-6):355-363
Quinones may be toxic by a number of mechanisms. including arylation and oxidative stress caused by redox cycling. Using isolated hepatocytes, we have studied the cytotoxicity of four quinones. with differing abilities to arylate cellular nucleophiles and redox cycle. in relation to their effects on cellular pyridine nucleotides. High concentrations of menadione (redox cycles and arylates). 2-hydroxy-1,4-naphthoquinone (neither arylates nor redox cycles via a one electron reduction) 2.3-dimethoxy-1.4-naphthoquinone (a pure redox cycler) and p-benzoquinone (a pure arylator) caused an initial decrease in NAD+ and loss of viability, which was not prevented by 3-aminobenzamide. an inhibitor of poly(ADP-ribose)polymerase. In contrast. 3-aminobenzamide inhibited the loss of NAD' and viability caused by dimethyl sulphate so implicating poly(ADP-ribose)polymerase in its toxicity but not that of the quinones. Non-toxic concentrations of menadione. 2.3-dimethoxy-1.4-naphthoquinone and 2-hydroxy-1.4-naphthoquinone all caused markedly similar changes in cellular pyridine nucleotides. An initial decrease in NAD+ was accompanied by a small. transient increase in NADP+ and followed by a larger. prolonged increase in NADPH and total NADP+ + NADPH. Nucleotide changes were not observed with non-toxic concentrations of p-benzoquinone. Our findings suggest that a primary event in the response of the cell to redox cycling quinones is to bring about an interconversion of pyridine nucleotides. in an attempt to combat the effects of oxidative stress  相似文献   

14.
The effects of menadione (2-methyl-1,4-naphthoquinone) metabolism on intracellular soluble and protein-bound thiols were investigated in freshly isolated rat hepatocytes. Menadione was found to cause a dose-dependent decrease in intracellular glutathione (GSH) level by three different mechanisms: (a) Oxidation of GSH to glutathione disulfide (GSSG) accounted for 75% of the total GSH loss; (b) About 15% of the cellular GSH reacted directly with menadione to produce a GSH-menadione conjugate which, once formed, was excreted by the cells into the medium; (c) A small amount of GSH (approximately 10%) was recovered by reductive treatment of cell protein with NaBH4, indicating that GSH-protein mixed disulfides were also formed as a result of menadione metabolism. Incubation of hepatocytes with high concentrations of menadione (greater than 200 microM) also induced a marked decrease in protein sulfhydryl groups; this was due to arylation as well as oxidation. Binding of menadione represented, however, a relatively small fraction of the total loss of cellular sulfhydryl groups, since it was possible to recover about 80% of the protein thiols by reductive treatments which did not affect protein binding. This suggests that the loss of protein sulfhydryl groups, like that of GSH, was mainly a result of oxidative processes occurring within the cell during the metabolism of menadione.  相似文献   

15.
Quinone-induced cell death is often attributed to oxidative stress during which the formation of DNA strand breaks is thought to play an important role. In this study, extensive DNA damage was observed in human chronic myelogenous leukemic cells (K562) exposed for 15 minutes to low concentrations (15–100 μM) of the redox cycling quinones 2,3-dimethoxy-1,4-naphthoquinone (2,3-diOMe-1,4-NQ) and menadione. However, DNA strand breakage and cell death could not be attributed to oxidative stress as the intracellular level and redox status of the reducing equivalents NADP(H) and GSH were unaffected. The intracellular level of NAD+ was found to correlate well with the extent of DNA repair (r = 0.93, P < 0.02) and cell proliferation (r = 0.96, P < 0.01) in cells exposed to the quinones. In contrast, a significant decrease in the level of intracellular ATP was only observed in cells exposed to menadione (50–100 μM). These results suggest that redox cycling quinones are capable of inducing DNA damage in mammalian cells by a mechanism that does not involve oxidative stress. Following DNA damage, cell death is dependent on the availability of NAD+, which may be key to the rapid repair of strand breaks. © 1995 John Wiley & Sons, Inc.  相似文献   

16.
Monoamine oxidase (MAO) catalyzes the oxidative deamination of biogenic and exogenous amines and its inhibitors have therapeutic value for several conditions including affective disorders, stroke, neurodegenerative diseases and aging. The discovery of 2,3,6-trimethyl-1,4-naphthoquinone (TMN) as a nonselective and reversible inhibitor of MAO, has suggested 1,4-naphthoquinone (1,4-NQ) as a potential scaffold for designing new MAO inhibitors. Combining molecular modeling tools and biochemical assays we evaluate the kinetic and molecular details of the inhibition of human MAO by 1,4-NQ, comparing it with TMN and menadione. Menadione (2-methyl-1,4-naphthoquinone) is a multitarget drug that acts as a precursor of vitamin K and an inducer of mitochondrial permeability transition. Herein we show that MAO-B was inhibited competitively by 1,4-NQ (Ki = 1.4 μM) whereas MAO-A was inhibited by non-competitive mechanism (Ki = 7.7 μM). Contrasting with TMN and 1,4-NQ, menadione exhibited a 60-fold selectivity for MAO-B (Ki = 0.4 μM) in comparison with MAO-A (Ki = 26 μM), which makes it as selective as rasagiline. Fluorescence and molecular modeling data indicated that these inhibitors interact with the flavin moiety at the active site of the enzyme. Additionally, docking studies suggest the phenyl side groups of Tyr407 and Tyr444 (for MAO-A) or Tyr398 and Tyr435 (for MAO-B) play an important role in the interaction of the enzyme with 1,4-NQ scaffold through forces of dispersion as verified for menadione, TMN and 1,4-NQ. Taken together, our findings reveal the molecular details of MAO inhibition by 1,4-NQ scaffold and show for the first time that menadione acts as a competitive and reversible inhibitor of human MAO.  相似文献   

17.
In perfused rat liver menadione elicits substantial oxidation in both the NADPH and GSH redox systems. Biliary excretion of GSSG is increased several-fold. Menadione derivatives appear in the bile predominantly as the menadione-S-glutathione conjugate, thiodione (60%), or as conjugates derived therefrom (17%). About 10% appear as menadione glucuronides. The excretion of taurocholate into bile is strongly inhibited upon menadione infusion. The inhibition of taurocholate excretion is small in livers with a low content of Se-GSH-peroxidase and in glutathione-depleted livers. In these livers intracellular GSSG and biliary GSSG release remain at low values, although menadione still imposes oxidative stress as indicated by an oxidation of intracellular NADPH. Under anoxic conditions menadione has little influence on both the NADPH and GSH redox systems and also on biliary taurocholate excretion. The amount of thiodione released into bile is similar to that found under normoxia, whereas the amount of glucuronidated products almost doubled. We conclude (a) that intracellular formation of GSSG by menadione occurs via the generation of hydrogen peroxide; (b) that the inhibition of biliary taurocholate excretion by menadione is related to the increased formation of glutathione disulfide; and (c) that menadione derivatives show little, if any, contribution to the inhibition of taurocholate excretion.  相似文献   

18.
The cytotoxic effects of many quinones are thought to be mediated through their one-electron reduction to semiquinone radicals, which subsequently enter redox cycles with molecular oxygen to produce active oxygen species and oxidative stress. The two-electron reduction of quinones to diols, mediated by DT-diaphorase (NAD(P)H: (quinone-acceptor) oxidoreductase), may therefore represent a detoxifying pathway which protects the cell from the formation of these reactive intermediates. By using menadione (2-methyl-1,4-naphthoquinone) and isolated hepatocytes, the relative contribution of the two pathways to quinone metabolism has been studied and a protective role for DT-diaphorase demonstrated. Moreover, in the presence of cytotoxic concentrations of menadione rapid changes in intracellular thiol and Ca2+ homeostasis were observed. These were associated with alterations in the surface structure of the hepatocytes which may be an early indication of cytotoxicity.  相似文献   

19.
Quinones may be toxic by a number of mechanisms. including arylation and oxidative stress caused by redox cycling. Using isolated hepatocytes, we have studied the cytotoxicity of four quinones. with differing abilities to arylate cellular nucleophiles and redox cycle. in relation to their effects on cellular pyridine nucleotides. High concentrations of menadione (redox cycles and arylates). 2-hydroxy-1,4-naphthoquinone (neither arylates nor redox cycles via a one electron reduction) 2.3-dimethoxy-1.4-naphthoquinone (a pure redox cycler) and p-benzoquinone (a pure arylator) caused an initial decrease in NAD+ and loss of viability, which was not prevented by 3-aminobenzamide. an inhibitor of poly(ADP-ribose)polymerase. In contrast. 3-aminobenzamide inhibited the loss of NAD' and viability caused by dimethyl sulphate so implicating poly(ADP-ribose)polymerase in its toxicity but not that of the quinones. Non-toxic concentrations of menadione. 2.3-dimethoxy-1.4-naphthoquinone and 2-hydroxy-1.4-naphthoquinone all caused markedly similar changes in cellular pyridine nucleotides. An initial decrease in NAD+ was accompanied by a small. transient increase in NADP+ and followed by a larger. prolonged increase in NADPH and total NADP+ + NADPH. Nucleotide changes were not observed with non-toxic concentrations of p-benzoquinone. Our findings suggest that a primary event in the response of the cell to redox cycling quinones is to bring about an interconversion of pyridine nucleotides. in an attempt to combat the effects of oxidative stress  相似文献   

20.
Prostaglandin A2 (PGA2) stimulates the biosynthesis of gamma-glutamylcysteine synthetase and elevates glutathione (GSH) contents in cultured mammalian cells. To clarify the importance of gamma-glutamylcysteine synthetase induction in the defence of endothelial cells against oxidative stress, the effect of PGA2 on menadione (2-methyl-1,4-naphthoquinone)-induced cell injury was examined. Incubation of porcine aorta endothelial cells with menadione produced marked loss of cellular GSH and protein sulfhydryl groups, followed by leakage of lactic dehydrogenase (LDH) into the culture medium. The LDH leakage and modification of protein thiol was, however, completely prevented by pretreatment of the cells with PGA2. The protective effect of PGA2 was more potent than that of cysteine delivery agents such as methionine, N-acetylcysteine or 2-oxo-4-thiazolidine carboxylic acid (OTC). The results suggest that cellular GSH plays an important role in the defence against oxidative stress, and induction of gamma-glutamylcysteine synthetase is effective for protecting vascular endothelial cells.  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