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 46 毫秒
1.
2.
3.
4.
5.
We report that transfection of insulin-like growth factor-binding protein-3 (IGFBP-3) cDNA in human breast cancer cell lines expressing either mutant p53 (T47D) or wild-type p53 (MCF-7) induces apoptosis. IGFBP-3 also increases the ratio of pro-apoptotic to anti-apoptotic members of the Bcl-2 family. In MCF-7, an increase in Bad and Bax protein expression and a decrease in Bcl-x(L) protein and Bcl-2 protein and mRNA were observed. In T47D, Bax and Bad proteins were up-regulated; Bcl-2 protein is undetectable in these cells. As T47D expresses mutant p53 protein, these modulations of pro-apoptotic proteins and induction of apoptosis are independent of p53. The effect of IGFBP-3 on the response of T47D to ionizing radiation (IR) was examined. These cells do not G(1) arrest in response to IR and are relatively radioresistant. Transfection of IGFBP-3 increased the radiosensitivity of T47D and increased IR-induced apoptosis but did not effect a rapid G(1) arrest. IR also caused a much greater increase in Bax protein in IGFBP-3 transfectants compared with vector controls. Thus, IGFBP-3 increases the expression of pro-apoptotic proteins and apoptosis both basally and in response to IR, suggesting it may be a p53-independent effector of apoptosis in breast cancer cells via its modulation of the Bax:Bcl-2 protein ratio.  相似文献   

6.
7.
After an initial burst of cell proliferation, the type 1 insulin-like growth factor receptor (IGF-IR) induces granulocytic differentiation of 32D IGF-IR cells, an interleukin-3-dependent murine hemopoietic cell line devoid of insulin receptor substrate-1 (IRS-1). The combined expression of the IGF-IR and IRS-1 (32D IGF-IR/IRS-1 cells) inhibits IGF-I-mediated differentiation, and causes malignant transformation of 32D cells. Because of the role of IRS-1 in changing the fate of 32D IGF-IR cells from differentiation (and subsequent cell death) to malignant transformation, we have looked for differences in IGF-IR signaling between 32D IGF-IR and 32D IGF-IR/IRS-1 cells. In this report, we have focused on p70(S6K), which is activated by the IRS-1 pathway. We find that the ectopic expression of IRS-1 and the inhibition of differentiation correlated with a sustained activation of p70(S6K) and an increase in cell size. Phosphorylation in vivo of threonine 389 and, to a lesser extent, of threonine 421/serine 424 of p70(S6K) seemed to be a requirement for inhibition of differentiation. A role of IRS-1 and p70(S6K) in the alternative between transformation or differentiation of 32D IGF-IR cells was confirmed by findings that inhibition of p70(S6K) activation or IRS-1 signaling, by rapamycin or okadaic acid, induced differentiation of 32D IGF-IR/IRS-1 cells. We have also found that the expression of myeloperoxidase mRNA (a marker of differentiation, which sharply increases in 32D IGF-IR cells), does not increase in 32D IGF-IR/IRS-1 cells, suggesting that the expression of IRS-1 in 32D IGF-IR cells causes the extinction of the differentiation program initiated by the IGF-IR, while leaving intact its proliferation program.  相似文献   

8.
9.
To examine the p53-mediated biological activities and signalling pathways, we generated stable transfectants of the p53-null IW32 murine erythroleukemia cells expressing the temperature-sensitive p53 mutant DNA, tsp53(val135). Two clones with different levels of p53 protein expression were selected for further characterization. At permissive temperature, clone 1-5 cells differentiated along the erythroid pathway, and clone 3-2 cells that produced greater levels (3.5-fold) of p53 underwent apoptosis. Apoptosis of 3-2 cells was accompanied by mitochondrial cytochrome c release and caspase activation as well as by cleavage of caspase substrates. Bax protein was induced to a similar extent in these clones by wild-type p53; expression of p21(Cip1/Waf1) and p27(Kip1) proteins was also increased. However, significantly lesser extent of induction for both CDK inhibitors was detected in the apoptotic 3-2 clone. The general caspase inhibitor benzyloxycarbonyl-Val-Ala-Asp-fluoromethyl ketone (z-VAD.fmk) blocked the p53-induced apoptosis in 3-2 cells, with a concomitant elevation of p27(Kip1), suggesting that p27(Kip1) protein underwent caspase-dependent proteolysis in the apoptotic 3-2 cells. Together these results linked a pathway involving cytochrome c release, caspase activation and p27(Kip1) degradation to the p53-induced apoptosis in IW32 erythroleukemia cells.  相似文献   

10.
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   

11.
12.
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   

13.
This study examined the effects of p53 gene status on DNA damage-induced cell death and chemosensitivity to various chemotherapeutic agents in non-small cell lung cancer (NSCLC) cells. A mutant p53 gene was introduced into cells carrying the wild-type p53 gene and also vice versa to introduce the wild-type p53 gene into cells carrying the mutant p53 gene. Chemosensitivity and DNA damage-induced apoptosis in these cells were then examined. This study included five cell lines, NCI-H1437, NCI-H727, NCI-H441 and NCI-H1299 which carry a mutant p53 gene and NCI-H460 which carries a wild-type p53 gene. Mutant p53-carrying cells were transfected with the wild-type p53 gene, while mutant p53 genes were introduced into NCI-H460 cells. These p53 genes were individually mutated at amino acid residues 143, 175, 248 and 273. The representative cell line NCI-H1437 cells transfected with wild-type p53 gene (H1437/wtp53) showed a dramatic increase in susceptibility to three anticancer agents (7-fold to cisplatin, 21-fold to etoposide, and 20-fold to camptothecin) compared to untransfected or neotransfected H1437 cells. An increase in chemosensitivity was also observed in wild-type p53 transfectants of H727, H441, H1299 cells. The results of chemosensitivity were consistent with the observations on apoptotic cell death. H1437/wtp53 cells, but not H1437 parental cells, exhibited a characteristic feature of apoptotic cell death that generated oligonucleosomal-sized DNA fragments. In contrast, loss of chemosensitivity and lack of p53-mediated DNA degradation in response to anticancer agents were observed in H460 cells transfected with mutant p53. These observations suggest that the increase in chemosensitivity was attributable to wild-type p53 mediation of the process of apoptosis. In addition, our results also suggest that p53 gene status modulates the extent of chemosensitivity and the induction of apoptosis by different anticancer agents in NSCLC cells.  相似文献   

14.
The adenovirus E1A oncogene products stimulate DNA synthesis and cell proliferation but fail to transform primary baby rat kidney (BRK) cells because of the induction of p53-mediated programmed cell death (apoptosis). Overexpression of dominant mutant p53 (to abrogate wild-type p53 function) or introduction of apoptosis inhibitors, such as adenovirus E1B 19K or Bcl-2 oncoproteins, prevents E1A-induced apoptosis and permits transformation of BRK cells. The ability of activated Harvey-ras (H-ras) to cooperate with E1A to transform BRK cells suggests that H-ras is capable of overcoming the E1A-induced, p53-dependent apoptosis. We demonstrate here that activated H-ras was capable of suppressing apoptosis induced by E1A and wild-type p53. However, unlike Bcl-2 and the E1B 19K proteins, which completely block apoptosis but not p53-dependent growth arrest, H-ras expression permitted DNA synthesis and cell proliferation in the presence of high levels of wild-type p53. The mechanism by which H-ras regulates apoptosis and cell cycle progression is thereby strikingly different from that of the E1B 19K and Bcl-2 proteins. BRK cells transformed with H-ras and the temperature sensitive murine mutant p53(val 135), which lack E1A, underwent growth arrest at the permissive temperature for wild-type p53. p53-dependent growth arrest, however, could be relieved by E1A expression. Thus, H-ras alone was insufficient and cooperation of H-ras and E1A was required to override growth suppression by p53. Our data further suggest that two complementary growth signals from E1A plus H-ras can rescue cell death and thus permit transformation.  相似文献   

15.
Overexpression of wild-type p53 in p53-deficient leukemic cells induces apoptosis, which can be inhibited by hematopoietic survival factors. This suggests that p53 may contribute to survival factor dependence. To assess the role of wild-type p53 in mediating apoptosis following survival factor withdrawal, we interfered with endogenous p53 activity in interleukin-3 (IL-3)-dependent cells. Extended survival without IL-3 was conferred by recombinant retroviruses encoding either a full-length p53 mutant or a C-terminal p53 miniprotein, both of which can act as negative-dominant inhibitors of wild-type p53. On the other hand, excess wild-type p53 activity failed to elicit apoptosis as long as IL-3 was present. We propose that p53 is a positive, though not exclusive, mediator of survival factor dependence in hematopoietic cells.  相似文献   

16.
17.
Upon treatment with some DNA damaging agents, human H1299 tumor-derived cells expressing inducible versions of wild-type or mutant p53 with inactive transactivation domain I (p53Q22/S23) undergo apoptosis. In cells expressing either version of p53, caspase 2 activation is required for release of cytochrome c and cell death. Furthermore, silencing of PIDD (a factor previously shown to be required for caspase 2 activation) by siRNA suppresses apoptosis by both wild-type p53 and p53Q22/S23. Despite the finding that caspase 2 is essential for DNA damage-facilitated, p53-mediated apoptosis, induction of wild-type p53 (with or without DNA damage) resulted in a reduction of caspase 2 mRNA and protein levels. In this study we sought to provide a mechanism for the negative regulation of caspase 2 by p53 as well as provide insight as to why p53 may repress a key mediator of p53-dependent apoptosis. Mechanistically, we show that DNA binding and/or transactivation domains of p53 are crucial for mediating transrepression. Further, expression of p21 (in p53-null cells inducibly expressing p21) is sufficient to mediate repression of caspase 2. Deletion of p21 or E2F-1 not only abrogated repression of caspase 2, but also stimulated the expression of caspase 2 above basal levels, implicating the requirement for an intact p21/Rb/E2F pathway in the down-regulation of caspase 2. As this p53/p21-dependent repression of caspase 2 can occur in the absence of DNA damage, caspase 2 repression does not simply seem to be a consequence of the apoptotic process. Down-regulation of caspase 2 levels by p53 may help to determine cell fate by preventing cell death when unnecessary.  相似文献   

18.
19.
Utilizing a temperature sensitive p53 mutant (pLTRp53cGval135) which expresses mutant p53 at 37 degrees C and a wild-type like p53 at 32 degrees C, we transfected a human ovarian cancer cell line (SKOV3) which does not express endogenous p53. Among the different clones obtained, we selected three clones. Two were obtained from simultaneous transfection of p53 and neomycin resistance expression plasmids (SK23a and SK9), the other was obtained from transfection experiments utilizing the neomycin resistance gene only (SKN). Introduction of mutant p53 did not alter the morphology or growth characteristics of this ovarian cancer cell line. Upon shifting to the permissive temperature, a dramatic change in morphology and growth rate was observed in SK23a and SK9 cells that is associated with the presence of a wild-type like p53. SKN and SKOV3 cells maintained at 32 degrees C did not change morphology and only slightly reduced proliferation. Both SK23a and SK9 cells did not show evidence of apoptosis when measured up to 72 hours of maintenance at 32 degrees C. In contrast to what observed in other cell lines, SK23a and SK9 cells maintained at 32 degrees C were not blocked in G1, but they were accumulated in G2-M. This accumulation was transient and could be due either to a blockade or to a delay in the G2 progression. No down-regulation of c-myc was observed in p53 expressing clones when shifted to the permissive temperature. In these conditions gadd45 mRNA expression was highly stimulated in SK9 and SK23a cells but not in SKN cells. In both clones Gas1 mRNA was not detected either at 37 degrees C or 32 degrees C. This system represents a new and useful model for studying the effect of the absence of p53 (SKOV3 or SKN), presence of mutated p53 (SK23a and SK9 kept at 37 degrees C) or wild type p53 (SK23a and SK9 kept at 32 degrees C) on the mechanism of response of cancer cells to DNA damaging agents.  相似文献   

20.
Wild-type p53 triggers two distinct biological responses, cell cycle arrest and apoptosis. Several small DNA tumor viruses encode proteins that bind p53 and thus block the function of p53. This probably reflects the need of these viruses to prevent p53-induced cell cycle arrest and apoptosis to allow viral DNA replication. Unlike SV40 large T, polyoma virus large T does not bind p53, and it is still unclear how polyoma virus blocks p53 function. To address this question, we transfected polyoma virus middle T or small t alone or middle T and small t together into J3D mouse T-lymphoma cells carrying temperature-sensitive p53 (ts p53). Induction of wild-type p53 by temperature shift to 32 degrees C triggered both G1 cell cycle arrest and apoptosis in parental J3D-ts p53 cells. In contrast, J3D-ts p53 cells coexpressing middle T and small t showed only a weak G1 cell cycle arrest response after induction of wild-type p53 at 32 degrees C. Fluorescence-activated cell sorter analysis revealed that nearly half of the middle T-expressing cells, 30% of the small t-expressing cells, and a majority of the cells coexpressing middle T and small t were resistant to p53-induced apoptosis. The phosphatidylinositol 3-kinase inhibitor wortmannin partially abrogated the protective effect of middle T but not small t on p53-induced apoptosis, indicating that middle T prevents p53-induced apoptosis through the phosphatidylinositol 3-kinase signal transduction pathway. Our results thus establish a mechanism for polyoma virus-mediated inhibition of p53 function.  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