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 203 毫秒
1.
2.
TP53 is the most frequently mutated gene in cancers. Mutations lead to loss of p53 expression or expression of a mutant protein. Mutant p53 proteins commonly lose wild-type function, but can also acquire novel functions in promoting metastasis and chemoresistance. Previously, we uncovered a role for Rab-coupling protein (RCP) in mutant p53-dependent invasion. RCP promotes endosomal recycling and signalling of integrins and receptor tyrosine kinases. In a screen to identify novel RCP-interacting proteins, we discovered P-glycoprotein (P-gp). Thus, we hypothesised that mutant p53 could promote chemoresistance through RCP-dependent recycling of P-gp. The interaction between RCP and P-gp was verified endogenously and loss of RCP or mutant p53 rendered cells more sensitive to cisplatin and etoposide. In mutant p53 cells we detected an RCP-dependent delivery of P-gp to the plasma membrane upon drug treatment and decreased retention of P-gp substrates. A co-localisation of P-gp and RCP was seen in mutant p53 cells, but not in p53-null cells upon chemotherapeutic exposure. In conclusion, mutant p53 expression enhanced co-localisation of P-gp and RCP to allow for rapid delivery of P-gp to the plasma membrane and increased resistance to chemotherapeutics.Subject terms: Tumour-suppressor proteins, Preclinical research  相似文献   

3.
Human cytomegalovirus (HCMV) encodes several proteins that can modulate components of the cell cycle machinery. The UL123 gene product, IE1-72, binds the Rb-related, p107 protein and relieves its repression of E2F-responsive promoters; however, it is unable to induce quiescent cells to enter S phase in wild-type (p53(+/+)) cells. IE1-72 also induces p53 accumulation through an unknown mechanism. We present here evidence suggesting that IE1-72 may activate the p53 pathway by increasing the levels of p19(Arf) and by inducing the phosphorylation of p53 at Ser15. Phosphorylation of this residue by IE1-72 expression alone or HCMV infection is found to be dependent on the ataxia-telangiectasia mutated kinase. IE2-86 expression leads to p53 phosphorylation and may contribute to this phenotype in HCMV-infected cells. We also found that IE1-72 promotes p53 nuclear accumulation by abrogating p53 nuclear shuttling. These events result in the stimulation of p53 activity, leading to a p53- and p21-dependent inhibition of cell cycle progression from G(1) to S phase in cells transiently expressing IE1-72. Thus, like many of the small DNA tumor viruses, the first protein expressed upon HCMV infection activates a p53 response by the host cell.  相似文献   

4.
Using the differential display method combined with a cell line that carries a well-controlled expression system for wild-type p53, we isolated a p53-inducible gene, termed p53DINP1 (p53-dependent damage-inducible nuclear protein 1). Cell death induced by DNA double-strand breaks (DSBs), as well as Ser46 phosphorylation of p53 and induction of p53AIP1, were blocked when we inhibited expression of p53DINP1 by means of an antisense oligonucleotide. Overexpression of p53DINP1 and DNA damage by DSBs synergistically enhanced Ser46 phosphorylation of p53, induction of p53AIP1 expression, and apoptotic cell death. Furthermore, the protein complex interacting with p53DINP1 was shown to phosphorylate Ser46 of p53. Our results suggest that p53DINP1 may regulate p53-dependent apoptosis through phosphorylation of p53 at Ser46, serving as a cofactor for the putative p53-Ser46 kinase.  相似文献   

5.
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   

6.
The tumor suppressor p53 gene product is an essential component of the cytotoxic pathway triggered by DNA-damaging stimuli such as chemotherapeutic agents and ionizing radiation. We previously demonstrated that adenovirus-mediated wild-type p53 gene transfer could enhance the cytotoxic actions of chemotherapeutic drugs both in vitro and in vivo; however, the molecular mechanism of this chemosensitization is still unclear. Cyclin D1 is a major regulator of the progression of cells into the proliferative stage of the cell cycle. Here we show that infection with an adenovirus vector expressing the wild-type p53 gene (Ad-p53) caused an increase in cyclin D1 protein levels in human colorectal cancer cell lines DLD-1 and SW620; treatment with the anti-cancer drug adriamycin, however, down-regulated their cyclin D1 protein expression in a dose-dependent manner. The suppression of cyclin D1 expression following adriamycin treatment could be blocked by simultaneous Ad-p53 infection. Furthermore, DLD-1 and SW620 cells transfected with the cyclin D1 expression construct displayed increased sensitivity to adriamycin compared to that of the vector-transfected control. Our results suggest that ectopic wild-type p53 gene transfer results in increased cyclin D1 expression and, consequently, sensitizes human colorectal cancer cells to chemotherapeutic agents.  相似文献   

7.
Inoue T  Wu L  Stuart J  Maki CG 《FEBS letters》2005,579(22):4978-4984
Wild-type p53 accumulates in the nucleus following stress. Current models suggest this nuclear accumulation involves phosphorylation at p53 N-terminal sites, and inhibition of murine double minute (MDM)2-dependent nuclear export. We monitored the effects of stress on MDM2-dependent nuclear export of wild-type p53 and a mutant lacking N-terminal phosphorylation sites. Etoposide and ionizing radiation inhibited nuclear export of wild-type p53 and the phosphor-mutant to comparable extents, indicating nuclear export inhibition does not require N-terminal phosphorylation. Cytoplasmic p53 accumulated in the nucleus of transfected cells treated with the nuclear export-inhibitor leptomycin B (LMB). Interestingly, LMB caused less p53 nuclear accumulation than stress treatment, suggesting stress-induced nuclear accumulation of p53 does not result solely from inhibited nuclear export.  相似文献   

8.
A number of genotoxic chemicals and agents, such as benzo(a)pyrene and ultraviolet light, are able to induce nuclear accumulation of p53 protein. Usually, this response is transient and a consequence of stabilization of the wild-type p53 protein. After withdrawal of the exposure, the amount of p53 protein returns to a normal level within hours or a few days. We have studied the p53 response to the exposure of crocidolite asbestos in A-549 lung carcinoma cells using three different methods, i.e., p53 immunohistochemistry, Western blotting and metabolic labelling followed by p53 immunoprecipitation. With these techniques we demonstrate a dose-dependent p53 nuclear response to crocidolite exposure. The half-life of p53 protein in A-549 lung carcinoma cells cultured in serum-free media increased from 30 up to 80 min, and the protein reacted with a wild-type specific antibody suggesting that it was in a wild-type conformation. In situ 3′-end labelling of A-549 cells demonstrated a dose-dependent increase in apoptotic activity. Our data support the idea that increased apoptotic activity, induced by crocidolite, is mediated by p53. This revised version was published online in June 2006 with corrections to the Cover Date.  相似文献   

9.
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   

10.
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   

11.
Tumors expressing the ABL oncoproteins (BCR/ABL, TEL/ABL, v-ABL) can avoidapoptosis triggered by DNA damaging agents. The tumor suppressor protein p53 is animportant activator of apoptosis in normal cells; conversely its functional loss may causedrug resistance. The ABL oncoprotein - p53 paradigm represents the relationship between anoncogenic tyrosine kinase and a tumor suppressor gene. Here we show that BCR/ABLoncoproteins employ p53 to induce resistance to DNA damage in myeloid leukemia cells.Cells transformed by the ABL oncoproteins displayed accumulation of p53 upon DNAdamage. In contrast, only a modest increase of p53 expression followed by activation ofcaspase-3 were detected in normal cells expressing endogenous c-ABL. Phosphatidylinositol-3 kinase-like protein kinases (ATR and also ATM) -dependent phosphorylation of p53-Ser15residue was associated with the accumulation of p53, and stimulation of p21Waf-1 andGADD45, resulting in G2/M delay in BCR/ABL cells after genotoxic treatment. Inhibition ofp53 by siRNA or by the temperature-sensitive mutation reduced G2/M accumulation anddrug resistance of BCR/ABL cells. In conclusion, accumulation of the p53 proteincontributed to prolonged G2/M checkpoint activation and drug resistance in myeloid cellsexpressing the BCR/ABL oncoproteins.  相似文献   

12.
13.
14.
15.
《The Journal of cell biology》1995,128(6):1185-1196
The p53 tumor suppressor protein has been implicated as a mediator of programmed cell death (PCD). A series of nontransformed mammary epithelial cell (MEC) lines were used to correlate p53 function with activation of PCD. Treatment of MECs expressing mutant, inactive, or no p53 with DNA-damaging agents did not induce apoptosis. Upon introduction of temperature-sensitive p53 into HC11 cells, which lack wild-type (wt) p53, PCD was observed after mitomycin treatment at 32 degrees, when the ts p53 protein is in wt conformation. Thus, wt p53 mediates activation of PCD in response to mitomycin in HC11 cells. Treatment of the MCF10-A cells, which express wt p53, with various DNA- damaging agents led to nuclear accumulation of p53. Only mitomycin treatment led to an increase in the number of apoptotic nuclei. ErbB-2- transformed MCF10-A cells responded to mitomycin, cisplatin, and 5-Fl- uracil, suggesting that signaling from activated ErbB-2 enhances the cells ability to respond to DNA damage. A combination of high cell density and serum-free medium induces apoptosis in all MECs tested, irrespective of their p53 status. Under these conditions, EGF or insulin act as survival factors in preventing PCD. These data might elucidate some aspects of breast involution and tumorigenesis.  相似文献   

16.
Although the p53 tumor-suppressor gene product plays a critical role in apoptotic cell death induced by DNA-damaging chemotherapeutic agents, human glioma cells with functional p53 were more resistant to gamma-radiation than those with mutant p53. U-87 MG cells with wild-type p53 were resistant to gamma-radiation. U87-W E6 cells that lost functional p53, by the expression of type 16 human papillomavirus E6 oncoprotein, became susceptible to radiation-induced apoptosis. The formation of ceramide by acid sphingomyelinase (A-SMase), but not by neutral sphingomyelinase, was associated with p53-independent apoptosis. SR33557 (2-isopropyl-1-(4-[3-N-methyl-N-(3,4-dimethoxybphenethyl)amino]propyloxy)benzene-sulfonyl) indolizine, an inhibitor of A-SMase, suppressed radiation-induced apoptotic cell death. In contrast, radiation-induced A-SMase activation was blocked in glioma cells with endogenous functional p53. The expression of acid ceramidase was induced by gamma-radiation, and was more evident in cells with functional p53. N-oleoylethanolamine, which is known to inhibit ceramidase activity, unexpectedly downregulated acid ceramidase and accelerated radiation-induced apoptosis in U87-W E6 cells. Moreover, cells with functional p53 could be sensitized to gamma-radiation by N-oleoylethanolamine, which suppressed radiation-induced acid ceramidase expression and then enhanced ceramide formation. Sensitization to gamma-radiation was also observed in U87-MG cells depleted of functional p53 by retroviral expression of small interfering RNA. These results indicate that ceramide may function as a mediator of p53-independent apoptosis in human glioma cells in response to gamma-radiation, and suggest that p53-dependent expression of acid ceramidase and blockage of A-SMase activation play pivotal roles in protection from gamma-radiation of cells with endogenous functional p53.  相似文献   

17.
18.
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   

19.
20.
Wild-type p53 protein is abnormally sequestered in the cytoplasm of a subset of primary human tumors including neuroblastomas (NB) (U. M. Moll, M. LaQuaglia, J. Benard, and G. Riou, Proc. Natl. Acad. Sci. USA 92:4407-4411, 1995; U. M. Moll, G. Riou, and A. J. Levine, Proc. Natl. Acad. Sci.USA 89:7262-7266, 1992). This may represent a nonmutational mechanism for abrogating p53 tumor suppressor function. To test this hypothesis, we established the first available in vitro model that accurately reflects the wild-type p53 sequestration found in NB tumors. We characterized a series of human NB cell lines that overexpress wild-type p53 and show that p53 is preferentially localized to discrete cytoplasmic structures, with no detectable nuclear p53. These cell lines, when challenged with a variety of DNA strand-breaking agents, all exhibit impaired p53-mediated G1 arrest. Induction analysis of p53 and p53-responsive genes show that this impairment is due to suppression of nuclear p53 accumulation. Thus, this naturally occurring translocation defect compromises the suppressor function of p53 and likely plays a role in the tumorigenesis of these tumors previously thought to be unaffected by p53 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