肿瘤发生及治疗过程中p53蛋白线粒体迁移及相关细胞凋亡途径的研究
详细信息    本馆镜像全文|  推荐本文 |  |   获取CNKI官网全文
摘要
p53蛋白作为一种重要的肿瘤抑制因子,在细胞或病毒性致癌基因(Oncogene)被表达的时候会被自动激活并启动细胞凋亡途径杀死突变细胞,是生物体对抗癌变的主要机制。p53蛋白在被外界条件激活后,通过转录途径(transcriptional-dependent pathway)和非转录途径(transcriptional-independent pathway)共同诱导细胞凋亡的发生。p53蛋白的非转录活性主要来自于线粒体内的p53蛋白。虽然目前对线粒体p53蛋白功能方面仍有很多争议,但p53蛋白在激活后的线粒体迁移(p53 mitochondrial translocation)已经有大量数据支持并被人们所接受了。尽管如此,目前关于p53线粒体迁移机制方面的认识还很少,且存在很多争议性的观点。本文工作的重点就是研究肿瘤发生过程中以及癌症细胞中的p53蛋白线粒体迁移机制,详细考察了线粒体生理功能对皮肤肿瘤发生过程(skin carcinogenesis)中p53蛋白线粒体迁移的影响,并研究了突变p53蛋白在淋巴瘤细胞(lymphoma cells)中的线粒体迁移状况以及相关的凋亡功能。本文首次揭示了线粒体渗透性转换(mitochondrial permeability transition)以及线粒体解偶联效应(mitochondrial uncoupling effect)对p53蛋白线粒体迁移的调控作用,证明了突变p53蛋白线粒体迁移在癌症细胞中的存在及潜在的凋亡诱导能力,这些实验结果有助于更好的理解p53蛋白的线粒体迁移机制以及p53非转录凋亡途径,同时也为进一步设计以线粒体p53为靶标的癌症预防与治疗手段提供理论基础。
The tumor suppresser p53 is known to trigger apoptosis in response to DNA damage, oncogene activation, or certain chemotherapeutic drugs. Potentially, as a defending mechanism against carcinogenesis, p53 induces apoptosis via transcriptional-dependent and transcriptional-independent pathways, two fundamentally different, but synergistic mechanisms of action. In addition to its transcriptional activation, it has been shown that a fraction of p53 translocates into mitochondria at the very early stage of apoptosis which, in several conditions, is necessary for the activation of p53-apoptosis network. However, the mitochondrial events that affect p53 translocation and associated apoptosis are still unclear.
     Herein, we studied whether p53 mitochondrial translocation and subsequent apoptosis were affected by mitochondrial functional alternation including mitochondrial uncoupling and blocking permeability transition pore during skin tumor promotion. Our studies revealed several interesting facts: (1) TPA induced p53 mitochondrial translocation and associated mitochondrial dysfunction and apoptosis in JB6 P+ cells. (2) Mitochondrial uncoupling and blocking permeability transition pore both suppressed TPA-induced p53 mitochondrial translocation in JB6 P+ cells. (3) Mitochondrial uncoupling and blocking permeability transition pore both prevented mitochondrial dysfunction and apoptosis associated with p53 mitochondrial translocation in JB6 P+ cells. (4) Blocking permeability transition pore reduced the mitochondrial p53 level in DBA2/J mouse skin epidermal tissues after DMBA/TPA treatment. (5) UCP2-knockdown JB6 P+ cells showed enhanced p53 mitochondrial translocation and decreased colony formation in response to TPA treatment.
     All these results suggested that mitochondrial functional alternation including mitochondrial uncoupling and permeability transition could modulate p53 mitochondrial translocation and further affect downstream mitochondrial dysfunction and apoptosis pathway during early skin tumor promotion. More over, these effects existed both in vivo and in vitro, and were important for the tumor formation process.
     The above mentioned p53 apoptotic activities belong to wt p53 and have been well documented for years. Nevertheless, as a pivotal tumor suppresser, p53 is the most commonly mutated gene in human cancers. Accumulated studies reveal that more than 50% of human cancers contain mutations in p53 gene. These mutations strongly select for p53 proteins that fail to bind to DNA in a sequence-specific fashion and thus block its transcriptional activities. However, the DNA binding activity is not necessary for p53 to exert its transcriptional-independent activities and a number of studies have suggested mitochondrial localization of mutant p53 in cancer cells. These results indicate a potential role of mutant p53 in the mitochondria of cancer cells. Since our above results have revealed a potential relationship between mitochondrial metabolism status and p53 mitochondrial translocation, we chose several lymphoma cell line with different p53 status (wt, mutant, or null) to further investigated whether mutant p53 translocates into mitochondria during cancer therapy, whether mutant p53 affects mitochondrial functions including complex activities and apoptosis, and whether restoring mutant p53 by ellipticine enforces mutant p53 to affect mitochondrial functions.. Our results showed that: (1) The expression levels of both wt and mutant p53 were increased in lymphoma cells upon doxorubicin treatment. (2) Both wt and mutant p53 translocated into mitochondria upon doxorubicin treatment. (3) Mutant p53-bearing DHL-4 cells were more resistant to doxorubicin treatment than wt p53-containing DoHH2 cells. (4) Decreases in mitochondrial Complex I and Complex II activities were associated with translocation of wt but not mutant p53. (5) A small molecule, ellipticine, restored the activity of mutant p53 which further affect mitochondrial function and apoptosis.
     These results from lymphoma cells demonstrated that during cancer therapy, both wt and mutant p53 could translocate into mitochondria while only wt p53 mitochondrial translocation could further induce mitochondrial dysfunction and apoptosis. This means, in addition to the blocking effect on transcriptional-dependent pathway, mutations in p53 can also block the transcriptional-independent pathway, and very interestingly, status of p53 doesn’t affect its mitochondrial translocation. Even though the mechanism this phenomenon is still unclear, these blocking effect might result from conformational alternation caused by mutation, since a small molecule, ellipticine, can interact with mutant p53 and restore its apoptotic activity in response to doxorubicin treatment.
     Taken together, our studies provide a deep insight into the mechanism of p53 mitochondrial translocation and associated alternation in apoptotic network. Our studies concerned both cancer cells and normal epithermal cells during early tumor promotion. Our results and conclusions may provide novel information regarding the role of p53 transcriptional-independent pathway in carcinogenesis, which may contribute to the understanding of the relationship between p53 status and cancer chemoresistance from a new angle, helping to design a better strategy for p53-mediated therapy.
引文
[1] Eliyahu D, Michalovitz D, Eliyahu S, et al. p53-Oncogene or anti-oncogene.[J] Oncogenes in Cancer Diagnosis, 1990, 39:125-134.
    [2] Lane DP and Benchimol S. p53: Oncogene or anti-oncogene?[J] Genes and Development, 1990, 4:1-8.
    [3] Soussi T and Béroud C. Assessing TP53 status in human tumours to evaluate clinical outcome.[J] Nat Rev Cancer, 2001, 1(3):233-240.
    [4] Michalovitz D, Halevy O and Oren M. p53 mtation-gain or losses.[J] J Cell Biochem, 1991, 45:22-29.
    [5] Milner J. Flexibility: the key to p53 function?[J] Trends Biochem Sci, 1995, 20(2):49-51.
    [6] Dittmer D, Pati S, Zambetti G, et al. Gain of function mutation in p53.[J] Nat Genet, 1993, 4(1):42-46.
    [7] Halevy O, Michalovitz D and Oren M. Different tumor-derived p53 mutants exhibit distinct biological activities.[J] Science, 1990, 250(4977):113-116.
    [8] Chang C, Simmons DT, Martin MA, et al. Identification and partial characterization of new antigens from simian virus 40-transformed mouse cells.[J] J Virol, 1979, 31:463-471.
    [9] Kress M, May E, Cassingena R, et al. Simian virus 40-transformed cells express new species of proteins precipitable by anti-simian virus 40 serum.[J] J Virol, 1979, 31:472-483.
    [10] Lane DP and Crawford LV. T antigen is bound to a host protein in SV40-transformed cells.[J] Nature, 1979, 278:261-263.
    [11] Linzer DI and Levine AJ. Characterization of a 54 K dalton cellular SV40 tumor antigen resent in SV40-transformed cells and in infected embryonal carcinoma cells.[J] Cell, 1979, 1:43-52.
    [12] Melero JA, Stitt DT, Mangel WF, et al. Identification of new polypeptide species (48-55K) immunoprecipitable by antiserum to purified large T antigen and present in simian virus 40-infected and transformed cells.[J] J Virol, 1979, 93:466-480.
    [13] De Leo AB, Jay G, Appella E, et al. Detection of a transformation-related antigen in chemically induced sarcomas and other transformed cells of the mouse.[J] Proc Natl Acad Sci USA, 1979, 76:2420-2424.
    [14] Rotter V, Witte ON, Coffman R, et al. Abelson murine leukemia virus-induced tumors elicit antibodies against a host cell protein, p53.[J] J Virol, 1980, 36:547-555.
    [15] Crawford LV, Pim DC and Bulbrook RD. Detection of antibodies against the cellular protein p53 in sera from patients with breast cancer.[J] Int J Cancer, 1982, 30:403-408.
    [16] Caron de Fromentel C, May-Levin F, Mouriesse H, et al. Presence of circulating antibodies anainst cellular protein p53 in a notable proportion of children with B-cell lymphoma.[J] Int J Cancer, 1987, 39:185-189.
    [17] Reich NC and Levine AJ. Growth regulation of a cellular tumour antigen, p53, in nontransformed cells.[J] Nature, 1984, 308:199-201.
    [18] Milner J and Mccornick F. Lymphocyte stimulation: concanavalin A induces the expression of a 53k protein.[J] Cell Biol Int Rep, 1980, 4:663-667.
    [19] Mercer WE, Avignolo C and Baserga R. Role of the p53 protein in cell proliferation as studied by microinjection of monoclonal antibodies.[J] Mol Cell Biol, 1984, 4:276-281.
    [20] Calabretta B, Kaczmarek LL, Selleri L, et al. Growth-dependent expression of human Mr 53,000 tumor antigen messenger RNA in normal and neoplastic cells.[J] Cancer Res, 1986, 46:5738-5742.
    [21] Mercer WE, Nelson D, DeLeo AB, et al. Microinjection of monoclonal antibody to protein p53 inhibits serum-induced DNA synthesis in 3T3 cells.[J] Proc Natl Acad Sci USA, 1982, 79:6309-6312.
    [22] Steinmeyer K, Maacke H and Deppert W. Cell cycle control by p53 in normal (3T3) and chemically transformed (meth-A) mouse cells. I. Regulation of p53 expression.[J] Oncogene, 1990, 5:1691-1699.
    [23] Deppert W, Buschhausendenker G, Patschinsky T, et al. Cell cycle control of p53 in normal (3T3) and chemically transformed (Meth-A) mouse cells. II. Requirement for cell cycle progression.[J] Oncogene, 1990, 5:1701-1706.
    [24] Shohat O, Greenberg M, Reisman D, et al. Inhibition of cell growth mediated by plasmids encoding p53 anti-sense.[J] Oncogene, 1987, 1:277-283.
    [25] Eliyahu D, Raz A, Gruss P, et al. Participation of p53 cellular tumour antigen in transformation of normal embryonic cells.[J] Nature, 1984, 312:646-649.
    [26] Parada LF, Land H, Weinberg RA, et al. Cooperation between gene encoding p53 tumour antigen and ras in cellular transformation.[J] Nature, 1984, 312:649-651.
    [27] Jenkins JR, Rudge K, Chumakov P, et al. The cellular oncogene p53 can be activated by mutagenesis.[J] Nature, 1985, 317:816-818.
    [28] Jenkins JR, Rudge K and Currie GA. Cellular immortalization by a cDNA clone encodinig the transformation-associated phosphoprotein p53.[J] Nature, 1984, 312:651-654.
    [29] Mowat M, Cheng A, Kimura N, et al. Rearrangements of the cellular p53 gene in erythroleukaemic cells transformed by Friend virus.[J] Nature, 1985, 314:633-636.
    [30] Munroe DG, Rovinski B, Bernstein A, et al. Loss of highly conserved domain on p53 as aresult of gene deletion during friend virus-induced erythroleukemia.[J] Oncogene, 1988, 2:621-624.
    [31] Eliyahu D, Michalovitz D, Eliyahu S, et al. Wild-Type p53 can inhibit oncogene-mediated focus formation.[J] Proc Natl Acad Sci USA, 1989, 86:8763-8767.
    [32] Finlay CA, Hinds PW and Levine AJ. The p53 proto-oncogene can act as a suppressor of transformation.[J] Cell, 1989, 57:1083-1093.
    [33] Hupp TR, Sparks A and Lane DP. Small peptides activate the latent sequence-specific DNA binding function of p53.[J] Cell, 1995, 83(2):237-245.
    [34] Vogelstein B, Fearon ER, Hamilton SR, et al. Genetic alterations during colorectal-tumor development.[J] N Engl J Med, 1988, 319:525-532.
    [35] Baker SJ, Fearon ER, Nigro J, et al. Chromosome 17 deletions and p53 gene mutations in colorectal carcinomas.[J] Science, 1989, 244:217-221.
    [36] Takahashi T, Nau MM, Chibal I, et al. p53-a frequent target for genetic abnormalities in lung cancer.[J] Science, 1989, 246:491-494.
    [37] Milner J and Medcalf EA. Cotranslation of activated mutant p53 with wild type drives the wild-type p53 protein into the mutant conformation.[J] Cell, 1991, 65:765-774.
    [38] Lane DP. Cancer. p53. guardian of the genome.[J] Nature, 1992, 358(6381):15-16.
    [39] Vousden KH. p53: death star.[J] Cell, 2000, 103(5):691-694.
    [40] Harris CC. p53: at the crossroads of molecular carcinogenesis and risk assessment.[J] Science, 1993, 262(5142):1980-1981.
    [41] Vogelstein B, Lane D and Levine AJ. Surfing the p53 network.[J] Nature, 2000, 408:307-310.
    [42] Giaccia AJ and Kastan MB. The complexity of p53 modulation: emerging patterns from divergent signals.[J] Genes Dev, 1998, 12:2973-2983.
    [43] Gudkov AV and Komarova EA. The role of p53 in determining sentitivity to radiotherapy.[J] Nat Rev Cancer, 2003, 3:117-129.
    [44] Oren M. Decision making by p53: life, death and cancer.[J] Cell Death, 2003, 10:431-442.
    [45] Feng Z, Zhang H, Levine AJ, et al. The coordinate regulation of the p53 and mTOR pathways in cells.[J] Proc Natl Acad Sci USA, 2005, 102:8204-8209.
    [46] Jones RG, Plas DR, Kubek S, et al. AMP-activated protein kinase induces a p53-dependent metabolic checkpoint.[J] Mol Cell, 2005, 18:283-293.
    [47] Soussi T. The p53 pathway and human cancer.[J] Br J Surg, 2005, 92:1331-1332.
    [48] Harris SL and Levine AJ. The p53 pathway: positive and negative feedback loops.[J] Oncogene, 2005, 24:2899-2908.
    [49] Leng RP, Lin Y, Ma W, et al. Pirh2, a p53-induced ubiquitin-protein ligase, promotes p53 degradation.[J] Cell, 2003, 112:779-791.
    [50] Dornan D, Wertz I, Shimizu H, et al. The ubiqutin ligase COP1 is a critical negative regulator of p53.[J] Nature, 2004, 429:86-92.
    [51] Stommel JM and Wahl GM. Accelerated MDM2 auto-degradation induced by DNA-damage kinases is required for p53 activation.[J] EMBO J, 2004, 23:1547-1556.
    [52] Appella E and Anderson CW. Post-translational modifications and activation of p53 by genotoxic stresses.[J] Eur J Biochem, 2001, 268:2764-2772.
    [53] Xirodimas DP, Saville MK, Bourdon JC, et al. Mdm2-mediated NEDD8 conjugation of p53 inhibits its transcriptional activity.[J] Cell, 2004, 118:83-97
    [54] Brooks CL and Gu W. Ubiquitination, phosphorylation and acetylation: the molecular basis for p53 regulation.[J] Cruu Opin Cell Biol, 2003, 15:164-171.
    [55] Xu Y. Regulation of p53 responses by post-translational modifications.[J] Cell Death Differ, 2003, 10:400-403.
    [56] Vassilev LT, Vu BT, Graves B, et al. In vivo activation of the p53 pathway by small-molecule antagonists of MDM2.[J] Science, 2004, 303:844-848.
    [57] Mckinney K, Mattia M, Gottifredi V, et al. p53 linear diffusion along DNA requires its C terminus.[J] Mol Cell, 2004, 16:413-424.
    [58] Mckinney K and Prives C. Regulation of p53 DNA binding[M]. 25 years of p53 research. Dordrecht: Springer pp:27-52.
    [59] Lowe SW and Sherr CJ. Tumor suppression by Ink4a-Arf: progress and puzzles.[J] Curr Opin Genet Dev, 2003, 13:77-83.
    [60] Hemann MT and Lowe S. p53 links tumor development th cancer therapy.[M] 25 years of p53 research. Dordrecht: Springer. Pp:339-352.
    [61] Lev Bar-Or R, Maya R, Segel LA, et al. Generation of oscillations by the p53-Mdm2 feedback loop: a theoretical and experimental study.[J] Proc Natl Acad Sci USA, 2000, 97:11250-11255.
    [62] Lahav G, Rosenfeld N, Sigal A, et al. Dynamics of the p53-Mdm2 feedback loop in individual cells.[J] Nat Genet, 2004, 36:147-150.
    [63] Momand J, Zambetti GP, Olson DC, et al. The mdm-2 oncogene product forms a complex with the p53 protein and inhibits p53-mediated transactivation.[J] Cell, 1992, 69:1237-1245.
    [64] Oliner JD, Kinzler KW, Meltzer PS, et al. Amplification of a gene encoding a p53-associated protein in human sarcomas.[J] Nature, 1992, 358:80-83.
    [65] Montes de Oca Luna R, Wagner DS and Lozano G. Rescue of early embryonic lethality in mdm2-deficient mice by deletion of p53.[J] Nature, 1995, 378:203-206.
    [66] Parant J, Chavez-Reyes A, Little NA, et al. Rescue of embryonic lethality in Mdm4-null mice by loss of Trp53 suggests a nonoverlapping pathway with MDM2 to regulate p53.[J] Nat Genet, 2001, 29:92-95.
    [67] Tao W and Levine AJ. Nucleocytoplasmic shuttling of oncoprotein Hdm2 is required for Hdm2-mediated degradation of p53.[J] Proc Natl Acad Sci USA, 1999, 96: 3077-3080.
    [68] Boyd SD, Tsai KY and Jacks T. An intact HDM2 RING-finger domain is required for nuclear exclusion of p53.[J] Nat Cell Biol, 2000, 2:563-568.
    [69] Mayo LD and Donner DN. A phosphatidylinasitol 3-kinase/Akt pathway promotes translocation of Mdm2 from the cytoplasm to the nucleus.[J] Proc Natl Acad Sci USA, 2001, 98:11598-11603.
    [70] Zhou BP, Liao Y, Xia W, et al. HER-2/neu induces p53 ubiquitination via Akt-mediated MDM2 phosphorylation.[J] Nat Cell Biol, 2001, 3:973-982.
    [71] Okamoto K, Li H, Jensen MR, et al. Cyclin G recruits PP2A to dephosphorylate Mdm2.[J] Mol Cell, 2002, 9:761-771.
    [72] Levine AJ, Feng Z, Mak T, et al. Coordination and communication between the p53 and IGF-1-AKT-TOR signal transduction pathway.[J] Genes Dev, 2006, 20:267-275.
    [73] Chen J, Ruan H, Ng SM, et al. Loss of function of def selectively up regulates Delta-113p53 expression to arrest expansion growth of digestive organs in zebrafish.[J] Genes Dev, 2005, 19:2900-2911.
    [74] el-Deiry WS, Kern SE, Pietenpol JA, et al. Definition of a consensus binding site for p53.[J] Nat Genet, 1992, 1:45-49.
    [75] Hoh J, Jin S, Parrado T, et al. The p53MH algorithm and its application in detectingp53-responsive genes.[J] Proc Natl Acad Sci USA, 2002, 99:8467-8472.
    [76] Feng Z, Jin S, Zupnick A, et al. p53 tumor suppressor protein regulates the levels of huntingtin gene expression.[J] Oncogene, 2005, 24:1-7.
    [77] Zhao R, Gish K, Murphy M, et al. The transcriptional program following p53 activation.[J] Cold Spring Herb Symp Quant Biol, 2000, 65:475-482.
    [78] Moll UM, Wolff S, Speidel D, et al. Transcriptional-independent pro-apoptotic functions of p53.[J] Curr Opin Cell Biol, 2005, 17:631-636.
    [79] Chipuk JE, Kuwana T, Bouchier-Hayers L, et al. Direct activation of Bax by p53 mediates mitochondrial membrane permeabilization and apoptosis.[J] Science, 2004, 303:1010-1014.
    [80] Donehower LA. Does p53 affect organismal aging?[J] J Cell Physiol, 2002, 192:23-33.
    [81] Budanov AV, Sablina AA, Feinstein E, et al. Regeneration of peroxiredoxin by p53-regulated sestrins, homologs of bacterial AhpD.[J] Science, 2004, 304:596-600.
    [82] Buckbinder L, Talbott R, Velasco-Miguel S, et al. Induction of the growth inhibitor IGF-binding protein 3 by p53.[J] Nature, 1995, 377:646-649.
    [83] Kunz C, Pebler S, Otte J, et al. Differential regulation of plasminogen activator and inhibitor gene transcription by the tumor suppressor p53.[J] Nucl Acids Res, 1995, 23:3710-3717.
    [84] Zou Z, Gao C, Nagaich AK, et al. p53 regulates the expression of the tumor suppressor gene maspin.[J] J Biol Chem, 2000, 275:6051-6054.
    [85] Dameron KM, Volpert OV, Tainsky MA, et al. Control of angiogenesis in fibroblasts by p53 regulation of thrombospondin-1.[J] Science, 1994, 265:1582-1584.
    [86] Amzallag N, Passer BJ, Allanic D, et al. TSAP6 facilitates the secretion of translationally controlled tumor protein/histamine-releasing factor via a nonclassical pathway.[J] J Biol Chem, 2004, 279:46104-46112.
    [87] Deng C, Zhang P, Harper JW, et al. Mice lacking p21CIP1/WAF1 undergo normal development, but are defective in G1 checkpoint control.[J] Cell, 1995, 82:675-684.
    [88] Brugarolas J, Chandrasekaran C, Gordon JI, et al. Radiation-induced cell cycle arrest compromised by p21 deficiency.[J] Nature, 1995, 377: 552-557.
    [89] Vaziri C, Saxena S, Jeon Y, et al. A p53-dependent checkpoint pathway prevents rereplication.[J] Mol Cell, 2003, 11:997-1008.
    [90] Wang XJ, Greenhalgh DA, Jiang A, et al. Analysis of centrosome abnormalities andangiogenesis in epidermaltargeted p53172 H mutant and p53-knockout mice after chemical carcinogenesis: evidence for a gain of function.[J] Mol Carcinog, 1998, 23:185-192.
    [91] Passer BJ, Nancy-Portebois V, Amzallag N, et al. The p53-inducible TSAP6 gene product regulates apoptosis and the cell cycle and interacts with Nix and the Myt1 kinase.[J] Proc Natl Acad Sci USA, 2003, 100:2284-2289.
    [92] Caelles C, Helmberg A and Karin M. p53-dependent apoptosis in the absence of transcriptional activation of p53-target genes.[J] Nature, 1994, 370:220-223.
    [93] Haupt Y, Rowan S, Shaulian E, et al. Induction of apoptosis in HeLa cells by trans-activation-deficient p53.[J] Genes Dev, 1995, 9:2170-2183.
    [94] Wagner AJ, Kokontis JM and Hay N. Myc-mediated apoptosis requires wild-type p53 in a manner independent of cell cycle arrest and the ability of p53 to induce p21waf1/cip1.[J] Genes Dev, 1994, 8:2817-2830.
    [95] Chipuk JE and Green DR. p53’s believe it or not: lessons on transcription-independent death.[J] J Clin Immunol, 2003, 23:355-361.
    [96] Bennett M, Macdonald K, Chan SW, et al. Cell surface trafficking of Fas: a rapid mechanism of p53-mediated apoptosis.[J] Science, 1998, 282:290-293.
    [97] Leu JI, Dumont P, Hafey M, et al. Mitochondrial p53 activates Bak and causes disruption of a Bak-Mcl1 complex.[J] Nat Cell Biol, 2004, 6:443-450.
    [98] Mihara M, Erater S, Zaika A, et al. p53 has a direct apoptogenic role at the mitochondria.[J] Mol Cell, 2003, 11:577-590.
    [99] Dumont P, Leu JI, Della Pietra AC 3rd, et al. The codon 72 polymorphic variants of p53 have markedly different apoptotic potential.[J] Nat Genet, 2003, 33:357-365.
    [100] Marchenko ND, Zaika A and Moll UM. Death signal-induced localization of p53 protein to mitochondria. A potential role in apoptotic signaling.[J] J Biol Chem, 2000, 275:16202-16212.
    [101] Sansome C, Zaika A, Marchenko ND, et al. Hypoxia death stimulus induces translocation of p53 protein to mitochondria. Detection by immunofluorescence on whole cells.[J] FEBS Lett, 2001, 488:110-115.
    [102] Essmann F, Pohlmann S, Gillissen B, et al. Irradiation-induced translocation of p53 to mitochondria in the absence of apoptosis.[J] J Biol Chem, 2005, 280:37169-37177.
    [103] Wolff S, Erster S, Palacios G, et al. p53’s mitochondrial translocation and MOMP action isindependent of Puma and Bax and severely disrupts mitochondrial membrane integrity.[J] Cell Res, 2008, 18:733-744.
    [104] Pietsch EC, Perchiniak E, Canutescu AA, et al. Oligomerization of BAK by p53 utilizes conserved residues of the p53 DNA binding domain.[J] J Biol Chem, 2008, 283:21294-21304.
    [105] Sykes SM, Stanek TJ, Frank A, et al. Acetylation of the DNA binding domain regulates transcription-independent apoptosis by p53.[J] J Biol Chem, 2009, 284:20197-20205.
    [106] Sot B, Freund SM and Fersht AR. Comparative biophysical characterization of p53 with the pro-apoptotic BAK and the anti-apoptotic BCL-xL.[J] J Biol Chem, 2007, 282:29193-29200.
    [107] Tomita Y, Marchenko N, Erster S, et al. WT p53, but not tumor-derived mutants, bind to Bcl2 via the DNA binding domain and induce mitochondrial pemeabilization.[J] J Biol Chem, 2006, 281:8600-8606.
    [108] Petros AM, Gunasekera A, Xu N, et al. Defining the p53 DNA-binding domain/Bcl-x(L)-binding interface usinig NMR.[J] FEBS Lett, 2004, 559:171-174.
    [109] Akhtar RS, Geng Y, Klocke BJ, et al. Neural precursor cells possess multiple p53-dependent apoptotic pathways.[J] Cell Death Differ, 2006, 13:1727-1739.
    [110] Geng Y, Akhtar RS, Shacka JJ, et al. p53 transcription-dependent and–independent regulation of cerebellar neural precursor cell apoptosis.[J] J Neuropathol Exp Neurol, 2007, 66-74.
    [111] Speidel D, Helmbold H and Deppert W. Dissection of transcriptional and non-transcriptional p53 activities in the response to genotoxic stress.[J] Oncogene, 2006, 25:940-953.
    [112] Todorovicc V, Chen CC, Hay N, et al. The matrix protein CCN1 (CYR61) induces apoptosis in fibroblasts.[J] J Cell Biol, 2005, 171:559-568.
    [113] Tan J, Zhuang L, Leong HS, et al. Pharmacologic modulation of glycogen synthase kinase-3beta promotes p53-dependent apoptosis through a direct Bax-mediated mitochondrial pathway in colorectal cancer cells.[J] Cancer Res, 2005, 65:9012-9020.
    [114] Waster PK and Ollinger KM. Redox-dependent translocation of p53 to mitochondria or nucleus in human melanocytes after UVA- and UVB-induced apoptosis.[J] J Invest Dermatol, 2009, 129:1769-1781.
    [115] Raffo AJ, Kim AL and Fine RL. Formation of nuclear Bax/p53 comnplexex is associated with chemotherapy induced apoptosis.[J] Oncogene, 2000, 19:6216-6228.
    [116] Chipuk JE, Maurer U, Green DR, et al. Pharmacologic activation of p53 elicitsBax-dependent apoptosis in the absence of transcription.[J] Cancer Cell, 2003, 4:371-381.
    [117] Jiang P, Du W, Heese K, et al. The Bad guy cooperates with good cop p53: Bad is transcriptionally up-regulated by p53 and forms a Bad/p53 complex at the mitochondria to induce apoptosis.[J] Mol Cell Biol, 2006, 26:9071-9082.
    [118] Xu H, Tai J, Ye H, et al. The N-terminal domain of tumor suppressor p53 is involved in the molecular interaction with the anti-apoptotic protein Bcl-XL.[J] Biochem Biophys Res Commun, 2006, 341:938-944.
    [119] Deng X, Gao F, Flagg T, et al. Bcl2’s flexible loop domain regulates p53 binding and survival.[J] Mol Cell Biol, 2006, 26:4421-4434.
    [120] Fletcher JI, Meusburger S, Hawkins CJ, et al. Apoptosis is triggered when prosurvival Bcl-2 proteins cannot restrain Bax.[J] Proc Natl Acad Sci USA, 2008, 105:18081-18087.
    [121] Jiang M and Milner J. Bcl-2 constitutively suppresses p53-dependent apoptosis in colorectal cancer cells.[J] Genes Dev, 2003, 17:832-837.
    [122] Chipuk JE, Bouchier-Hayes L, Kuwana T, et al. PUMA couples the nuclear and cytoplasmic proapoptotic function of p53.[J] Science, 2005, 309:1732-1735.
    [123] Palacios G and Moll UM. Mitochondrially targeted wild-type p53 suppresses growth of mutant p53 lymphomas in vivo.[J] Oncogene, 2006,25:6133-6139.
    [124] Talos F, Petrenko O, Mena P, et al. Mitochondrially targetd p53 has tumor suppressor activities in vivo.[J] Cancer Res, 2005, 65:9971-9981.
    [125] Ferecatu I, Rincheval V, Mignotte B, et al. Tickets for p53 journey among organelles.[J] Front Biosci, 2009, 14:4214-4228.
    [126] You H, Yamamoto K and Mak TW. Regulation of transactivation-independent proapoptotic activity of p53 by FOXO3a.[J] Proc Natl Acad Sci USA, 2006, 103:9051-9056.
    [127] Li M, Brooks CL, Wu-Baer F, et al. Mono- versus polyubiquitination: differential control of p53 fate by Mdm2.[J] Science, 2003, 302:1972-1975.
    [128] Marine JC, Francoz S, Maetens M, et al. Keeping p53 in check: essential and synergistic functions of Mdm2 and Mdm4.[J] Cell Death Differ, 2006, 13:927-934.
    [129] Marchenko ND, Wolff S, Erster S, et al. Monoubiquitylation promotes mitochondrial p53 translocation.[J] EMBO J, 2007, 26:923-934.
    [130] Dhar SK and St Clair DK. Nucleophosmin blocks mitochondrial localization of p53 andapoptosis.[J] J Biol Chem, 2009, 284:16409-16418.
    [131] Pietsch EC, Leu JI, Frank A, et al. The tetramerization domain of p53 is required for efficient BAK oligomerization.[J] Cancer Biol Ther, 2007, 6:1576-1583.
    [132] Moll UM, LaQuaglia M, Bénard J, et al. Wild-type p53 protein undergoes cytoplasmic sequestration in undifferentiated neuroblastomas but not in differentiated tumors.[J] Proc Natl Acad Sci USA, 1995, 92:4407-4411.
    [133] Moll UM, Ostermeyer AG, Haladay R, et al. Cytoplasmic sequestration of wild-type p53 protein impairs the G1 checkpoint after DNA damage.[J] Mol Cell Biol, 1996, 16:1126-1137.
    [134] Moll UM, Riou G and Levine AJ. Two distinct mechanisms alter p53 in breast cancer: mutation and nuclear exclusion.[J] Proc Natl Acad Sci USA, 1992, 89:7262-7266.
    [135] Ueda H, Ullrich SJ, Gangemi JD, et al. Functional inactivation but not structural mutation of p53 causes liver cancer.[J] Nat Genet, 1995, 9:41-47.
    [136] Mahyar-Roemer M, Fritzsche C, Wagner S, et al. Mitochondrial p53 levels parallel total p53 levels independent of stress response in human colorectal carcinoma and glioblastoma cells.[J] Oncogene, 2004, 23:6226-6236.
    [137] Ferecatu I, Bergeaud M, Rodríguez-Enfedaque A, et al. Mitochondrial localization of the low level p53 protein in proliferative cells.[J] Biochem Biophys Res Commun, 2009, 387:772-777.
    [138] Tasdemir E, Maiuri MC, Galluzzi L, et al. Regulation of autophagy by cytoplasmic p53.[J] Nat Cell Biol, 2008, 10:676-687.
    [139] Green DR and Kroemer G. Cytoplasmic functions of the tumour suppressor p53.[J] Nature, 2009, 458:1127-1130.
    [140] Morselli E, Tasdemir E, Maiuri MC, et al. Mutant p53 protein localized in the cytoplasm inhibits autophagy.[J] Cell Cycle, 2008, 7:3056-3061.
    [141] Nemajerova A, Erster S and Moll UM. The post-translational phosphorylation and acetylation modification profile is not the determining factor in targeting endogenous stress-induced p53 to mitochondria.[J] Cell Death Differ, 2005, 12:197-200.
    [142] Heyne K, Schmitt K, Mueller D, et al. Resistance of mitochondrial p53 to dominant inhibition.[J] Mol Cancer, 2008, 7:54.
    [143] Deng X, Gao F and May WS. Protein phosphatase 2A inactivates Bcl2's antiapoptoticfunction by dephosphorylation and up-regulation of Bcl2-p53 binding.[J] Blood, 2009, 113:422-428.
    [144] Sawada M, Sun W, Hayes P, et al. Ku70 suppresses the apoptotic translocation of Bax to mitochondria.[J] Nat Cell Biol, 2003, 5:320-329.
    [145] Yamaguchi H, Woods NT, Piluso LG, et al. p53 acetylation is crucial for its transcription-independent proapoptotic functions.[J] J Biol Chem, 2009, 284:11171-11183.
    [146] Leu JI and George DL. Hepatic IGFBP1 is a prosurvival factor that binds to BAK, protects the liver from apoptosis, and antagonizes the proapoptotic actions of p53 at mitochondria.[J] Genes Dev, 2007, 21:3095-3109.
    [147] Ohtsuka T, Ryu H, Minamishima YA, et al. ASC is a Bax adaptor and regulates the p53-Bax mitochondrial apoptosis pathway.[J] Nat Cell Biol, 2004, 6:121-128.
    [148] Zhang H, Kim JK, Edwards CA, et al. Clusterin inhibits apoptosis by interacting with activated Bax.[J] Nat Cell Biol, 2005, 7:909-915.
    [149] Guo B, Zhai D, Cabezas E, et al. Humanin peptide suppresses apoptosis by interfering with Bax activation.[J] Nature, 2003, 423:456-461.
    [150] Li N, Zheng Y, Chen W, et al. Adaptor protein LAPF recruits phosphorylated p53 to lysosomes and triggers lysosomal destabilization in apoptosis.[J] Cancer Res, 2007, 67, 11176-11185.
    [1] Yuspa SH and Harris CC. Altered differentiation of mouse epidermal cells treated with retinyl acetate in vitro.[J] Exptl Cell Res, 1974, 86:95-105.
    [2] Yuspa SH, Hennings H, Dermer P, et al. Dimethyl sulfoxide-induced enhancement of 7,12 dimethylbenz[a]anthracene metabolism and DNA binding in differentiating mouse epidermal cell cultures.[J] Cancer Res, 1976, 36:947-951.
    [3] Yuspa SH, Lichti Y, Ben T, et al. Phorbol-ester tumor promoters stimulate DNA synthesis and ornithine decarboxylase activity in mouse epidermal cell cultures.[J] Nature, 1976, 262:402-404.
    [4] Yuspa SH, Morgan DL and Levy JA. In vitro cultivation of a chemically induced epidermal carcinoma: establishment of three cell lines and isolation of murine leukemia virus.[J] J Natl Cancer Inst, 1973, 50:1561-1570.
    [5] Sun T and Green H. Differentiation of the epidermal keratinocyte in cell culture: formation of the cornified envelop.[J] Cell, 1976, 9:511-521.
    [6] Colburn NH, Bruegge WF, Bates JR, et al. Correlation of anchorage-independent growth with tumorigenicity of chemically transformed mouse epidermal cells.[J] Cancer Res, 1978, 38:624-634.
    [7] Colburn NH, Former BF, Nelson KA, et al. Tumor promoter induces anchorage independence irreversibly.[J] Nature, 1979, 281:589-591.
    [8] Nakamura Y, Colburn NH and Gindhart TD. Role of reactive oxygen in tumor promotion: implication of superoxide anion in promotion of neoplastic transformation in JB-6 cells by TPA.[J] Carcinogenesis, 1985, 6:229-235.
    [9] Ghosh R, Amstad P and Cerutti P. UVB-induced DNA breaks interfere with transcriptional induction of c-fos.[J] Mol Cell Biol, 1993, 13:6992-6999.
    [10] Amstad PA, Liu H, Ichimiya M, et al. Bcl-2 enhancement of malignant transformation in mouse epidermal JB6 cells.[J] Mol Carcinog, 1997, 20:231-239.
    [11] Gupta A, Rosenberger SF and Bowden GT. Increased ROS levels contribute to elevated transcription factor and MAP kinase activities in malignantly progressed mouse keratinocyte cell lines.[J] Carcinogenesis, 1999, 20:2063-2073.
    [12] Hsu TC, Nair R, Tulsian P, et al. Transformation nonresponsive cells owe their resistance to lack of p65/nuclear factor-kappaB activation.[J] Cancer Res, 2001, 61:4160-4168.
    [13] Trachootham D, Alexandre J and Huang P. Targeting cancer cells by ROS-mediated mechanisms: a radical therapeutic approach?[J] Nat Rev Drug Discov, 2009, 8:579-591.
    [14] Clerkin JS, Naughton R, Quiney C, et al. Mechanisms of ROS modulated cell survival during carcinogenesis.[J] Cancer Lett, 2008, 266:30-36.
    [15] Zhao Y, Oberley TD, Chaiswing L, et al. Manganese superoxide dismutase deficiency enhances cell turnover via tumor promoter-induced alterations in AP-1 and p53-mediated pathways in a skin cancer model.[J] Oncogene, 2002, 21:3836-3846.
    [16] Zhao Y, Chaiswing L, Velez JM, et al. p53 translocation to mitochondria precedes its nuclear translocation and targets mitochondrial oxidative defense protein-manganese superoxide dismutase.[J] Cancer Res, 2005, 65:3745-3750.
    [17] Birch-Machin MA. Jackson S, Kler RS, et al. 1993. Mitochondrial dysfunction. San Diego: Academic Press: 51-69.
    [18] Zhao Y, Wang LM, Chaiswing L, et al. Tamoxifen protects against acute tumor necrosis factor alpha-induced cardiac injury via improving mitochondrial functions.[J] Free Radic Biol Med, 2006, 40:1234-1241.
    [19] Chipuk JE and Green DR. Cytoplasmic p53: bax and forward.[J] Cell Cycle, 2004,3:429-431.
    [20] Levine AJ, Hu W and Feng Z. The p53 pathway: what questions remain to be explored?[J] Cell Death Differ, 2006, 13:1027-1036.
    [21] Murphy ME, Leu JI and George DL. p53 moves to mitochondria.[J] Cell Cycle, 2004, 3:836-839.
    [1] Zhao Y, Oberley TD, Chaiswing L, et al. Manganese superoxide dismutase deficiency enhances cell turnover via tumor promoter-induced alterations in AP-1 and p53-mediated pathways in a skin cancer model.[J] Oncogene, 2002, 21:3836-3846.
    [2] Zhao Y, Chaiswing L, Oberley TD, et al. A mechanism based antioxidant approach for the reduction of skin carcinogenesis.[J] Cancer Res, 2005, 65:1401-1405.
    [3] Zhao Y, Chaiswing L, Velez JM, et al. p53 translocation to mitochondria precedes its nuclear translocation and targets mitochondrial oxidative defense protein-manganese superoxide dismutase.[J] Cancer Res, 2005, 65:3745-3750.
    [4] Zhao Y, Chaiswing L, Bakthavatchalu V, et al. Ras mutation promotes p53 activation and apoptosis of skin keratinocytes.[J] Carcinogenesis, 2006, 27:1692-1698.
    [5] Manfredi JJ. p53 and apoptosis: it’s not just in the nucleus anymore.[J] Mol Cell, 2003, 11:552-554.
    [6] Chipuk JE and Green DR. p53’s believe it or not: lessons on transcription-iindependent death.[J] J Clin Immunol, 2003, 23:355-361.
    [7] Murphy ME, Leu JI and George DL. p53 moves to mitochondria.[J] Cell Cycle, 2004, 3(7):836-839.
    [8] Moll UM, Wolff S, Speidel D, et al. Transcription-independent pro-apoptotic functions of p53.[J] Curr Opin Cell Biol, 2005, 17:631-636.
    [9] Li PF, Dietz R and von Harsdorf R. p53 regulates mitochondrial membrane potential through reactive oxygen species and induces cytochrome c-independent apoptosis blocked by Bcl-2.[J] EMBO J, 1999, 18:6027-6036.
    [10] Marchenko ND, Zaika AI and Moll UM.. Death signal-induced localization of p53 protein to mitochondria. A potential role in apoptotic signaling.[J] J Biol Chem, 2001, 275:16202-16212.
    [11] Erster S, Mihara M, Kim RH, et al. In vivo mitochondrial p53 translocation triggers a rapidfirst wave of cell death in response to DNA damage that can precede p53 target gene activation.[J] Mol Cell Biol, 2004, 24:6728-6741.
    [12] Colburn NH, Former BF, Nelson KA, et al. Tumor promoter induces anchorage independence irreversibly.[J] Nature, 1979, 281:589-591.
    [13] Birch-Machin MA, Jackson S, Kler RS, et al. Mitochondrial Dysfunction. San Diego: Academic Press; 1993:51-69.
    [14] Zhao Y, Wang LM, Chaiswing L, et al. Tamoxifen protects against acute tumor necrosis factor alpha-induced cardiac injury via improving mitochondrial functions.[J] Free Radic Biol Med, 2006, 40:1234-1241.
    [15] Palacios G and Moll UM. Mitochondrially targeted wild-type p53 suppresses growth of mutant p53 lymphomas in vivo.[J] Oncogene, 2006, 25:6133-6139.
    [16] Mihara M, Erster S, Zaika A, et al. p53 has a direct apoptogenic role at the mitochondria.[J] Mol Cell, 2003, 11:577-590.
    [17] Leu JI, Dumont P, Hafey M, et al. Mitochondrial p53 activates Bak and causes disruption of a Bak-Mcll complex.[J] Nat Cell Biol, 2004, 6:443-450.
    [18] Chipuk JE, Kuwana T, Bouchier-Hayes L, et al. Direct activation of Bax by p53 mediates mitochondrial membrane permeabilization and apoptosis.[J] Science, 2004, 303:1010-1014.
    [19] Marchenko ND, Wolff S, Erster S, et al. Monoubiquitylation promotes mitochondrial p53 translocation.[J] EMBO J, 2007, 26:923-934.
    [20] Raymond MA, Mollica L, Vigneault N, et al. Blockade of apoptotic machinery by cyclosporine A redirects cell death toward necrosis in arterial endothelial cells: regulation by reactive oxygen species and cathepsin D.[J] FASEB J, 2003, 17:515-517.
    [1] Brand MD. The proton leak across the mitochondrial innermembrane.[J] Biochim Biophys Acta, 1990, 1018:128-133.
    [2] Brand MD, Chien LF, Ainscow EK, et al. The causes and functions of mitochondrial proton leak.[J] Biochim Biophys Acta, 1994, 1187:132-139.
    [3] Brown GC and Brand MD. On the nature of the mitochondrial proton leak.[J] Biochim Biophys Acta, 1991, 1059:55-62.
    [4] Brand MD, Brindle KM, Buckingham JA, et al. The significance and mechanism of mitochondrial proton conductance.[J] Int J Obes Relat Metab Disord, 1999, 23:S4-S11.
    [5] Porter RK and Brand MD. Causes of differences in respiration rate of hepatocytes from mammals of different body mass.[J] Am J Physiol, 1995, 269:R1213-R1224.
    [6] Rolfe DFS, Newman JMB, Buckingham JA, et al. Contribution of mitochondrial proton leak to respiration rate in working skeletal muscle and liver and to SMR.[J] Am J Physiol, 1999, 276:C692-C699.
    [7] Samudio I, Fiegl M and Andreeff M. Mitochondrial uncoupling and the warburg effect: molecular basis for the reprogramming of cancer cell metabolism.[J] Cancer Res, 2009, 69:2163-2166.
    [8] Martindale JL and Holbrook NJ. Cellular response to oxidative stress: signaling for suicide and survival.[J] Cell Physiol, 2002, 192:1-15.
    [9] Schumacker PT. Reactive oxygen species in cancer cells: live by the sword, die by the sword.[J] Cancer Cell, 2006, 10:175-176.
    [10] Halliwell B. Oxidative stress and cancer: have we moved forward?[J] Biochem J, 2007, 401:1-11.
    [11] Nègre-Salvayre A, Hirtz C, Carrera G, et al. A role for uncoupling protein-2 as a regulator of mitochondrial hydrogen peroxide generation.[J] FASEB J, 1997, 11:809-815.
    [12] Kowaltowski AJ, Costa AD and Vercesi AE. Activation of the potato plant uncoupling mitochondrial protein inhibits reactive oxygen species generation by the respiratory chain.[J]FEBS Lett, 1998, 425:213-216.
    [13] Kowaltowski AJ, de Souza-Pinto NC, Castilho RF, et al. Mitochondria and reactive oxygen species.[J] Free Radic Bio Med, 2009, 47:333-343.
    [14] H Harper ME, Antoniou A, Villalobos-menuey E, et al. Characterization of a novel metabolic strategy used by drug-resistant tumor cells.[J] FASEB J, 2002, 16:1550-1557.
    [15] Horimoto M, Resnick MB, Konkin TA, et al. Expression of uncoupling protein-2 in human colon cancer.[J] Clin Cancer Res, 2004, 10:6203-6207.
    [16] Levine JJ. p53, the cellular gatekeeper for growth and division.[J] Cell, 1997, 88:323-331.
    [17] Hollstein M, Rice K, Greenblatt MS, et al. Datebase of p53 gene somatic mutations in human tumors and cell lines.[J] Nucleic Acids Res, 1994, 22:3551-3555.
    [18] Samudio I, Fiegl M, McQueen T, et al. The Warburg effect in leukemia-stroma cocultures is mediated by mitochondrial uncoupling associated with uncoupling protein 2 activation.[J] Cancer Res, 2008, 68:5198-5205.
    [19] Derdak Z, Mark NM, Beldi G, et al. The mitochondrial uncoupling protein-2 promotes chemoresistance in cancer cells.[J] Cancer Res, 2008, 68:2813-2819.
    [20] Zhao Y, Oberley TD, Chaiswing L, et al. Manganese superoxide dismutase deficiency enhances cell turnover via tumor promoter-induced alterations in AP-1 and p53-mediated pathways in a skin cancer model.[J] Oncogene, 2002, 21:3836-3846.
    [21] Zhao Y, Chaiswing L. Oberley TD, et al. A mechanism based antioxidant approach for the reduction of skin carcinogenesis.[J] Cancer Res, 2005, 65:1401-1405.
    [22] Zhao Y, Chaiswing L, Velez JM, et al. p53 translocation to mitochondria precedes its nuclear translocation and targets mitochondrial oxidative defense protein-manganese superoxide dismutase.[J] Cancer Res, 2005, 65:3745-3750.
    [23] Zhao Y, Chaiswing L, Bakthavatchalu V, et al. Ras mutation promotes p53 activation and apoptosis of skin keratinocytes.[J] Carcinogenesis, 2006, 27:1692-1698.
    [24] Hwang PM, Bunz F, Yu J, et al. Ferredoxin reductase affects p53-dependent, 5-fluorouracil-induced apoptosis in colorectal cancer cells.[J] Nat Med, 2001, 7:1111-1117.
    [25] Colburn NH, Former BF, Nelson KA, et al. Tumor promoter induces anchorage independence irreversibly.[J] Nature, 1979, 281:589-591.
    [26] Birch-Machin MA, Jackson S, Kler RS, et al. Mitochondrial Dysfunction. San Diego:Academic Press; 1993:51-69.
    [27] Zhao Y, Wang LM, Chaiswing L, et al. Tamoxifen protects against acute tumor necrosis factor alpha-induced cardiac injury via improving mitochondrial functions.[J] Free Radic Biol Med, 2006, 40:1234-1241.
    [28] Colburn NH, Vorder Bruegge WF, Bates JR, et al. Correlation of anchorage-independent growth with tumorigenicity of chemically transformed mouse epidermal cells.[J] Cancer Res, 1978, 38:624-634.
    [29] Moll UM, Wolff S, Speidel D, et al. Transcription-independent pro-apoptotic functions of p53.[J] Curr Opin Cell Biol, 2005, 17:631-636.
    [30] Murphy ME, Leu JI and George DL. p53 moves to mitochondria.[J] Cell Cycle 2004, 3(7):836-839.
    [31] Echtay KS. Mitochondrial uncoupling proteins– what is their physiological role.[J] Free Radic Biol Med, 2007, 43:1351-1371.
    [32] Moll UM and Zaika A. Nuclear and mitochondrial apoptotic pathways of p53.[J] FEBS Lett, 2001, 493:65-69.
    [33] Marchenko ND, Zaika AI and Moll UM. Death signal-induced localization of p53 protein to mitochondria. A potential role in apoptotic signaling.[J] J Biol Chem, 2001, 275:16202-16212.
    [34] Mihara M, Erster S, Zaika A, et al. p53 has a direct apoptogenic role at the mitochondria.[J] Mol Cell, 2003, 11:577-590.
    [35] Liu J, St. Clair DK, Gu X, et al. Blocking mitochondrial permeability transition prevents p53 mitochondrial transloation during skin tumor promotion.[J] FEBS Lett, 2008, 582:1319-1324.
    [36] Gogvadze V, Orrenius S and Zhivotovsky B. Mitochondria in cancer cells: what is so special about them?[J] Trends Cell Biol, 2008, 18:165-173.
    [37] Collins P, Jones C, Choudhury S, et al. Increased expression of uncoupling protein 2 in HepG2 cells attenuates oxidative damage and apoptosis.[J] Liver Int, 2005, 25:880-887.
    [1] Levine JJ. p53, the cellular gatekeeper for growth and division.[J] Cell, 1997, 88:323-331.
    [2] Hollstein M, Rice K, Greenblatt MS, et al. Datebase of p53 gene somatic mutations in human tumors and cell lines.[J] Nucleic Acids Res, 1994, 22:3551-3555.
    [3] Garcia-Cao I, Garcia-Cao M, Martin-Caballero J, et al. "Super p53" mice exhibit enhanced DNA damage response, are tumor resistant and age normally.[J] EMBO J, 2002, 21:6225-6235.
    [4] Tyner SD, Venkatachalam S, Choi J, et al. p53 mutant mice that display early ageing-associated phenotypes.[J] Nature, 2002, 415:45-53.
    [5] Jia HP. Controversial chinese gene-therapy drug entering unfamiliar territory.[J] Nat Rev Drug Discov, 2006, 4:269-270.
    [6] Tokino T and Nakamura Y. The role of p53-target genes in human cancer.[J] Crit Rev Oncol Hematol, 2000, 33:1-6.
    [7] Moll UM and Zaika A. Nuclear and mitochondrial apoptotic pathways of p53.[J] FEBS Lett, 2001, 493:65-69.
    [8] Li PF, Dietz R and von Harsdorf R. p53 regulates mitochondrial membrane potential through reactive oxygen species and induces cytochrome c-independent apoptosis blocked by Bcl-2.[J] EMBO J, 1999, 18:6027-6036.
    [9] Marchenko ND, Zaika AI and Moll UM. Death signal-induced localization of p53 protein to mitochondria. A potential role in apoptotic signaling.[J] J Biol Chem, 2000, 275:16202-16212.
    [10] Erster S, Mihara M, Kim RH, et al. In vivo mitochondrial p53 translocation triggers a rapid first wave of cell death in response to DNA damage that can precede p53 target gene activation.[J] Mol Cell Biol, 2004, 24:6728-6741.
    [11] Zhao Y, Chaiswing L, Velez JM, et al. p53 translocation to mitochondria precedes its nuclear translocation and targets mitochondrial oxidative defense protein-manganese superoxide dismutase.[J] Cancer Res, 2005, 65:3745-3750.
    [12] Mihara M, Erster S, Zaika A, et al. p53 has a direct apoptogenic role at the mitochondria.[J] Mol Cell, 2003, 11:577-90.
    [13] Heyne K, Schmitt K, Mueller D, et al. Resistance of mitochondrial p53 to dominant inhibition.[J] Mol Cancer, 2008, 7:54.
    [14] Tang X, Zhu Y, Han L, et al. CP-31398 restores mutant p53 tumor suppressor function and inhibits UVB-induced skin carcinogenesis in mice.[J] J Clin Invest, 2007, 117:3753-3764.
    [15] Mahyar-Roemer M, Fritzsche C, Wagner S, et al. Mitochondrial p53 levels parallel total p53 levels independent of stress response in human colorectal carcinoma and glioblastoma cells.[J] Oncogene, 2004, 23:6226-6236.
    [16] Allman R, Errington RJ and Smith PJ. Delayed expression of apoptosis in human lymphoma cells undergoing low-dose taxol-induced mitotic stress.[J] Br J Cancer, 2003, 88:1649-1658.
    [17] Gao C, Nakajima T, Taya Y, et al. Activation of p53 in MDM2-overexpressing cells through phosphorylation.[J] Biochem Biophys Res Commun, 1999, 264:860-864.
    [18] Karpova MB, Sanmun D, Henter JI, et al. Betulinic acid, a natural cytotoxic agent, fails to trigger apoptosis in human Burkitt's lymphoma-derived B-cell lines.[J] Int J Cancer, 2006, 118:246-252.
    [19] Yen HC, Oberley TD, Vichitbandha S, et al. The protective role of manganese superoxide dismutase against adriamycin-induced acute cardiac toxicity in transgenic mice.[J] J Clin Invest, 1996, 98:1253-1260.
    [20] Zhao Y, Wang LM, Chaiswing L, et al. Tamoxifen protects against acute tumor necrosis factor alpha-induced cardiac injury via improving mitochondrial functions.[J] Free Radic Bio Med, 2006, 40:1234-1241.
    [21] Bykov VJ, Selivanova G and Wiman KG. Small molecules that reactivate mutant p53.[J] Eur J Cancer, 2003, 39:1828-1834.
    [22] Duthu A, Debuire B, Romano J, et al. p53 mutations in raji cells– characterization and localization relative to other burkitt’s lymphomas.[J] Oncogene, 1992, 7:2161-2167.
    [23] Donahue RJ, Razmara M, Hoek JB, et al. Direct influence of the p53 tumor suppressor on mitochondrial biogenesis and function.[J] FASEB J, 2001, 15:635-644.
    [24] Matoba S, Kang JG, Patino WD, et al. p53 regulates mitochondrial respiration.[J] Science, 2006, 312:1650-1653.
    [25] Soussi T. p53 alterations in human cancer: more questions than answers.[J] Oncogene, 2007, 26:2145-2156.
    [26] Achanta G, Sasaki R, Feng L, et al. Novel role of p53 in maintaining mitochondrial genetic stability through interaction with DNA Pol gamma.[J] EMBO J, 2005, 24:3482-3492.
    [27] Yoshida Y, Izumi H, Torigoe T, et al. p53 physically interacts with mitochondrial transcription factor A and differentially regulates binding to damaged DNA.[J] Cancer Res, 2003, 63:3729-3734.
    [28] Brown CR, Hong-Brown LQ and Welch WJ. Correcting temperature-sensitive protein folding defects.[J] J Clin Invest, 1997, 99:1432-1444.
    [29] Selivanova G, Kawasaki T, Ryabchenko L, et al. Reactivation of mutant p53: a new strategy for cancer therapy.[J] Semin Cancer Biol, 1998, 8:369-378.
    [30] Sugikawa E and Hosoi T. Mutant p53 mediated induction of cell cycle arrest and apoptosis at G1 phase by 9-hydroxyellipticine.[J] Anticancer Res, 1999, 19:3099-3108.
    [31] Huff AC and Kreuzer KN. Evidence for a common mechanism of action for antitumor and antibacterial agents that inhibit type II DNA topoisomerases.[J] J Biol Chem, 1990, 265:20496–20505.
    [32] Peng Y, Li C, Chen L, et al. Rescue of mutant p53 transcription function by ellipticine.[J] Oncogene, 2003, 22:4478-4487.
    [33] StiborováM, Sejbal J, Borek-DohalskáL, et al. The anticancer drug ellipticine forms covalent DNA adducts, mediated by human cytochromes P450, through metabolism to 13-hydroxyellipticine and ellipticine N2-oxide.[J] Cancer Res, 2004, 64: 8374-8380.
    [34] Fornari FA, Randolph JK, Yalowich JC, et al. Interference by doxorubicin with DNA unwinding in MCF-7 breast tumor cells.[J] Mol Pharmacol, 1994, 45:649-656.
    [35] Pigram WJ, Fuller W and Hamilton LD. Stereochemistry of intercalation: interaction of daunomycin with DNA.[J] Nature New Biol, 1972, 235:17-19.
    [36] Ohashi M, Sugikawa E and Nakanishi N. Inhibition of p53 protein phosphorylation by 9-hydroxyellipticine: a possible anticancer mechanism.[J] Jpn J Cancer Res, 1995, 86:819-827.
    [37] Hubert A, Paris S, Piret JP, et al. Casein kinase 2 inhibition decreases hypoxia-inducible factor-1 activity under hypoxia through elevated p53 protein level.[J] J Cell Sci, 2006, 119:3351-3362.

© 2004-2018 中国地质图书馆版权所有 京ICP备05064691号 京公网安备11010802017129号

地址:北京市海淀区学院路29号 邮编:100083

电话:办公室:(+86 10)66554848;文献借阅、咨询服务、科技查新:66554700