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.
Our recent study has demonstrated that cellular redox imbalance can directly initiate apoptosis in a mitotic competent PC-12 cell line without the involvement of reactive oxygen species (ROS). However, whether cell apoptosis induced by ROS is, in fact, mediated by a loss of redox balance caused by the oxidant is unresolved. The linkage between oxidant-mediated apoptosis and the induction of cellular redox was examined in PC-12 cells using the oxidant, tert-butylhydroperoxide (TBH). TBH caused cell apoptosis in 24 h that was preceded by an early increase (30 min) in oxidized glutathione (GSSG). Pretreatment with N-acetyl cysteine prevented TBH-induced GSSG increases and cell apoptosis. Altered Bax/BcL-2 expression and release of mitochondrial cytochrome c occurred post-redox imbalance and was kinetically linked to caspase-3 activation and poly ADP-ribose polymerase cleavage. Moreover, cell apoptosis was attenuated by inhibition of caspase-9, but not caspase-8, and blockade of mitochondrial ROS generation and permeability transition pore attenuated caspase 3 activation and cell apoptosis. Collectively, these results show that TBH-induced GSSG elevation is associated with the disruption of mitochondrial integrity, activation of caspase-3 and cell apoptosis. This redox induction of the apoptotic cascade was dissociated from cellular GSH efflux.  相似文献   

2.
We previously showed that tert-butyl hydroperoxide (TBH) induced apoptosis in na?ve rat pheochromocytoma (nPC12) cells that correlated with cellular redox imbalance and mitochondrial apoptotic signaling. In this study, we tested the hypothesis that differentiation of nPC12 cells results in altered susceptibility to TBH utilizing a model of differentiated PC12 (dPC12) cells induced by nerve growth factor. TBH (100 microM) induced dPC12 apoptosis (12% at 24 h) at levels lower than na?ve cells (35%). This resistance was associated with elevated GSH, NADPH (reduced nicotinamide adenine dinucleotide phosphate), TBH metabolism, redox enzyme activities, reduced cellular GSH/GSSG (glutathione disulfide) status and preservation of mitochondrial membrane potential. Altering cellular GSH with ethacrynic acid or N-acetylcysteine, respectively, exacerbated or protected against dPC12 apoptosis. dPC12 apoptosis was mediated by caspase-9 and -3 activation and apoptosis protease activator protein-1 (Apaf-1) expression. These results show that nPC12 transition to dPC12 cells afforded protection against oxidative challenge due to maintenance of reduced GSH/GSSG and decreased Apaf-1 expression.  相似文献   

3.
We have shown previously that sulforaphane (SFN), a constituent of many edible cruciferous vegetables including broccoli, suppresses growth of prostate cancer cells in culture as well as in vivo by causing apoptosis, but the sequence of events leading to cell death is poorly defined. Using PC-3 and DU145 human prostate cancer cells as a model, we now demonstrate, for the first time, that the initial signal for SFN-induced apoptosis is derived from reactive oxygen species (ROS). Exposure of PC-3 cells to growth-suppressive concentrations of SFN resulted in ROS generation, which was accompanied by disruption of mitochondrial membrane potential, cytosolic release of cytochrome c, and apoptosis. All these effects were significantly blocked on pretreatment with N-acetylcysteine and overexpression of catalase. The SFN-induced ROS generation was significantly attenuated on pretreatment with mitochondrial respiratory chain complex I inhibitors, including diphenyleneiodonium chloride and rotenone. SFN treatment also caused a rapid and significant depletion of GSH levels. Collectively, these observations indicate that SFN-induced ROS generation is probably mediated by a nonmitochondrial mechanism involving GSH depletion as well as a mitochondrial component. Ectopic expression of Bcl-xL, but not Bcl-2, in PC-3 cells offered significant protection against the cell death caused by SFN. In addition, SFN treatment resulted in an increase in the level of Fas, activation of caspase-8, and cleavage of Bid. Furthermore, SV40-immortalized mouse embryonic fibroblasts (MEFs) derived from Bid knock-out mice displayed significant resistance toward SFN-induced apoptosis compared with wild-type MEFs. In conclusion, the results of the present study indicate that SFN-induced apoptosis in prostate cancer cells is initiated by ROS generation and that both intrinsic and extrinsic caspase cascades contribute to the cell death caused by this highly promising cancer chemopreventive agent.  相似文献   

4.
Arsenic trioxide (ATO) can regulate many biological functions such as apoptosis and differentiation in various cells. We investigated an involvement of ROS such as H(2)O(2) and O(2)(*-), and GSH in ATO-treated Calu-6 cell death. The levels of intracellular H(2)O(2) were decreased in ATO-treated Calu-6 cells at 72 h. However, the levels of O(2)(*-) were significantly increased. ATO reduced the intracellular GSH content. Many of the cells having depleted GSH contents were dead, as evidenced by the propidium iodine staining. The activity of CuZn-SOD was strongly down-regulated by ATO at 72 h while the activity of Mn-SOD was weakly up-regulated. The activity of catalase was decreased by ATO. ROS scavengers, Tiron and Trimetazidine did not reduce levels of apoptosis and intracellular O(2)(*-) in ATO-treated Calu-6 cells. Tempol showing a decrease in intracellular O(2)(*-) levels reduced the loss of mitochondrial transmembrane potential (DeltaPsi(m)). Treatment with NAC showing the recovery of GSH depletion and the decreased effect on O(2)(*-) levels in ATO-treated cells significantly inhibited apoptosis. In addition, BSO significantly increased the depletion of GSH content and apoptosis in ATO-treated cells. Treatment with SOD and catalase significantly reduced the levels of O(2)(*-) levels in ATO-treated cells, but did not inhibit apoptosis along with non-effect on the recovery of GSH depletion. Taken together, our results suggest that ATO induces apoptosis in Calu-6 cells via the depletion of the intracellular GSH contents rather than the changes of ROS levels.  相似文献   

5.
Atmospheric pressure room temperature plasma jets (APRTP-Js) that can emit a mixture of different active species have recently found entry in various medical applications. Apoptosis is a key event in APRTP-Js-induced cellular toxicity, but the exact biological mechanisms underlying remain elusive. Here, we explored the role of reactive oxygen species (ROS) and reactive nitrogen species (RNS) in APRTP-Js-induced apoptosis using in vitro model of HepG2 cells. We found that APRTP-Js facilitated the accumulation of ROS and RNS in cells, which resulted in the compromised cellular antioxidant defense system, as evidenced by the inactivation of cellular antioxidants including glutathione (GSH), superoxide dismutase (SOD) and catalase. Nitrotyrosine and protein carbonyl content analysis indicated that APRTP-Js treatment caused nitrative and oxidative injury of cells. Meanwhile, intracellular calcium homeostasis was disturbed along with the alteration in the expressions of GRP78, CHOP and pro-caspase12. These effects accumulated and eventually culminated into the cellular dysfunction and endoplasmic reticulum stress (ER stress)-mediated apoptosis. The apoptosis could be markedly attenuated by N-acetylcysteine (NAC, a free radical scavenger), which confirmed the involvement of oxidative and nitrative stress in the process leading to HepG2 cell apoptosis by APRTP-Js treatment.  相似文献   

6.
Inhibition of oxidative stress has been reported to be involved in the cardioprotective effects of hydrogen sulfide (H(2)S) during ischemia/reperfusion (I/R). However, the mechanism whereby H(2)S regulates the level of cardiac reactive oxygen species (ROS) during I/R remains unclear. Therefore, we investigated the effects of H(2)S on pathways that generate and scavenge ROS. Our results show that pretreating rat neonatal cardiomyocytes with NaHS, a H(2)S donor, reduced the levels of ROS during the hypoxia/reoxygenation (H/R) condition. We found that H(2)S inhibited mitochondrial complex IV activity and increased the activities of superoxide dismutases (SODs), including Mn-SOD and CuZn-SOD. Further studies indicated that H(2)S up-regulated the expression of Mn-SOD but not CuZn-SOD. Using a cell-free system, we showed that H(2)S activates CuZn-SOD. An isothermal titration calorimetry (ITC) analysis indicated that H(2)S directly interacts with CuZn-SOD. Taken together, H(2)S inhibits mitochondrial complex IV and activates SOD to decrease the levels of ROS in cardiomyocytes during I/R.  相似文献   

7.
Arsenic trioxide (ATO) can affect many biological functions such as apoptosis and differentiation in various cells. We investigated the involvement of ROS and GSH in ATO-induced HeLa cell death using ROS scavengers, especially N-acetylcysteine (NAC). ATO increased intracellular O(2)(*-) levels and reduced intracellular GSH content. The ROS scavengers, Tempol, Tiron and Trimetazidine, did not significantly reduce levels of ROS or GSH depletion in ATO-treated HeLa cells. Nor did they reduce the apoptosis induced by ATO. In contrast, treatment with NAC reduced ROS levels and GSH depletion in the ATO-treated HeLa cells and prevented ATO-induced apoptosis. Treatment with exogenous SOD and catalase reduced the depletion of GSH content in ATO-treated cells. Catalase strongly protected the cells from ATO-induced apoptosis. In addition, treatment with SOD, catalase and NAC slightly inhibited the G1 phase accumulation induced by ATO. In conclusion, NAC protects HeLa cells from apoptosis induced by ATO by up-regulating intracellular GSH content and partially reducing the production of O(2)(*-).  相似文献   

8.
We examined the acute toxicity of dieldrin, a possible environmental risk factor of Parkinson's disease, in a dopaminergic cell model, PC12 cells, to determine early cellular events underlying the pesticide-induced degenerative processes. EC(50) for 1 h dieldrin exposure was 143 microM for PC12 cells, whereas EC(50) for non-dopaminergic cells was 292-351 microM, indicating that dieldrin is more toxic to dopaminergic cells. Dieldrin also induced rapid, dose-dependent releases of dopamine and its metabolite, DOPAC, resulting in depletion of intracellular dopamine. Additionally, dieldrin exposure caused depolarization of mitochondrial membrane potential in a dose-dependent manner. Flow cytometric analysis showed generation of reactive oxygen species (ROS) within 5 min of dieldrin treatment, and significant increases in lipid peroxidation were also detected following 1 h exposure. ROS generation was remarkably inhibited in the presence of SOD. Dieldrin-induced apoptosis was significantly attenuated by both SOD and MnTBAP (SOD mimetic), suggesting that dieldrin-induced superoxide radicals serve as important signals in initiation of apoptosis. Furthermore, pretreatment with deprenyl (MAO-inhibitor) or alpha-methyl-L-p-tyrosine (TH-inhibitor) also suppressed dieldrin-induced ROS generation and DNA fragmentation. Taken together, these results suggest that rapid release of dopamine and generation of ROS are early cellular events that may account for dieldrin-induced apoptotic cell death in dopaminergic cells.  相似文献   

9.
Although the induction of myocyte apoptosis by ischemia-reperfusion (I/R) is attenuated by ischemic preconditioning (IPC), the underlying mechanism is not fully understood. Phosphatase and tensin homologs deleted on chromosome 10 (PTEN) promotes apoptosis through Akt-dependent and -independent mechanisms. We tested the hypothesis that IPC attenuates the mitochondrial localization of PTEN in the myocardium induced by I/R. Isolated hearts from wild-type mice were exposed to IPC or normal perfusion followed by 30 min of ischemia and reperfusion. IPC attenuated myocardial infarct size and apoptosis after I/R. Heart fractionation showed that mitochondrial PTEN and Bax protein levels and the physical association between them were increased by 30 min of I/R and that IPC attenuated all of these effects of I/R. Muscle-specific PTEN knockout decreased mitochondrial Bax protein levels in the reperfused myocardium and increased cell survival. To determine whether PTEN relocalization to mitochondria was influenced by I/R-induced production of ROS, hearts were perfused with N-acetylcysteine (NAC) to scavenge ROS or H(2)O(2) to mimic I/R-induced ROS. Mitochondrial PTEN protein levels were decreased by NAC and increased by H(2)O(2). PTEN protein overexpression was generated in mouse hearts by adenoviral gene transfer. PTEN overexpression increased mitochondrial PTEN and Bax protein levels and ROS production, whereas muscle-specific PTEN knockout produced the opposite effects. In conclusion, myocardial I/R causes PTEN localization to the mitochondria, related to the generation of ROS; IPC attenuates the mitochondrial localization of PTEN after I/R, potentially inhibiting the translocation of Bax to the mitochondria and resulting in improved cell viability.  相似文献   

10.
Lee YH  Kim SY  Kim JR  Yoh KT  Baek SH  Kim MJ  Ryu SH  Suh PG  Kim JH 《Life sciences》2000,67(7):827-837
Oxidative stress has been implicated in a wide range of cellular damage which includes DNA oxidation, membrane lipid peroxidation, and apoptosis. In our study, we found that overexpression of PLC-beta1 in NIH3T3 fibroblasts protected them from cell death occuring in response to oxidative stress. Cell death caused by treatment with prooxidant tert-butylhydroperoxide (TBH), H2O2, or CdCl2 was considerably suppressed in PLC-beta1 overexpressed NIH/beta1-14 cells in comparison to control NIH/neo cells. However, overexpression of PLC-beta1 failed to protect the cells from toxicity by diamide or KCN. In addition, while accumulation of c-fos mRNA was observed within 30 min of TBH treatment in vector transfected NIH/neo cells, TBH-induced c-fos mRNA generation was completely suppressed in NIH/beta1-14 cells, while that of c-jun and GAPDH was not affected. These findings suggest that PLC-beta1 may play a role in process that can protect cells from oxidative stress-induced cell death.  相似文献   

11.
Low levels of reactive oxygen species (ROS) can function as redox-active signaling messengers, whereas high levels of ROS induce cellular damage. Menadione generates ROS through redox cycling, and high concentrations trigger cell death. Previous work suggests that menadione triggers cytochrome c release from mitochondria, whereas other studies implicate the activation of the mitochondrial permeability transition pore as the mediator of cell death. We investigated menadione-induced cell death in genetically modified cells lacking specific death-associated proteins. In cardiomyocytes, oxidant stress was assessed using the redox sensor RoGFP, expressed in the cytosol or the mitochondrial matrix. Menadione elicited rapid oxidation in both compartments, whereas it decreased mitochondrial potential and triggered cytochrome c redistribution to the cytosol. Cell death was attenuated by N-acetylcysteine and exogenous glutathione or by overexpression of cytosolic or mitochondria-targeted catalase. By contrast, no protection was observed in cells overexpressing Cu,Zn-SOD or Mn-SOD. Overexpression of antiapoptotic Bcl-X(L) protected against staurosporine-induced cell death, but it failed to confer protection against menadione. Genetic deletion of Bax and Bak, cytochrome c, cyclophilin D, or caspase-9 conferred no protection against menadione-induced cell death. However, cells lacking PARP-1 showed a significant decrease in menadione-induced cell death. Thus, menadione induces cell death through the generation of oxidant stress in multiple subcellular compartments, yet cytochrome c, Bax/Bak, caspase-9, and cyclophilin D are dispensable for cell death in this model. These studies suggest that multiple redundant cell death pathways are activated by menadione, but that PARP plays an essential role in mediating each of them.  相似文献   

12.
13.
Copper/zinc-superoxide dismutase (CuZn-SOD) transgenic mice overexpress the gene for human CuZn-SOD. To assess the effects of the overexpression of CuZn-SOD on the brain scavenging systems, we have measured the activities of manganese-SOD (Mn-SOD), catalase, and glutathione peroxidase (GSH-Px) in various regions of the mouse brain. In nontransgenic mice, cytosolic CuZn-SOD activity was highest in the caudate-putamen complex; this was followed by the brainstem and the hippocampus. The lowest activity was observed in the cerebellum. In transgenic mice, there were significant increases of cytosolic CuZn-SOD activity in all of these regions, with ratios varying from a twofold increase in the brainstem to 3.42-fold in the cerebellum in comparison with nontransgenic mice. Particulate Mn-SOD was similarly distributed in all brain regions, and its levels also were significantly increased in superoxide dismutase (SOD)-transgenic mice. In the brains of nontransgenic mice, cytosolic catalase activity was similar in all brain regions except the cortex, which showed less than 50% of the activity observed in the other regions. In transgenic mice, cytosolic catalase activity was significantly increased, with the cortex showing the greatest changes (133%) in comparison with nontransgenic mice. The smallest increases were observed in the hippocampus (34%). In contrast to what was observed for SOD and catalase, there were no significant changes in cytosolic GSH-Px activity in any of the brain regions examined. The present results indicate that, in addition to displaying marked increases in the levels of brain CuZn-SOD activity, SOD-transgenic mice also exhibit increases in other enzymes that scavenge oxygen-based radicals.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

14.
Superoxide dismutases (SODs) are important antioxidant enzymes responsible for the elimination of superoxide radical (O(2)(-)). The manganese-containing SOD (Mn-SOD) has been suggested to have tumor suppressor function and is located in the mitochondria where the majority of O(2)(-) is generated during respiration. Although increased reactive oxygen species (ROS) in cancer cells has long been recognized, the expression of Mn-SOD in cancer and its role in cancer development remain elusive. The present study used a human tissue microarray to analyze Mn-SOD expression in primary ovarian cancer tissues, benign ovarian lesions, and normal ovary epithelium. Significantly higher levels of Mn-SOD protein expression were detected in the malignant tissues compared with normal tissues (p < 0.05). In experimental systems, suppression of Mn-SOD expression by small interfering RNA caused a 70% increase of superoxide in ovarian cancer cells, leading to stimulation of cell proliferation in vitro and more aggressive tumor growth in vivo. Furthermore, stimulation of mitochondrial O(2)(-) production induced an increase of Mn-SOD expression. Our findings suggest that the increase in Mn-SOD expression in ovarian cancer is a cellular response to intrinsic ROS stress and that scavenging of superoxide by SOD may alleviate the ROS stress and thus reduce the simulating effect of ROS on cell growth.  相似文献   

15.
S C Grace 《Life sciences》1990,47(21):1875-1886
Three isozymes of superoxide dismutase (SOD) have been identified and characterized. The iron and manganese isozymes (Fe-SOD and Mn-SOD, respectively) show extensive primary sequence and structural homology, suggesting a common evolutionary ancestor. In contrast, the copper/zinc isozyme (CuZn-SOD) shows no homology with Fe-SOD or Mn-SOD, suggesting an independent origin for this enzyme. The three isozymes are unequally distributed throughout the biological kingdoms and are located in different subcellular compartments. Obligate anaerobes and aerobic diazotrophs contain Fe-SOD exclusively. Facultative aerobes contain either Fe-SOD or Mn-SOD or both. Fe-SOD is found in the cytosol of cyanobacteria while the thylakoid membranes of these organisms contain a tightly bound Mn-SOD. Similarly, most eukaryotic algae contain Fe-SOD in the chloroplast stroma and Mn-SOD bound to the thylakoids. Most higher plants contain a cytosol-specific and a chloroplast-specific CuZn-SOD, and possibly a thylakoid-bound Mn-SOD as well. Plants also contain Mn-SOD in their mitochondria. Likewise, animals and fungi contain a cytosolic CuZn-SOD and a mitochondrial Mn-SOD. The Mn-SOD found in the mitochondria of eukaryotes shows strong homology to the prokaryotic form of the enzyme. Taken together, the phylogenetic distribution and subcellular localization of the SOD isozymes provide strong support for the hypothesis that the chloroplasts and mitochondria of eukaryotic cells arose from prokaryotic endosymbionts.  相似文献   

16.
J Liu  H M Shen  C N Ong 《Life sciences》2001,69(16):1833-1850
Recent studies have demonstrated that induction of apoptosis is related to the cell growth inhibition potential of Salvia Miltiorrhiza (SM), a traditional herbal medicine. In the present study, we further explore the mechanistic pathway involved in SM-induced apoptosis in human hepatoma HepG2 cells. A rapid decline of intracellular glutathione (GSH) and protein thiol content was found in SM-treated cells. Moreover. SM exposure resulted in mitochondrial dysfunction as demonstrated by: (i) the onset of mitochondrial permeability transition (MPT); (ii) the disruption of mitochondrial membrane potential (MMP); and (iii) the release of cytochrome c from mitochondria into the cytosol. Subsequently, elevated level of intracellular reactive oxygen species (ROS) was observed prior to the onset of DNA fragmentation. However, no caspase-3 cleavage was observed throughout the whole period of SM treatment, while a caspase-3-independent poly(ADP-ribose) polymerase (PARP) cleavage was noted at the late stage in SM-induced apoptosis. Pretreatment of cells with N-acetylcysteine (NAC), the GSH synthesis precursor, conferred complete protection against MMP loss, ROS generation and apoptosis induced by SM. MPT inhibitors, cyclosporin A plus trifluoperazine, partially restored intracellular GSH content, and reduced SM-induced ROS formation and subsequently inhibited cell death. Moreover, antioxidants NAC, deferoxamine and catalase had little effect on GSH depletion and mitochondrial dysfunction, yet still were able to completely protect cells from SM-induced apoptosis. Taken together, our results suggest that SM deplete intracellular thiols, which, in turn, causes MPT and subsequent increase in ROS generation, and eventually apoptotic cell death.  相似文献   

17.
Fan S  Li L  Chen S  Yu Y  Qi M  Tashiro S  Onodera S  Ikejima T 《Free radical research》2011,45(11-12):1307-1324
Silibinin, as the major active constituent of silymarin, has its various biological effects. Here, we investigated the inhibitory effects of silibinin on HeLa cell growth in relation to autophagy and apoptosis induced by reactive oxygen species (ROS) and reactive nitrogen species (RNS) generation. Silibinin dose and time-dependently decreased cell growth cultured in medium containing 10% fetal bovine serum or in serum free media (SFM) with an IC(50) of approximately 80-100 and 40-60 μM at 24 h, respectively. Silibinin induced autophagy at 12 h, confirmed by monodansylcadervarine (MDC) staining and up-regulation of beclin-1, and induced apoptosis at 24 h, detected by observation of apoptotic bodies and activation of caspase-3. 3-methyladenine (3-MA) inhibited silibinin-induced autophagy and attenuated the silibinin's inhibitory effect on cell viability, suggesting that autophagy enhanced silibinin-induced cell death. Silibinin increased ROS levels at 12 h, and ROS scavenger, N-acetylcysteine (NAC), significantly reversed the cytotoxicity of silibinin through inhibiting both autophagy and apoptosis. Specific antioxidants were applied and results indicated that hydroxyl radical (·OH) was the major ROS induced by silibinin, and OH scavenger glutathione (GSH) inhibited apoptosis and autophagy. Silibinin also generated RNS production in the cells at 12 h. High concentration of N omega-nitro-l-arginine methyl ester (L-NAME) as nitric oxide synthase (NOS) inhibitor attenuated the cytotoxicity of silibinin by decreasing ROS levels, leading to down-regulation of apoptosis. Silibinin also could interrupt the respiring functions of mitochondria, leading to ROS production and oxidative damage.  相似文献   

18.
In previous works, we have established a correlation between antioxidant system response and tolerance to drought, osmotic stress and photooxidative stress of different wheat cultivars with contrasting drought tolerance. In the present work, a protocol to obtain and transform wheat protoplasts was established. Transgenic protoplasts with Manganese Superoxide Dismutase (Mn-SOD) (E.C.: 1.15.1.1) and Glutathione Reductase (GR) (E.C.: 1.6.4.2) overexpression in chloroplasts were obtained, and their responses to photooxidative stress were characterized. Protoplasts with Mn-SOD or GR overexpression, showed different responses and tolerance to photooxidative stress. Protoplasts with Mn-SOD overexpression showed lower levels of oxidative damage, higher level of endogenous hydrogen peroxide and a great induction of total SOD and GR activities during photooxidative treatments. In protoplasts with GR overexpression the oxidative damage provoked by the photooxidative treatment was similar to control protoplasts, the GSH content and GSH/GSH + GSSG ratio were higher than control and Mn-SOD transformed protoplast, and total SOD and GR activities were not induced. Our results suggest that the differential responses and tolerance to photooxidative stress given by Mn-SOD or GR overexpression, also depend on the effects of these enzyme activities over the cellular redox state balance, which modulate the responses to photooxidative stress.  相似文献   

19.
UV-induced apoptosis in keratinocytes is a highly complex process in which various molecular pathways are involved. These include the extrinsic pathway via triggering of death receptors and the intrinsic pathway via DNA damage and reactive oxygen species (ROS) formation. In this study we investigated the effect of catalase and CuZn-superoxide dismutase (SOD) overexpression on apoptosis induced by UVB exposure at room temperature or 4 degrees C on normal human keratinocytes. Irradiation at low temperature reduced UV-induced apoptosis by 40% in normal keratinocytes independently of any change in p53 and with a decrease in caspase-8 activation. Catalase overexpression decreased apoptosis by 40% with a reduction of caspase-9 activation accompanied by a decrease in p53. Keeping cells at low temperature and catalase overexpression had additive effects. CuZn-SOD overexpression had no significant effect on UVB-induced apoptosis. UVB induced an increase in ROS levels at two distinct stages: immediately following irradiation and around 3 h after irradiation. Catalase overexpression inhibited only the late increase in ROS levels. We conclude that catalase overexpression has a protective role against UVB irradiation by preventing DNA damage mediated by the late ROS increase.  相似文献   

20.
alpha-Hederin, a pentacyclic triterpene saponin isolated from the seeds of Nigella sativa, was recently reported to have potent in vivo antitumor activity against LL/2 (Lewis Lung carcinoma) in BDF1 mice. In this study we observed that alpha-hederin caused a dose- and time-dependent increase in apoptosis of murine leukemia P388 cells. In order to evaluate the possible mechanisms for apoptosis, the effects of alpha-hederin on intracellular thiol concentration, including reduced glutathione (GSH), and protein thiols, and the effects of pretreatment with N-acetlycysteine (NAC), a precursor of intracellular GSH synthesis, or buthionine sulfoxime (BSO), a specific inhibitor of intracellular GSH synthesis, on alpha-hederin-induced apoptosis were investigated. It was found that alpha-hederin rapidly depleted intracellular GSH and protein thiols prior to the occurrence of apoptosis. NAC significantly alleviated alpha-hederin-induced apoptosis, while BSO augmented alpha-hederin-induced apoptosis significantly. The depletion of cellular thiols observed after alpha-hederin treatment caused disruption of mitochondrial membrane potential (deltapsi(m)) and subsequently increased the production of reactive oxygen species (ROS) in P388 cells at an early time point. Bongkrekic acid (BA), a ligand of the mitochondrial adenine nucleotide translocator, and cyclosporin (CsA) attenuated the alpha-hederin-induced loss of deltapsi(m), and ROS production. Thus, oxidative stress after alpha-hederin treatment is an important event in alpha-hederin-induced apoptosis. As observed in this study, permeability transition of mitochondrial membrane occurs after depletion of GSH and precedes a state of reactive oxygen species (ROS) generation. Further, we observed that alpha-hederin caused the release of cytochrome c from the mitochondria to cytosol, leading to caspase-3 activation. Our findings thus demonstrate that changes in intracellular thiols and redox status leading to perturbance of mitochondrial functions are important components in the mechanism of alpha-hederin-induced cell death.  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