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 role of the tumor suppressor protein p53 in apoptosis of mouse hepatoma cells was studied. Different lines were used which were either p53 wild-type or carried various types of heterozygous or homozygous p53 mutations. The presence of mutations was demonstrated to correlate with a lack in transactivating activity of p53. While UV-light effectively produced apoptosis in cells of all lines, irrespective of their p53 mutational status, gamma-irradiation induced the formation of micronuclei but failed to induce apoptosis. Both UV- and gamma-irradiation led to nuclear accumulation and increases in p53 protein in p53 wild-type cells. Similarly, no significant differences in apoptotic response between p53 wild-type and p53 mutated cells were seen with other apoptotic stimuli like CD95/APO-1/Fas or TNFalpha. These data suggest that wild-type p53 is not required for induction of apoptosis in mouse hepatoma cells which may explain the apparent lack of p53 mutations in mouse liver tumors.  相似文献   

2.
Cisplatin treatment of Epstein-Barr virus-immortalized human B lymphoblastoid cell lines (LCLs) results in p53-mediated apoptosis which occurs largely in a population of cells at the G1/S boundary of the cell cycle. Cell cycle progression appears to be required for this apoptosis because arresting cells earlier in G1 inhibited apoptosis despite the accumulation of p53. Overexpression of wild-type p53 also induces apoptosis in an LCL. Therefore six mutant genes derived from Burkitt's lymphoma (BL) cells were assayed for their ability to induce apoptosis when similarly overexpressed. The same genes were analysed in transient transfection assays for their ability to transactivate appropriate reporter plasmids. A correlation between the ability of p53 to transactivate and induce apoptosis was revealed. The only mutant capable of transactivation also induced apoptosis. Further analysis of the BL lines in which p53 had been characterized showed that whereas some lines were essentially resistant to cisplatin, three were rapidly induced to undergo apoptosis. All three have a single p53 allele encoding a mutant which is incapable of transactivation or (for two tested) mediating apoptosis when expressed in an LCL. Cell cycle analysis revealed that this apparently p53-independent apoptosis did not follow G1 arrest but in fact occurred largely in cells distributed in the G2/M phase of the cell cycle. These data suggest the existence of a second checkpoint in the G2 or M phase which, in the absence of a functional p53, is the primary point of entry into the apoptosis programme following DNA damage.  相似文献   

3.
Activation of the p53 protein can lead to apoptosis and cell cycle arrest. In contrast, activation of the signalling pathway controlled by the Kit receptor tyrosine kinase prevents apoptosis and promotes cell division of a number of different cell types in vivo. We have investigated the consequences of activating the Kit signalling pathway by its ligand Steel factor on these opposing functions of the p53 protein in Friend erythroleukemia cells. A temperature-sensitive p53 allele (Val-135) was introduced into the Friend erythroleukemia cell line (DP-16) which lacks endogenous p53 expression. At 38.5 degrees C, the Val-135 protein maintains a mutant conformation and has no effect on cell growth. At 32 degrees C, the mutant protein assumes wild-type properties and induces these cells to arrest in G1, terminally differentiate, and die by apoptosis. We demonstrate that Steel factor inhibits p53-mediated apoptosis and differentiation but has no effect on p53-mediated G1/S cell cycle arrest. These results demonstrate that Steel factor functions as a cell survival factor in part through the suppression of differentiation and apoptosis induced by p53 and suggest that cell cycle arrest and apoptosis may be separable functions of p53.  相似文献   

4.
The tumor suppressor p53 has been implicated in gamma irradiation-induced apoptosis. To investigate possible consequences of wild-type p53 loss in leukemia, we studied the effect of a single dose of gamma irradiation upon p53-deficient human T-ALL (acute lymphoblastic leukemia) CCRF - CEM cells. Exposure to 3 - 96 Gy caused p53-independent cell death in a dose and time-dependent fashion. By electron microscopic and other criteria, this cell death was classified as apoptosis. At low to intermediate levels of irradiation, apoptosis was preceded by accumulation of cells in the G2/M phase of the cell division cycle. Expression of Bcl-2 and Bax were not detectably altered after irradiation. Expression of the temperature sensitive mouse p53 V135 mutant induced apoptosis on its own but only slightly increased the sensitivity of CCRF - CEM cells to gamma irradiation. Thus, in these, and perhaps other leukemia cells, a p53- and Bcl-2/Bax-independent mechanism is operative that efficiently senses irradiation effects and translates this signal into arrest in the G2/M phase of the cell cycle and subsequent apoptosis.  相似文献   

5.
The calcium ionophore ionomycin cooperates with the S100B protein to rescue a p53-dependent G(1) checkpoint control in S100B-expressing mouse embryo fibroblasts and rat embryo fibroblasts (REF cells) which express the temperature-sensitive p53Val135 mutant (C. Scotto, J. C. Deloulme, D. Rousseau, E. Chambaz, and J. Baudier, Mol. Cell. Biol. 18:4272-4281, 1998). We investigated in this study the contributions of S100B and calcium-dependent PKC (cPKC) signalling pathways to the activation of wild-type p53. We first confirmed that S100B expression in mouse embryo fibroblasts enhanced specific nuclear accumulation of wild-type p53. We next demonstrated that wild-type p53 nuclear translocation and accumulation is dependent on cPKC activity. Mutation of the five putative cPKC phosphorylation sites on murine p53 into alanine or aspartic residues had no significant effect on p53 nuclear localization, suggesting that the cPKC effect on p53 nuclear translocation is indirect. A concerted regulation by S100B and cPKC of wild-type p53 nuclear translocation and activation was confirmed with REF cells expressing S100B (S100B-REF cells) overexpressing the temperature-sensitive p53Val135 mutant. Stimulation of S100B-REF cells with the PKC activator phorbol ester phorbol myristate acetate (PMA) promoted specific nuclear translocation of the wild-type p53Val135 species in cells positioned in early G(1) phase of the cell cycle. PMA also substituted for ionomycin in the mediating of p53-dependent G(1) arrest at the nonpermissive temperature (37.5 degrees C). PMA-dependent growth arrest was linked to the cell apoptosis response to UV irradiation. In contrast, growth arrest mediated by a temperature shift to 32 degrees C protected S100B-REF cells from apoptosis. Our results suggest a model in which calcium signalling, linked with cPKC activation, cooperates with S100B to promote wild-type p53 nuclear translocation in early G(1) phase and activation of a p53-dependent G(1) checkpoint control.  相似文献   

6.
Many cell lines derived from neuroblastoma (NB) carry the wild-type p53 gene with a p53-dependent apoptotic pathway that is responsive to DNA damaging agents. A recent study has demonstrated that retinoic acid (RA) pretreatment of NB cells promotes chemoresistance to apoptosis induced by chemotherapeutic agents. We examine here the possible contribution of the p53 pathway to the chemoresistance response associated with the RA treatment in NB cells. Upon treatment with RA (1-10 microM) for 4 days, the human NB cells, SH-SY5Y, developed resistance selectively to p53-dependent apoptotic stimuli including gamma-irradiation, etoposide, and 1-(5-isoquinolinyl sulfonyl)-2-methylpiperazine (H-7). Interestingly, RA affected the ability of H-7 to induce nuclear accumulation of the p53 protein without altering its effect on elevating the steady-state level of p53, suggesting that drug-induced up-regulation and nuclear accumulation of the wild-type p53 protein are separable processes. The modulation of nuclear import of p53 protein by RA may thus represent a potential mechanism by which certain tumor cells with the wild-type p53 gene develop resistance to chemotherapeutic agents.  相似文献   

7.
Although DNA-damaging agents such as ultraviolet (UV) and X-ray can induce apoptosis, the difference in the apoptotic mechanism is not clearly understood. In the present study, we investigated the effects of these two genotoxic agents on the induction of DNA damage and subsequent apoptotic cell death from the viewpoint of cell cycle regulation by using WiDr cells. Transient G1 arrest was observed after UV exposure, whereas G2 but not G1 arrest was induced after X-ray irradiation. UV-exposure could induce G1 arrest in both mutant-type (mt-p53) and wild-type p53 (wt-p53) cells, but obvious G1 arrest was not observed in the cells lacking in p53 expression. An increase in the DNA fragmentation was observed at S phase in UV-irradiated cells and at G2 phase in X-irradiated cells, respectively. UV-irradiated cells showed an increase production of p53 protein and accumulation of p21 protein. On the contrary, both p53 and p21 proteins remained at a low level in X-irradiated cells. Treatment with aphidicolin, an S phase blocking agent, prolonged cell cycle arrest and reduced the rate of apoptotic cell death in both UV-irradiated and X-irradiated cells. From these results, it is suggested that UV-induced apoptosis occurs mainly at S phase and is regulated by increased production of p53 and p21 proteins, while X-ray-induced apoptosis occurs after G2 blockade and may be independent of p53.  相似文献   

8.
Murine erythroleukemia cells that lack endogenous p53 expression were transfected with a temperature-sensitive p53 allele. The temperature-sensitive p53 protein behaves as a mutant polypeptide at 37 degrees C and as a wild-type polypeptide at 32 degrees C. Three independent clones expressing the temperature-sensitive p53 protein were characterized with respect to p53-mediated G1 cell cycle arrest, apoptosis, and differentiation. Clone ts5.203 responded to p53 activation at 32 degrees C by undergoing G1 arrest, apoptosis, and differentiation. Apoptosis was seen in cells representative of all phases of the cell cycle and was not restricted to cells arrested in G1. The addition of a cytokine (erythropoietin, c-kit ligand, or interleukin-3) to the culture medium of ts5.203 cells blocked p53-mediated apoptosis and differentiation but not p53-mediated G1 arrest. These observations indicate that apoptosis and G1 arrest can be effectively uncoupled through the action of cytokines acting as survival factors and are consistent with the idea that apoptosis and G1 arrest represent separate functions of p53. Clones ts15.15 and tsCB3.4 responded to p53 activation at 32 degrees C by undergoing G1 arrest but not apoptosis. We demonstrate that tsCB3.4 secretes a factor with erythropoietin-like activity and that ts15.15 secretes a factor with interleukin-3 activity and suggest that autocrine secretion of these cytokines blocks p53-mediated apoptosis. These data provide a framework in which to understand the variable responses of cells to p53 overexpression.  相似文献   

9.
10.
The mismatch repair (MMR) system and p53 protein play a pivotal role in maintaining genomic stability and modulate cell chemosensitivity. Aim of this study was to examine the effects of either MMR-deficiency or p53 inactivation, or both, on cellular responses to bleomycin. The MMR-deficient colon carcinoma cell line HCT116 and its MMR-proficient subline HCT116/3-6, both expressing wild-type p53, were transfected with an expression vector encoding a dominant-negative p53 mutant, or with the empty vector. Four transfected clones, having the following phenotypes, MMR-proficient/p53 wild-type, MMR-proficient/p53 mutant, MMR-deficient/p53 wild-type, MMR-deficient/p53 mutant, were subjected to treatment with bleomycin. Loss of MMR function alone was associated with increased resistance to apoptosis, chromosomal damage and inhibition of colony formation caused by bleomycin. Loss of p53 alone resulted in abrogation of G1 arrest and increased sensitivity to apoptosis and chromosomal damage induced by the drug, but did not affect clonogenic survival after bleomycin treatment. Disabling both p53 and MMR function led to abrogation of G1 arrest and to a moderate impairment of drug-induced apoptosis. Chromosomal damage was reduced in the MMR-deficient/p53 mutant clone with respect to the MMR-proficient/p53 wild-type one, when evaluated 48 h after bleomycin treatment, but was comparable in both clones 96 h after drug exposure. Clonogenic survival of the MMR-deficient/p53 mutant clone was similar to that of the MMR-deficient/p53 wild-type one. The effects of MMR-deficiency on cellular responses to bleomycin were confirmed using the MMR-proficient lymphoblastoid cell line TK6 and its MMR-deficient subline MT1, both expressing wild-type p53. In conclusion, our data show that loss of MMR and p53 function exerts opposite and independent effects on apoptosis and chromosomal damage induced by bleomycin. Moreover, inactivation of MMR confers resistance to the cytotoxic activity of the anticancer agent in cells expressing either wild-type or mutant p53.  相似文献   

11.
We previously reported that nitric oxide (NO) released from S-nitrosoglutathione induces conformational change of the p53 tumor-suppressor protein that impairs its DNA-binding activity in vitro. We now demonstrate that MCF-7 cells preincubated in the presence of 0.5-1 mM S-nitrosoglutathione for 4 h before gamma-irradiation failed to arrest in the G1 phase of the cell cycle, whereas those gamma-irradiated without S-nitrosoglutathione exhibited a normal cell cycle arrest. The S-nitrosoglutathione-treated cells did not express the p53 target gene p21(waf-1) after gamma-irradiation, although p21(waf-1) was strongly expressed in cells irradiated in the absence of S-nitrosoglutathione. These results strongly suggest that NO impairs the function of p53 possibly via conformational change and/or amino acid modifications. On the other hand, cells incubated for 16 h in the presence of 1 mM S-nitrosoglutathione underwent apoptosis with accumulation of the pro-apoptotic protein Bax. This Bax accumulation, however, was shown to occur via a p53-independent pathway.  相似文献   

12.
Skp2 regulates G2/M progression in a p53-dependent manner   总被引:1,自引:0,他引:1  
Targeted proteasomal degradation mediated by E3 ubiquitin ligases controls cell cycle progression, and alterations in their activities likely contribute to malignant cell proliferation. S phase kinase-associated protein 2 (Skp2) is the F-box component of an E3 ubiquitin ligase complex that targets p27Kip1 and cyclin E1 to the proteasome. In human melanoma, Skp2 is highly expressed, regulated by mutant B-RAF, and required for cell growth. We show that Skp2 depletion in melanoma cells resulted in a tetraploid cell cycle arrest. Surprisingly, co-knockdown of p27Kip1 or cyclin E1 failed to prevent the tetraploid arrest induced by Skp2 knockdown. Enhanced Aurora A phosphorylation and repression of G2/M regulators cyclin B1, cyclin-dependent kinase 1, and cyclin A indicated a G2/early M phase arrest in Skp2-depleted cells. Furthermore, expression of nuclear localized cyclin B1 prevented tetraploid accumulation after Skp2 knockdown. The p53 status is most frequently wild type in melanoma, and the tetraploid arrest and down-regulation of G2/M regulatory genes were strongly dependent on wild-type p53 expression. In mutant p53 melanoma lines, Skp2 depletion did not induce cell cycle arrest despite up-regulation of p27Kip1. These data indicate that elevated Skp2 expression may overcome p53-dependent cell cycle checkpoints in melanoma cells and highlight Skp2 actions that are independent of p27Kip1 degradation.  相似文献   

13.
It has been convincingly demonstrated that genotoxic stresses cause the accumulation of the tumor suppressor gene p53. One important consequence of increased p53 protein levels in response to DNA damage is the activation of a G1-phase cell cycle checkpoint. It has also been shown that G1-phase cell cycle checkpoints are activated in response to other stresses, such as lack of oxygen. Here we show that hypoxia and heat, agents that induce cellular stress primarily by inhibiting oxygen-dependent metabolism and denaturing proteins, respectively, also cause an increase in p53 protein levels. The p53 protein induced by heat is localized in the cytoplasm and forms a complex with the heat shock protein hsc70. The increase in nuclear p53 protein levels and DNA-binding activity and the induction of reporter gene constructs containing p53 binding sites following hypoxia occur in cells that are wild type for p53 but not in cells that possess mutant p53. However, unlike ionizing radiation, the accumulation of cells in G1 phase by hypoxia is not strictly dependent on wild-type p53 function. In addition, cells expressing the human papillomavirus E6 gene, which show increased degradation of p53 by ubiquitination and fail to accumulate p53 in response to DNA-damaging agents, do increase their p53 levels following heat and hypoxia. These results suggest that hypoxia is an example of a "nongenotoxic" stress which induces p53 activity by a different pathway than DNA-damaging agents.  相似文献   

14.
Melanoma is the most aggressive of skin cancers because of its high resistance to currently available therapy. Although melanoma cells often retain wild-type p53 tumour suppressor protein and express it at high levels, the p53 mediated apoptosis pathway is suppressed. Histone deacetylase (HDAC) inhibitors are a promising group of compounds inducing differentiation, growth arrest and apoptosis in tumour cells in preclinical studies. We have studied the cellular effects of trichostatin A (TSA), a HDAC inhibitor, in a panel of melanoma cell lines and its mechanism of action in relation to p53. TSA stabilized wild-type p53, but p53 protein accumulation was overridden by simultaneous downregulation of p53 mRNA leading to a decrease in p53 protein. While growth arrest was induced in all cell lines studied and apoptosis in most (6/7), these cellular effects were independent of the p53 status of the cells. Inhibiting p53 function by a dominant negative p53 (p53(175His)) confirmed that the HDAC inhibitor induced apoptosis was independent of wild-type p53, even though TSA slightly activated p53 in a reporter assay. The results indicate that while the action of TSA is independent of p53, the activation of the apoptosis pathway by the HDAC inhibitors may provide therapeutic approaches for melanoma treatment.  相似文献   

15.
A p53-derived C-terminal peptide induced rapid apoptosis in breast cancer cell lines carrying endogenous p53 mutations or overexpressed wild-type (wt) p53 but was not toxic to nonmalignant human cell lines containing wt p53. Apoptosis occurred through a Fas/APO-1 signaling pathway involving increased extracellular levels of Fas/FasL in the absence of protein synthesis, as well as activation of a Fas/APO-1-specific protease, FLICE. The peptide activity was p53-dependent, and it had no effect in three tumor cell lines with null p53. Furthermore, the C-terminal peptide bound to p53 protein in cell extracts. Thus, p53-dependent, Fas/APO-1 mediated apoptosis can be induced in breast cancer cells with mutant p53 similar to the recently described Fas/APO-1 induced apoptosis by wt p53. However, mutant p53 without p53 peptide does not induce a Fas/APO-1 activation or apoptosis. Docking of the computed low energy conformations for the C-terminal peptide with those for a recently defined proline-rich regulatory region from the N-terminal domain of p53 suggests a unique low energy complex between the two peptide domains. The selective and rapid induction of apoptosis in cancer cells carrying p53 abnormalities may lead to a novel therapeutic modality.  相似文献   

16.
17.
BACKGROUND: Human colon cancers have a high frequency of p53 mutations, and cancer cells expressing mutant p53 tend to be resistant to current chemo- and radiation therapy. It is thus important to find therapeutic agents that can inhibit colon cancer cells with altered p53 status. beta-Lapachone, a novel topoisomerase inhibitor, has been shown to induce cell death in human promyelocytic leukemia and prostate cancer cells through a p53-independent pathway. Here we examined the effects of beta-lapachone on human colon cancer cells. MATERIALS AND METHODS: Several human colon cancer cell lines, SW480, SW620, and DLD1, with mutant or defective p53, were used. The antiproliferative effects of beta-lapachone were assessed by colony formation assays, cell cycle analysis, and apoptosis analysis, including annexin V staining and DNA laddering analysis. The effects on cell cycle and apoptosis regulatory proteins were examined by immunoblotting. RESULTS: All three cell lines, SW480, SW620, and DLD1, were sensitive to beta-lapachone, with an IC(50) of 2 to 3 microM in colony formation assays, a finding similar to that previously reported for prostate cancer cells. However, these cells were arrested in different stages of S phase. At 24 hr post-treatment, beta-lapachone induced S-, late S/G2-, and early S-phase arrest in SW480, SW620, and DLD1 cells, respectively. The cell cycle alterations induced by beta-lapachone were congruous with changes in cell cycle regulatory proteins such as cyclin A, cyclin B1, cdc2, and cyclin D1. Moreover, beta-lapachone induced apoptosis, as demonstrated by annexin V staining, flow cytometric analysis of DNA content, and DNA laddering analysis. Furthermore, down-regulation of mutant p53 and induction of p27 in SW480 cells, and induction of pro-apoptotic protein Bax in DLD1 cells may be pertinent to the anti-proliferative and apoptotic effects of beta-lapachone on these cells. CONCLUSIONS: beta-Lapachone induced cell cycle arrest and apoptosis in human colon cancer cells through a p53-independent pathway. For human colon cancers, which often contain p53 mutations, beta-lapachone may prove to be a promising anticancer agent that can target cancer cells, especially those with mutant p53.  相似文献   

18.
Alterations of the p53 gene have been attributed a major role in the development and resistance to therapy of several human cancers. Accumulation of p53 in tumor cells may result from mutations associated with prolonged half-life or from stabilization of wild-type p53 by different mechanisms. To address the role of p53 accumulation in the response of malignant glioma cells to radiochemotherapy, we expressed the p53 mutant p53(V143A) in five human malignant glioma cell lines with different genetic and functional p53 status. Accumulation of p53(V143A) modulated proliferation in three and clonogenicity in four of five cell lines without a clear pattern with regard to their endogenous p53 status. p53(V143A) inhibited the camptothecin-induced accumulation of p21(WAF1/CIP1) in cell lines with p53 functional wild-type activity, but not in cell lines lacking p53 activity, consistent with a transdominant-negative effect of p53(V143A). Irradiation induced a moderate G2/M arrest in all cell lines, irrespective of the p53 status, that was unaffected by p53(V143A). Radiosensitivity as well as sensitivity to BCNU, teniposide (VM26), topotecan, vincristine, Taxol, and cisplatin both in cytotoxic cell death and in clonogenic cell death was unchanged in p53(V143A)-transfected cells with few exceptions. These data do not support the hypothesis that accumulation of mutant p53 is a major determinant of the response to adjuvant radiochemotherapy in human malignant glioma cells.  相似文献   

19.
E1A expression during adenovirus infection induces apoptosis. E1A expression causes accumulation of the p53 tumor suppressor protein, and E1A-induced apoptosis is p53 mediated in primary rodent cells, implying that p53 induction may be linked to apoptosis induction by E1A. Adenoviruses containing mutations in the E1A gene were tested for the ability to trigger both p53 accumulation and the appearance of enhanced cytopathy (cyt phenotype) and degradation of DNA (deg phenotype), indicative of apoptosis in infected HeLa cells. The adenoviruses had mutations which disrupted the pRb- and/or p300-binding activities of E1A so that the relationship between p53 induction and apoptosis and binding to these cellular proteins by E1A could be determined. An E1A mutation that specifically disrupted the p300-binding activity failed to induce p53 accumulation, whereas mutations in E1A which affected pRb binding induced p53 accumulation. Thus, p300 binding was required and pRb binding was dispensable for E1A-mediated accumulation of p53 in HeLa cells. All the E1A mutant viruses, regardless of the ability to induce p53 accumulation, induced the cyt and deg phenotypes, suggesting that p53 induction in infected HeLa cells was not essential for apoptosis, nor was binding of E1A to the pRb and/or p300 protein. The possibility that E1A induced a p53-independent apoptosis pathway was tested by analyzing the appearance of the cyt and deg phenotypes in Saos-2 cells, which were null for both alleles of p53, upon adenovirus infection. An adenovirus expressing wild-type 12S E1A induced both the cyt and deg phenotypes in Saos-2 cells, as did all the E1A mutant viruses. Thus, E1A expression during infection of human cells may trigger redundant p53-independent and -dependent apoptotic pathways.  相似文献   

20.
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