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.
The nuclear factor Acinus has been suggested to mediate apoptotic chromatin condensation after caspase cleavage. However, this role has been challenged by recent observations suggesting a contribution of Acinus in apoptotic internucleosomal DNA cleavage. We report here that AAC‐11, a survival protein whose expression prevents apoptosis that occurs on deprivation of growth factors, physiologically binds to Acinus and prevents Acinus‐mediated DNA fragmentation. AAC‐11 was able to protect Acinus from caspase‐3 cleavage in vivo and in vitro, thus interfering with its biological function. Interestingly, AAC‐11 depletion markedly increased cellular sensitivity to anticancer drugs, whereas its expression interfered with drug‐induced cell death. AAC‐11 possesses a leucine‐zipper domain that dictates, upon oligomerization, its interaction with Acinus as well as the antiapoptotic effect of AAC‐11 on drug‐induced cell death. A cell permeable peptide that mimics the leucine‐zipper subdomain of AAC‐11, thus preventing its oligomerization, inhibited the AAC‐11–Acinus complex formation and potentiated drug‐mediated apoptosis in cancer cells. Our results, therefore, show that targeting AAC‐11 might be a potent strategy for cancer treatment by sensitization of tumour cells to chemotherapeutic drugs.  相似文献   

2.
Abstract The role of cytochrome c in insect cell apoptosis has drawn considerable attention and has been subject to considerable controversy. In Drosophila, the majority of studies have demonstrated that cytochrome c may not be involved in apoptosis, although there are conflicting reports. Cytochrome c is not released from mitochondria into the cytosol and activation of the initiator caspase Dronc or effector caspase Drice is not associated with cytochrome c during apoptosis in Drosophila SL2 cells or BG2 cells. Cytochrome c failed to induce caspase activation and promote caspase activation in Drosophila cell lysates, but remarkably caused caspase activation in extracts from human cells. Knockdown of cytochrome c does not protect cells from apoptosis and over‐expression of cytochrome c also does not promote apoptosis. Structural analysis has revealed that cytochrome c is not required for Dapaf‐1 complex assembly. In Lepidoptera, the involvement of cytochrome c in apoptosis has been demonstrated by the accumulating evidence. Cytochrome c release from mitochondria into cytosol has been observed in different cell lines such as Spodoptera frugiperda Sf9, Spodoptera litura Sl‐1 and Lymantria dispar LdFB. Silencing of cytochrome c expression significantly affected apoptosis and activation of caspase and the addition of cytochrome c to cell‐free extracts results in caspase activation, suggesting the activation of caspase is dependent on cytochrome c. Although Apaf‐1 has not been identified in Lepidoptera, the inhibitor of apoptosome formation can inhibit apoptosis and caspase activation. Cytochrome c may be exclusively required for Lepidoptera apoptosis.  相似文献   

3.
This study explored the effects involved in silencing CLIC4 on apoptosis and proliferation of mouse liver cancer Hca‐F and Hca‐P cells. A CLIC4‐target small interfering RNA (siRNA) was designed to compound into two individual complementary oligonucleotide chains. A process of annealing and connection to a pSilencer vector was followed by transfection with Hca‐F and Hca‐P cells. Quantitative real‐time polymerase chain reaction and Western blotting techniques were used to determine CLIC4 mRNA and protein expressions. CCK8 assay and flow cytometry were employed for analysis of the survival and apoptosis rate as well as the cell cycle in an octreotide‐induced apoptosis model. Expressions of caspase 3, caspase 9, and cleaved PARP were measured using Western blotting. The CLIC4 mRNA and protein expressions in Hca‐F and Hca‐P cells transfected by pSilencer‐CLIC4 siRNA plasmid in the blank group displayed remarkably decreased levels of expression, when compared with both the control and negative control (NC) groups. Decreased survival rates and cleaved PARP expression, increased cell apoptosis rate,expressions of caspase 3 and caspase 9 in Hca‐F and Hca‐P cells were detected in groups that had been cultured in a medium containing octreotide. The pSilencer‐CLIC4 siRNA‐2 group when compared with the control and NC groups exhibited decreased survival rates, cleaved PARP expression, increased cell apoptosis rates, and increased expressions of caspase 3 and caspase 9 of Hca‐F and Hca‐P cells. The results demonstrated that siRNA‐induced down‐regulation of CLIC4 could proliferation, while in turn promoting apoptosis of mouse liver cancer Hca‐F and Hca‐P cells. J. Cell. Biochem. 119: 659–668, 2018. © 2017 Wiley Periodicals, Inc.  相似文献   

4.
Cytotoxicity of many plant and bacterial toxins requires their endocytosis and retrograde transport from endosomes to the endoplasmic reticulum. Using cell fractionation and immunoblotting procedures, we have assessed the fate and action of the plant toxin ricin in rat liver in vivo, focusing on endosome‐associated events and induction of apoptosis. Injected ricin rapidly accumulated in endosomes as an intact A/B heterodimer (5–90 min) and was later (15–90 min) partially translocated to cytosol as A‐ and B‐chains. Unlike cholera and diphtheria toxins, which also undergo endocytosis in liver, neither in cell‐free endosomes loaded by ricin in vivo nor upon incubation with endosomal lysates did ricin undergo degradation in vitro. A time‐dependent translocation of ricin across the endosomal membrane occurred in cell‐free endosomes. Endosome‐located thioredoxin reductase‐1 was required for translocation as shown by its physical association with ricin chains and effects of its removal and inhibition. Ricin induced in vivo intrinsic apoptosis as judged by increased cytochrome c content, activation of caspase‐9 and caspase‐3, and enrichment of DNA fragments in cytosol. Furthermore, reduced ricin and ricin B‐chain caused cytochrome c release from mitochondria in vivo and in vitro, suggesting that the interaction of ricin B‐chain with mitochondria is involved in ricin‐induced apoptosis.  相似文献   

5.
6.
Chlamydia trachomatis is an obligate intracellular bacterial pathogen of medical importance. C. trachomatis develops inside a membranous vacuole in the cytosol of epithelial cells but manipulates the host cell in numerous ways. One prominent effect of chlamydial infection is the inhibition of apoptosis in the host cell, but molecular aspects of this inhibition are unclear. Tumour necrosis factor (TNF) is a cytokine with important roles in immunity, which is produced by immune cells in chlamydial infection and which can have pro‐apoptotic and non‐apoptotic signalling activity. We here analysed the signalling through TNF in cells infected with C. trachomatis. The pro‐apoptotic signal of TNF involves the activation of caspase‐8 and is controlled by inhibitor of apoptosis proteins. We found that in C. trachomatis‐infected cells, TNF‐induced apoptosis was blocked upstream of caspase‐8 activation even when inhibitor of apoptosis proteins were inhibited or the inhibitor of caspase‐8 activation, cFLIP, was targeted by RNAi. However, when caspase‐8 was directly activated by experimental over‐expression of its upstream adapter Fas‐associated protein with death domain, C. trachomatis was unable to inhibit apoptosis. Non‐apoptotic TNF‐signalling, particularly the activation of NF‐κB, initiates at the plasma membrane, while the activation of caspase‐8 and pro‐apoptotic signalling occur subsequently to internalization of TNF receptor and the formation of a cytosolic signalling complex. In C. trachomatis‐infected cells, NF‐κB activation through TNF was unaffected, while the internalization of the TNF–TNF‐receptor complex was blocked, explaining the lack of caspase‐8 activation. These results identify a dichotomy of TNF signalling in C. trachomatis‐infected cells: Apoptosis is blocked at the internalization of the TNF receptor, but non‐apoptotic signalling through this receptor remains intact, permitting a response to this cytokine at sites of infection.  相似文献   

7.
Natural substances have been attracted several researchers in the recent years, because of its potential antioxidant, anti‐inflammatory and anti‐cancer properties. We have investigated the effect of carnosine on cell viability, apoptosis, DNA damage, reactive oxygen species (ROS) and caspase 3 enzyme expression in human cervical carcinoma and Madin‐Darby Kidney Cells (MDCK) cells . Carnosine inhibited cancer cell growth up to 23%. ROS level was increased up to 30 and 31% in MDCK and HeLa cells respectively. Tunnel assay showed 42 and 14% of positive apoptotic cells in cancer and normal cells respectively. The alteration in mitochondrial and nuclear morphology was determined. The extended lace‐like network of normal mitochondria found in control cells. Carnosine treatment significantly altered the mitochondrial morphology of normal cervical carcinoma cell. Mitochondria were condensed clump structures in carnosine treated cancer cells. Carnosine reduced the number of colonies of cervical carcinoma cells. Caspase 3 expression was corresponded to the appearance of immunofluorescence in the cytoplasm. Caspase 3 expression was gradually increased in cervical carcinoma cells. In Silico, docking study was performed to recognize the binding activity of carnosine against a subunit of the caspase 3 , and carnosine was able to bind to the drug binding pocket of caspase 3. The glide energy is ?5.2 kcal/mol, suggesting the high binding affinity of carnosine to caspase 3. Taking all these data together, the natural dipeptide L‐carnosine could be a suitable antiproliferative agent in cervical carcinoma cells. Copyright © 2016 John Wiley & Sons, Ltd.  相似文献   

8.
Both neurons and glia succumb to programmed cell death (PCD) when deprived of growth factors at critical periods in development or following injury. Insulin‐like growth factor‐I (IGF‐I) prevents apoptosis in neurons in vitro. To investigate whether IGF‐I can protect Schwann cells (SC) from apoptosis, SC were harvested from postnatal day 3 rats and maintained in serum‐containing media until confluency. When cells were switched to serum‐free defined media (DM) for 12–72 h, they underwent PCD. Addition of insulin or IGF‐I prevented apoptosis. Bisbenzamide staining revealed nuclear condensation and formation of apoptotic bodies in SC grown in DM alone, but SC grown in DM plus IGF‐I had normal nuclear morphology. The phosphatidylinositol 3‐kinase (PI 3‐K) inhibitor LY294002 blocked IGF‐I–mediated protection. Caspase‐3 activity was rapidly activated upon serum withdrawal in SC, and the caspase inhibitor BAF blocked apoptosis. These results suggest that IGF‐I rescues SC from apoptosis via PI 3‐K signaling which is upstream from caspase activation. © 1999 John Wiley & Sons, Inc. J Neurobiol 41: 540–548, 1999  相似文献   

9.
Many viruses induce cell death and lysis as part of their replication and dissemination strategy, and in many cases features of apoptosis are observed. Attempts have been made to further increase productivity by prolonging cell survival via the over‐expression of anti‐apoptotic genes. Here, we extend the study to investigate the association between virus replication and apoptosis, pertinent to large‐scale vector production for gene therapy. Infection of an HEK293 cell line with a replication defective type‐5‐adenovirus expressing a GFP reporter (Ad5GFP) resulted in rapid decline in viability associated with increased virus titer. The over‐expression of bcl‐2 resulted in improved cell resistance to apoptosis and prolonged culture duration, but reduced virus specific and total productivity. In contrast, the over‐expression of pro‐caspase‐3 (Yama/CPP32/apopain) resulted in reduced cell survival but increased virus productivity. The treatment of infected cells with caspase inhibitors support the preposition that caspase‐3 dependent apoptosis, and to a lesser degree caspase‐9 dependent apoptosis, represent important steps in virus production, thus implicating the intrinsic apoptosis pathway in the production of adenovirus from HEK293 cells. The suppression of apoptosis by the over‐expression of XIAP (inhibitors of caspase family cell death proteases) further shows that caspase‐mediated activation plays an important role in virus infection and maturation. Biotechnol. Bioeng. 2009; 104: 752–765 © 2009 Wiley Periodicals, Inc.  相似文献   

10.
Quercetin, a widely distributed bioflavonoid, has been shown to induce growth inhibition in a variety of human cancer cells. However, the regulation of survivin and Bcl‐2 on the quercetin‐induced cell‐growth inhibition and apoptosis in cancer cells remains unclear. In the present study, we report that quercetin can inhibit proliferation and induce apoptosis in HepG2 cells in dose‐ and time‐dependent manner. Hoechst 33258 and acridine orange/ethidium bromide (AO/EB) staining showed that HepG2 cells underwent the typical morphologic changes of apoptosis characterized by nuclear shrinkage, chromatin condensation, or fragmentation after exposure to quercetin. Cell‐cycle analysis reveals a significant increase of the proportion of cells in G0/G1 phase. We also demonstrate that the levels of survivin and Bcl‐2 protein expression in HepG2 cells decreased concurrently, and the levels of p53 protein increased significantly after treatment with quercetin by immunocytochemistry analysis. Relative activity of caspase‐3 and caspase‐9 increased significantly. These data clearly indicate that quercetin‐induced apoptosis is associated with caspase activation, and the levels of survivin and Bcl‐2. Our results indicate that the expression of survivin may be associated with Bcl‐2 expression, and the inhibition expression of survivin, in conjunction with Bcl‐2, might cause more pronounced apoptotic effects. Together, concurrent down‐regulated survivin and Bcl‐2 play an important role in HepG2 cell apoptosis induced by quercetin.  相似文献   

11.
Corilagin is a component of Phyllanthus urinaria extract and has been found of possessing anti‐inflammatory, anti‐oxidative, and anti‐tumour properties in clinic treatments. However, the underlying mechanisms in anti‐cancer particularly of its induction of cell death in human breast cancer remain undefined. Our research found that corilagin‐induced apoptotic and autophagic cell death depending on reactive oxygen species (ROS) in human breast cancer cell, and it occurred in human breast cancer cell (MCF‐7) only comparing with normal cells. The expression of procaspase‐8, procaspase‐3, PARP, Bcl‐2 and procaspase‐9 was down‐regulated while caspase‐8, cleaved PARP, caspase‐9 and Bax were up‐regulated after corilagin treatment, indicating apoptosis mediated by extrinsic and mitochondrial pathways occurred in MCF‐7 cell. Meanwhile, autophagy mediated by suppressing Akt/mTOR/p70S6K pathway was detected with an increase in autophagic vacuoles and LC3‐II conversion. More significantly, inhibition of autophagy by chloroquine diphosphate salt (CQ) remarkably enhanced apoptosis, while the caspase inhibitor z‐VAD‐fmk failed in affecting autophagy, suggesting that corilagin‐induced autophagy functioned as a survival mechanism in MCF‐7 cells. In addition, corilagin induced intracellular reactive oxygen species (ROS) generation, when reduced by ROS scavenger NAC, apoptosis and autophagy were both down‐regulated. Nevertheless, in SK‐BR3 cell which expressed RIP3, necroptosis inhibitor Nec‐1 could not alleviate cell death induced by corilagin, indicating necroptosis was not triggered. Subcutaneous tumour growth in nude mice was attenuated by corilagin, consisting with the results in vitro. These results imply that corilagin inhibits cancer cell proliferation through inducing apoptosis and autophagy which regulated by ROS release.  相似文献   

12.
Swainsonine (SW) is an indolizidine alkaloid isolated from a number of poisonous plants. We have previously reported that SW inhibited luteal cell progesterone production by inducing caprine luteal cell apoptosis in vitro; however, the molecular mechanism of this phenomenon remains unclear. In this study, SW‐treated luteal cells showed apoptosis characteristics, including nuclear fragmentation, DNA ladder formation, and phosphatidylserine externalization. Further studies showed that SW activated caspase‐9 and caspase‐3, which subsequently cleaved poly(ADP‐ribose) polymerase. SW also increased in Bax/BcL‐2 ratios, promoted Bax translocation from the cytosol to mitochondria, and triggered the release of cytochrome c from mitochondria into the cytoplasm. However, Fas and Fas ligand induction or caspase‐8 activity did not appear any significant changes. Additional analysis also showed that pan‐caspase inhibitor, caspase‐9 inhibitor, or caspase‐3 inhibitor almost completely protected the cells from SW‐induced apoptosis, but not caspase‐8 inhibitor. Overall, these data demonstrated that SW induced luteal cells apoptosis through a mitochondrial‐mediated caspase‐dependent pathway.  相似文献   

13.
To investigate the effects of PA‐MSHA (Pseudomonas aeruginosa‐mannose sensitive hemagglutinin) on inhibiting proliferation of breast cancer cell lines and to explore its mechanisms of action in human breast cancer cells. MCF‐10A, MCF‐7, MDA‐MB‐468, and MDA‐MB‐231HM cells were treated with PA‐MSHA or PA (Heat‐killed P. aeruginosa) at different concentrations and different times. Changes of cell super‐microstructure were observed by transmission electron microscopy. Cell cycle distribution and apoptosis induced by PA‐MSHA were measured by flow cytometry (FCM) with PI staining, ANNEXIN V‐FITC staining and Hoechst33258 staining under fluorescence microscopy. Western blot was used to evaluate the expression level of apoptosis‐related molecules. A time‐dependent and concentration‐dependent cytotoxic effect of PA‐MSHA was observed in MDA‐MB‐468 and MDA‐MB‐231HM cells but not in MCF‐10A or MCF‐7 cells. The advent of PA‐MSHA changed cell morphology, that is to say, increases in autophagosomes, and vacuoles in the cytoplasm could also be observed. FCM with PI staining, ANNEXIN V‐FITC and Hoechst33258 staining showed that the different concentrations of PA‐MSHA could all induce the apoptosis and G0–G1 cell cycle arrest of breast cancer cells. Cleaved caspase 3, 8, 9, and Fas protein expression levels were strongly associated with an increase in apoptosis of the breast cancer cells. There was a direct relationship with increased concentrations of PA‐MSHA but not of PA. Completely different from PA, PA‐MSHA may impart antiproliferative effects against breast cancer cells by inducing apoptosis mediated by at least a death receptor‐related cell apoptosis signal pathway, and affecting the cell cycle regulation machinery. J. Cell. Biochem. 108: 195–206, 2009. © 2009 Wiley‐Liss, Inc.  相似文献   

14.
Ceramide is a member of the sphingolipid family of bioactive molecules demonstrated to have profound, diverse biological activities. Ceramide is a potential chemotherapeutic agent via the induction of apoptosis. Exposure to ceramide activates extracellular‐signal‐regulated kinases (ERK)1/2‐ and p38 kinase‐dependent apoptosis in human ovarian cancer OVCAR‐3 cells, concomitant with an increase in the expression of COX‐2 and p53 phosphorylation. Blockade of cyclooxygenase‐2 (COX‐2) activity by siRNA or NS398 correspondingly inhibited ceramide‐induced p53 Ser‐15 phosphorylation and apoptosis; thus COX‐2 appears at the apex of the p38 kinase‐mediated signaling cascade induced by ceramide. Induction of apoptosis by ceramide or resveratrol was inhibited by the endocytosis inhibitor, cytochalasin D (CytD); however, cells exposed to resveratrol showed greater sensitivity than ceramide‐treated cells. Ceramide‐treated cells underwent a dose‐dependent reduction in trans‐membrane potential. Although both ceramide and resveratrol induced the expressions of caspase‐3 and ‐7, the effect of inducible COX‐2 was different in caspase‐7 expression induced by ceramide compared to resveratrol. In summary, resveratrol and ceramide converge on an endocytosis‐requiring, ERK1/2‐dependent signal transduction pathway and induction of COX‐expression as an essential molecular antecedent for subsequent p53‐dependent apoptosis. In addition, expressions of caspase‐3 and ‐7 are observed. However, a p38 kinase‐dependent signal transduction pathway and change in mitochondrial potential are also involved in ceramide‐induced apoptosis. J. Cell. Biochem. 114: 1940–1954, 2013. © 2013 Wiley Periodicals, Inc.  相似文献   

15.
Background information. Caspase‐dependent and ‐independent death mechanisms are involved in apoptosis in a variety of human carcinoma cells treated with antineoplastic compounds. Our laboratory has reported that p53 is a key contributor of mitochondrial apoptosis in cervical carcinoma cells after staurosporine exposure. However, higher mitochondrial membrane potential dissipation and greater DNA fragmentation were observed in p53wt (wild‐type p53) HeLa cells compared with p53mt (mutated p53) C‐33A cells. Here, we have studied events linked to the mitochondrial apoptotic pathway. Results. Staurosporine can induce death of HeLa cells via a cytochrome c/caspase‐9/caspase‐3 mitochondrial‐dependent apoptotic pathway and via a delayed caspase‐independent pathway. In contrast with p53wt cells, p53mt C‐33A cells exhibit firstly caspase‐8 activation leading to caspase‐3 activation and Bid cleavage followed by cytochrome c release. Attenuation of PARP‐1 [poly(ADP‐ribose) polymerase‐1] cleavage as well as oligonucleosomal DNA fragmentation in the presence of z‐VAD‐fmk points toward a major involvement of a caspase‐dependent pathway in staurosporine‐induced apoptosis in p53wt HeLa cells, which is not the case in p53mt C‐33A cells. Meanwhile, the use of 3‐aminobenzamide, a PARP‐1 inhibitor known to prevent AIF (apoptosis‐inducing factor) release, significantly decreases staurosporine‐induced death in these p53mt carcinoma cells, suggesting a preferential implication of caspase‐independent apoptosis. On the other hand, we show that p53, whose activity is modulated by pifithrin‐α, isolated as a suppressor of p53‐mediated transactivation, or by PRIMA‐1 (p53 reactivation and induction of massive apoptosis), that reactivates mutant p53, causes cytochrome c release as well as mitochondrio—nuclear AIF translocation in staurosporine‐induced apoptosis of cervical carcinoma cells. Conclusions. The present paper highlights that staurosporine engages the intrinsic mitochondrial apoptotic pathway via caspase‐8 or caspase‐9 signalling cascades and via caspase‐independent cell death, as well as through p53 activity.  相似文献   

16.
17.
Background and Aims. H. pylori infection results in an increased epithelial apoptosis in gastritis and duodenal ulcer patients. We investigated the role and type of activation of caspases in H. pylori‐induced apoptosis in gastric epithelial cells. Methods. Differentiated human gastric cancer cells (AGS) and human gastric mucous cell primary cultures were incubated with H. pylori for 0.5–24 hours in RPMI 1640 medium, and the effects on cell viability, epithelial apoptosis, and activity of caspases were monitored. Apoptosis was analyzed by detection of DNA‐fragments by Hoechst stain®, DNA‐laddering, and Histone‐ELISA. Activities of caspases were determined in fluorogenic assays and by Western blotting. Cleavage of BID and release of cytochrome c were analyzed by Western blot. Significance of caspase activation was investigated by preincubation of gastric epithelial cells with cell permeable specific caspase inhibitors. Results. Incubation of gastric epithelial cells with H. pylori caused a time and concentration dependent induction of DNA fragmentation (3‐fold increase), cleavage of BID, release of cytochrome c and a concomittant sequential activation of caspase‐9 (4‐fold), caspase‐8 (2‐fold), caspase‐6 (2‐fold), and caspase‐3 (6‐fold). No effects on caspase‐1 and ‐7 were observed. Activation of caspases preceded the induction of DNA fragmentation. Apoptosis could be inhibited by prior incubation with the inhibitors of caspase‐3, ‐8, and ‐9, but not with that of caspase‐1. Conclusions. Activation of certain caspases and activation of the mitochondrial apoptotic pathway are essential for H. pylori induced apoptosis in gastric epithelial cells.  相似文献   

18.
Some short peptides discovered by phage display are found to be able to inhibit cancer growth and induce cancer cell apoptosis. In this study, a novel cancer‐targeting short peptide which was composed of 22 amino acids (ACHWPWCHGWHSACDLPMHPMC, abbreviated as sp22) and specifically bound to human CD59 was screened from a M13 phage display library so as to counteract tumor immune escape activity. The mechanism of exogenous sp22 peptide in inducing apoptosis of MCF‐7 cells was investigated. The results suggested that sp22 could lower CD59 expression level, downregulate Bcl‐2 expression, activate Fas and caspase‐3, and finally increase apoptotic cell numbers of MCF‐7 cells. However, sp22 had no obvious influence on normal human embryonic lung cells. In addition, the effects of endogenous sp22 gene on CD59 expression and NKM cell apoptosis were explored using the recombinant plasmid sp22‐PIRES. It showed that sp22 gene was efficiently expressed in transfected NKM cells. Compared with normal NKM cells, NKM cells transfected with sp22 displayed reduced mRNA and protein expression levels of CD59, increased sensitivity to complement‐mediated cytolysis, decreased cell survival ratio, changes of the expression of apoptosis associated proteins, increased number of apoptotic cells and the appearance of apoptotic morphology. The results suggested that sp22 protein could bind to CD59 and inhibit the expression of CD59. The cytolytic activity of complement on tumor cells strengthened and apoptosis signal was stepwise transferred which might be a potential way to kill tumor cells. J. Cell. Biochem. 113: 3810–3822, 2012. © 2012 Wiley Periodicals, Inc.  相似文献   

19.
Caspase activation and dependence on caspases has been observed in different paradigms of apoptotic cell death in vivo and in vitro. The present study examines the role of caspases in ionizing radiation‐induced apoptosis in the developing cerebellum of rats subjected to a single dose (2‐Gy γ rays) of whole‐body irradiation at postnatal day 3. Radiation‐induced apoptosis in the external granule cell layer, as defined by the presence of cells by extremely condensed, often fragmented nucleus, which were stained with the method of in situ end‐labeling of nuclear DNA fragmentation, first appeared at 3 h and peaked at 6 h following irradiation. Increased expression of the precursors of caspase 1 (ICE), 2 (Nedd2), 3 (CPP32), 6 (Mch2), and 8 (Mch5 and FLICE), and increased expression of active caspase 3, as revealed by immunohistochemistry, were observed in the external granule cell layer of the cerebellum. Radiation‐induced apoptosis was accompanied by an increase in the expression of the poly(ADP‐ribose) polymerase (PARP) fragment of about 89 kD, as revealed by Western blots of cerebellar homogenates. This was not associated with modifications of protein kinase Cδ and Lamin B. Concomitant injection in the culmen of the cerebellum in irradiated rats of high doses of Y‐VAD‐cmk, DEV‐fmk, or IETD‐fmk resulted in decreased expression of the PARP fragment in cerebellar homogenates. This was accompanied by a decrease in the expression of active caspase 3, as shown by immunohistochemistry. These observations suggest caspase activation following ionizing radiation. However, no differences in the number and morphological and biochemical characteristics of apoptotic cells, including strong nuclear and cytoplasmic c‐Jun/AP‐1 (N) expression, were observed between irradiated and both irradiated and caspase inhibitor–treated rats. Taken together, these observations suggest that the caspases examined are not essential for radiation‐induced apoptosis in the developing cerebellum. © 1999 John Wiley & Sons, Inc. J Neurobiol 41: 549–558, 1999  相似文献   

20.
Apoptosis ensures tissue homeostasis in response to developmental cues or cellular damage. Recently reported genome‐wide RNAi screens have suggested that several metabolic regulators can modulate caspase activation in Drosophila. Here, we establish a previously unrecognized link between metabolism and Drosophila apoptosis by showing that cellular NADPH levels modulate the initiator caspase Dronc through its phosphorylation at S130. Depletion of NADPH removed this inhibitory phosphorylation, resulting in the activation of Dronc and subsequent cell death. Conversely, upregulation of NADPH prevented Dronc‐mediated apoptosis upon DIAP1 RNAi or cycloheximide treatment. Furthermore, this CaMKII‐mediated phosphorylation of Dronc hindered Dronc activation, but not its catalytic activity. Blockade of NADPH production aggravated the death‐inducing activity of Dronc in specific neurons, but not in the photoreceptor cells of the eyes of transgenic flies; similarly, non‐phosphorylatable Dronc was more potent than wild type in triggering specific neuronal apoptosis. Our observations reveal a novel regulatory circuitry in Drosophila apoptosis, and, as NADPH levels are elevated in cancer cells, also provide a genetic model to understand aberrations in cancer cell apoptosis resulting from metabolic alterations.  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