• Keine Ergebnisse gefunden

II 3. Histologische und immunhistochemische Untersuchungen

II.5. Zellzyklusanalyse mittels FACS

Die FACS-Analyse (Fluorescence-Absorbance-Cell-Sorting) ermöglicht es, eine große Anzahl von Zellen in sehr kurzer Zeit einzeln auf ihre biochemischen oder biophysikalischen Eigenschaften zu untersuchen.

In einem sehr feinen Wasserstrahl wandern die zu untersuchenden Zellen hintereinander und werden dabei von einem Laser mit relativ kurzer Wellenlänge (!=488 nm) angeregt. Hierbei werden nun verschiedene Parameter bestimmt. Der Forward-Scatter (FSC) ist ein Maß für das Licht, welches im 180°-Winkel abgeschwächt aus der angeregten Zelle austritt und ist proportional zur Zellgröße. Der side-scatter (SSC) bezeichnet das im 90° Winkel abgegebene Streulicht und gilt als Maß für die Granularität. Aus diesen zwei Parametern können sehr gezielt Zellpopulationen bestimmt und definiert werden. Werden die Zellen zusätzlich mit einem Farbstoff oder einem fluoreszenzgekoppelten Antikörper inkubiert, so können neben den schon erwähnten Parametern weitere zur Fluoreszenz der Zellen erhoben werden, beispielsweise Rezeptorstatus, Differenzierungsmarker, etc.

Propidiumiodid interkaliert wie Ethydiumbromid mit DNA. Wird der Komplex aus DNA und Propidiumiodid mit kurzwelligem, blauen Licht angeregt, so emittiert dieser im roten Bereich. Diese Fluoreszenz wird über einen entsprechenden Filter detektiert. Die hierbei gemessene Lichtmenge ist direkt proportional zum DNA-Gehalt der Zelle. Visualisiert werden die erhobenen Werte in einem Histogramm, dessen x-Achse die Fluoreszenzintensität und dessen y-Achse die absolute Zellzahl repräsentieren. Über weitere statistische Analysen ist es nun möglich, die gemessenen Daten hinsichtlich Genauigkeit, Validität u.ä. zu untersuchen und die Anzahl der Zellen in den verschieden Abschnitten des Zyklus anzugeben.

p53cbs Promotor Luciferase-Gen

p53!

Luciferase Luciferin+ATP+H2O2+CoA

Licht+ADP+P+H2O

5.1. Zellzyklusanalysen von MCF-7-Zellen

Um ein möglichst große Anzahl von Zellen einer Population in der S- oder G2/M-Phase untersuchen zu können, wurden MCF-7 Zellen mit 0,5% FCS-supplementierten RPMI-Medium ohne den Zusatz von Insulin für 32 Stunden inkubiert. Hierdurch werden die Zellen in der G0/G1-Phase des Teilungszyklus arretiert. Nach der Zugabe von 10%igem FCS RPMI-Medium und Insulin wurde durch die im Medium enthaltenen Wachstumsfaktoren ein Teilungsstimulus induziert und die Zellen weitere 27 Stunden inkubiert. Nach dieser Zeit befinden sich etwa 65% der Zellen entweder in der S- oder G2/M-Phase. Bei einer weiteren Probe wurde direkt nach Zugabe des Wachstumsmediums mit unterschiedlichen Energiedosen bestrahlt, um einen Teilungsarrest zu induzieren und anschließend in An- oder Abwesenheit der modulatorischen Peptide MCF-7-Zellen auf ihre Teilungsaktivität mittels FACS untersuchen zu können. Andere Versuche, Zellen mit Hydroxyharnstoff (RNA-Polymerase-Inhibitor)oder Mimosin (Dihydrofolatreduktase-Inhibitor) synchronisieren zu können, erwiesen sich wegen der Toxizität der verwendeten Stoffe als nicht durchführbar.

H Literaturverzeichnis

Abraham, J., Spaner, D., and Benchimol, S. (1999). Phosphorylation of p53 protein in response to ionizing radiation occurs at multiple sites in both normal and DNA-PK deficient cells. Oncogene 18, 1521-1527.

Aints, A., and al., e. (1999). Intercellular spread of GFP-VP22. J Gene Med 1, 275-279.

Appella, A. (2001). Post-translational modifications and activation of p53 by genotoxic stresses. Eur J Biochem 268,, 2764-2772.

Balagurumoorthy, P., and al., e. (1995). Four p53 DNA-binding domain peptides bind natural p53-response elements and bend the DNA. Proc Natl Acad Sci USA 92, 8591-8595.

Barak, Y., Juven, T., Haffner, R., and Oren, M. (1993). mdm2 expression is induced by wild type p53 activity. Embo J 12, 461-468.

Berlose, J. P. (1996). Conformational and associated behaviour of the third helix of antennapedia homeodomain in membrane-mimetic environments. Eur J Biochem 242, 372-386.

Bourdon, J. C., Renzing, J., Robertson, P. L., Fernandes, K. N., and Lane, D. P. (2002). Scotin, a novel p53-inducible proapoptotic protein located in the ER and the nuclear membrane. J Cell Biol 158, 235-246.

Bradford, M. M. (1976). A rapid and sensitive method for the quantitation of microgramm quantities of protein utilizing the principle of protein-dye binding. Anal Biochem 72, 248-254.

Brandenburg, B., and al., e. (2001). Gentransfer unter Verwendung zellpermebaler, DNAbindender Fusionspeptide und zellpermeabler Nukleokapside; in Fakultät für Biologie, Chemieund Pharmazie; Freie Universität Berlin.

Brandenburg, B., Stockl, L., Gutzeit, C., Roos, M., Lupberger, J., Schwartlander, R., Gelderblom, H., Sauer, I. M., Hofschneider, P. H., and Hildt, E. (2005). A novel system for efficient gene transfer into primary human hepatocytes via cell-permeable hepatitis B virus-like particle. Hepatology 42, 1300-1309.

Buschmann, T., and al., e. (2000a). p53 phosphorylation and association with murine double minute 2, c-Jun NH2-terminal kinase, p14ARF, and p300(CBP during cell cycle and after exposure to ultraviolet irradiation. Cancer Res 60, 896-900.

Buschmann, T., and al., e. (2000b). SUMO-1 modification and activation of p53 by the coactivator protein TAFII31.

Cell 101, 753-763.

Canman, C. E., and al., e. (1998). Activation of the ATM kinase by ionizing radiation and phosphorylation of p53.

Science 281, 1677-1679.

Canman, C. E., Lim, D. S., Cimprich, K. A., Taya, Y., Tamai, K., Sakaguchi, K., Appella, E., Kastan, M. B., and Siliciano, J.

D. (1998). Activation of the ATM kinase by ionizing radiation and phosphorylation of p53. Science 281, 1677-1679.

Chiarugi, V., and al., e. (1998). Acetylation and phosphorylation of the carboxy-terminal domain of p53: regulative significance. Oncol Res 10, 55-57.

Dameron, K. M., Volpert, O. V., Tainsky, M. A., and Bouck, N. (1994). The p53 tumor suppressor gene inhibits

angiogenesis by stimulating the production of thrombospondin. Cold Spring Harb Symp Quant Biol 59, 483-489.

Deleo, A. B., and al., e. (1978). Detection of a transformation related antigen in chemically induced sarcomas and other transformed cells of the mouse. Proc Nat Acad Sci USA 76, 2420-2424.

Derossi, D., and al., e. (1994). The third helix of the antennapedia homeodomain translocates through biological membranes. J Biol Chem 269, 10444-10450.

Derossi, D., and al., e. (1996). Cell internalization of the third helix of the Antennapedia homeodomain is receptor independent. J Biol Chem 271, 18188-18193.

Derossi, D., and al., e. (1998). Trojan peptides: the penetratin system for intracelular delivery. Trends Cell Biol 8, 84-87.

Dilber, M. S., and al., e. (1999). Intercellular delivery oof thymidine kinase prodrug activating enzyme by the herpes simplex virus protein, VP22. Gene Ther 6, 12-21.

Dornan, D., Shimizu, H., Perkins, N. D., and Hupp, T. R. (2003). DNA-dependent acetylation of p53 by the transcription coactivator p300. J Biol Chem 278, 13431-13441.

Drin, G., and al., e. (2001). Translocation of the pAntp peptide and its amphipathic analogue AP-2AL. Biochemistry 40, 1824-1834.

Du, C., and al., e. (1998). Conformational and topological requirements of cell-permeable peptide function. J Peptide Res 51, 235-243.

Dumaz, N., Milne, D. M., and Meek, D. W. (1999). Protein kinase CK1 is a p53-threonine 18 kinase which requires prior phosphorylation of serine 15. FEBS Lett 463, 312-316.

Edelstein, M. L., Abedi, M. R., Wixon, J., and Edelstein, R. M. (2004). Gene therapy clinical trials worldwide 1989-2004-an overview. J Gene Med 6, 597-602.

Elliott, G., and O`Hare, P. (1997). Intercellular trafficking and protein delivery by by herpesvirus structural protein.

Cell 88, 223-233.

Elliott, G., and O`Hare, P. (1999). Intercellular trafficking of VP22-GFP fusion proteins., In Gene Ther., pp. 149-151.

Fischer, P. M., and al., e. (2000). Structure-activity relationship of truncated and substituted analogues of the intracellular delivery vector penetratin. J Peptide Res 55, 163-172.

Futaki, S., and al, e. (2001). Arginine-rich peptides. An abundant source of membrane-permeable peptides having potential as carriers for intracellular protein delivery. J Biol Chem 276, 5836-5840.

Gadea, G., and al., e. (2002). Regulation of Cdc-42-mediated morphological effects: a novel function for p53. EMBO J 21, 2373-2382.

Giannakakou, P., and al., e. (2000). p53 is associated with cellular microtubules and is transported to the nucleus by dynein. Nat Cell Biol 10, 709-717.

Gonin, P., Buchholz, C. J., Pallardy, M., and Mezzina, M. (2005). Gene therapy bio-safety: scientific and regulatory issues. Gene Ther 12 Suppl 1, S146-152.

Gottlieb, T. M., and al., e. (1998). p53 and apoptosis. Semin Cancer Biol 8, 359-368.

Gottlieb, T. M., and al., e. (2002). Cross-talk between Akt, p53 and mdm2: possible implications for the regulation of apoptosis. Oncogene 21, 1299-1303.

Grieger, J. C., and Samulski, R. J. (2005). Adeno-associated virus as a gene therapy vector: vector development, production and clinical applications. Adv Biochem Eng Biotechnol 99, 119-145.

Gu, W., and al., e. (1997). Activation of p53 sequence-specific DNA binding by acetylation of the p53 C-terminal domain. Cell 90, 595-606.

Hacker, G. (2000). The morphology of apoptosis. Cell Tissue Res 301, 5-17.

Hafner, A., Brandenburg, B., and Hildt, E. (2003). Reconstitution of gene expression from a regulatory-protein-deficient hepatitis B virus genome by cell-permeable HBx protein. EMBO Rep 4, 767-773.

Haupt, Y., and al., e. (1997). Mdm2 promotes the rapid degredation of p53. Nature 387, 299-303.

Hermeking, H., and al., e. (1997). 14-3-3sigma is a p53-regulated inhibitor of G2/M progression. Mol Cell 1, 3-11.

Hildt, E., and al., e. (1999). Identification of Grb2 as a novel binding partner of tumor necrosis factor (TNF) receptor I.

J Exp Med 189, 1707-1714.

Hillemann, A., Brandenburg, B., Schmidt, U., Roos, M., Smirnow, I., Lemken, M. L., Lauer, U. M., and Hildt, E. (2005).

Protein transduction with bacterial cytosine deaminase fused to the TLM intercellular transport motif induces profound chemosensitivity to 5-fluorocytosine in human hepatoma cells. J Hepatol 43, 442-450.

Hirao, A., and al., e. (2000). DNA damage-induced activation of p53 by the checkpoint kinase Chk2. Science 287, 1824-1827.

Ho, A., and al., e. (2001). Synthetic protein transduction domains: Enhanced transduction potential in vitro and in vivo. Cancer Res 61, 474-477.

Hollstein, M., and al., e. (1991). p53 mutations in human cancers. Science 253, 49-53.

Hong, T. M., and al., e. (2001). P53 amino acids 339-346 represent the minimal p53 repression domain. J Biol Chem 276, 1510-1515.

Horstmann, E., McCabe, M. S., Grochow, L., Yamamoto, S., Rubinstein, L., Budd, T., Shoemaker, D., Emanuel, E. J., and Grady, C. (2005). Risks and benefits of phase 1 oncology trials, 1991 through 2002. N Engl J Med 352, 895-904.

Hupp, T. R., Meek, D. W., Midgley, C. A., and Lane, D. P. (1992). Regulation of the specific DNA binding function of p53. Cell 71, 875-886.

Ihrie, R. A., Reczek, E., Horner, J. S., Khachatrian, L., Sage, J., Jacks, T., and Attardi, L. D. (2003). Perp is a mediator of p53-dependent apoptosis in diverse cell types. Curr Biol 13, 1985-1990.

Jeffrey, P. D., and al., e. (1995). Crystal structure of the tetramerization domain of the p53 tumor suppressor at 1,7 angstroms. Science 267, 1498-1502.

Joliot, A., and Prochiantz, A. (2004). Transduction peptides: from technology to physiology. Nat Cell Biol 6, 189-196.

Karbowski, M., and Youle, R. J. (2003). Dynamics of mitochondrial morphology in healthy cells and during apoptosis.

Cell Death Differ 10, 870-880.

Keller, D. M., and al., e. (2001). DNA damage-induced p53 serine 392 kinase complex contains CK2, hSpt16 and SSRP1. Mol Cell 7, 283-292.

Khanna, K. K., and al., e. (1998). ATM associates with and posphorylates p53: mapping the region of interaction. Nat Genet 20, 398-400.

Kishi, H., Nakagawa, K., Matsumoto, M., Suga, M., Ando, M., Taya, Y., and Yamaizumi, M. (2001). Osmotic shock induces G1 arrest through p53 phosphorylation at Ser33 by activated p38MAPK without phosphorylation at Ser15 and Ser20. J Biol Chem 276, 39115-39122.

Lakin, N. D., and al., e. (1999). The ataxia-teleangiectatica related protien ATR mediates DNA-dependent phosphorylation of p53. Oncogene 18, 3989-3995.

Lan, K. H., and al., e. (2002). HCV NS5A interacts with p53 and inhibits p53-mediated apoptosis. Oncogene 18, 4801-4811.

Langley, E., Pearson, M., Faretta, M., Bauer, U. M., Frye, R. A., Minucci, S., Pelicci, P. G., and Kouzarides, T. (2002).

Human SIR2 deacetylates p53 and antagonizes PML/p53-induced cellular senescence. Embo J 21, 2383-2396.

Li, L., Ljungman, M., and Dixon, J. E. (2000). The human Cdc14 phosphatases interact with and dephosphorylate the tumor suppressor protein p53. J Biol Chem 275, 2410-2414.

Li, M., and al., e. (2002). Deubiquitination of p53 by HAUSP is an important pathway for p53 stabilization. Nature 416, 648-653.

Li, Y., Raffo, A. J., Drew, L., Mao, Y., Tran, A., Petrylak, D. P., and Fine, R. L. (2003). Fas-mediated apoptosis is

dependent on wild-type p53 status in human cancer cells expressing a temperature-sensitive p53 mutant alanine-143. Cancer Res 63, 1527-1533.

Liang, S. H., and al., e. (1999). A bipartite nuclear localization signal is required for p53 nuclear import regulated by a carboxy-terminal domain. J Biol Chem 274, 32699-32703.

Lin, Y., and al., e. (2000). PIDD, a new death-domain-containing protein, is induced by p53 and promotes apoptosis.

Nat Genet 26, 122-127.

Liu, L., Scolnick, D. M., Trievel, R. C., Zhang, H. B., Marmorstein, R., Halazonetis, T. D., and Berger, S. L. (1999). p53 sites acetylated in vitro by PCAF and p300 are acetylated in vivo in response to DNA damage. Mol Cell Biol 19, 1202-1209.

Magzoub, M., and al., e. (2001). Interaction and structure induction of cell-penetrating peptides in the presence of phospholipid vesicles. Biochim Biophys Acta 1512, 77-89.

Matsuda, K., and al, e. (2002). p53AIP1 regulates the mitochondrial apoptotic pathway. Cancer Res 62, 2883-2889.

McLure, K. G., and al., e. (1998). How p53 bind DNA as a tetramer. EMBO J 17, 3342-3350.

Momand, J., and al., e. (2000). Mdm2- master regulator of the p53 tumor suppressor protein. Gene 242, 15-29.

Morris, M. C., and al., e. (1997). A new peptide vector for efficient delivery of oligonucleotides into mammalian cells.

Nucleic Acids Res 25, 2730-2736.

Nagahra, H., and al., e. (1998). Transduction of full length Tat fusion proteins into mammalian cells; Tat-p27Kip1 induces cell migration. Nat Med 4, 1449-1452.

Nakase, I., Niwa, M., Takeuchi, T., Sonomura, K., Kawabata, N., Koike, Y., Takehashi, M., Tanaka, S., Ueda, K., Simpson, J. C., et al. (2004). Cellular uptake of arginine-rich peptides: roles for macropinocytosis and actin rearrangement.

Mol Ther 10, 1011-1022.

Oda, E., and al., e. (2000). Noxa, a BH3-only member of the Bcl-2 family and candidate mediator of p53-induced apoptosis. Science 288, 1053-1058.

Oda, K., Arakawa, H., Tanaka, T., Matsuda, K., Tanikawa, C., Mori, T., Nishimori, H., Tamai, K., Tokino, T., Nakamura, Y., and Taya, Y. (2000). p53AIP1, a potential mediator of p53-dependent apoptosis, and its regulation by Ser-46-phosphorylated p53. Cell 102, 849-862.

Oehlke, J., and al., e. (1998). Cellular uptake of an alpha-helical amphiphatic model peptide with the potential to deliver polar compound into the cell interior non-endocytically. Biochim Biophys Acta 1414, 127-139.

Oess, S. (2000). Entdeckung und Charakterisierung eines neuen Zellpermeabilität vermittelnden Peptidmotives aus den Oberflächenproteinen des Hepatitis B-Virus; in Fakultät für Chemie und Pharmazie; Ludwig-Maximilians-Universität München.

Oess, S., and al., e. (2000). Novel cell permeable motif derived from the PreS2-domain of hepatitis-B virus surface antigens. Gene Ther 7, 750-758.

Oess, S., and Hildt, E. (2000). Novel cell permeable motif derived from the PreS2-domain of hepatitis-B virus surface antigens. Gene Ther 7, 750-758.

Ohiro, Y., and al., e. (2002). A novel p53 inducible apoptotic gene, PGR3, encodes a homologue of the apoptosis-inducing factor (AIF). FEBS Lett 524, 163-171.

Ohki, R., Nemoto, J., Murasawa, H., Oda, E., Inazawa, J., Tanaka, N., and Taniguchi, T. (2000). Reprimo, a new candidate mediator of the p53-mediated cell cycle arrest at the G2 phase. J Biol Chem 275, 22627-22630.

Okorokov, A. L., Rubbi, C. P., Metcalfe, S., and Milner, J. (2002). The interaction of p53 with the nuclear matrix is mediated by F-actin and modulated by DNA damage. Oncogene 21, 356-367.

Park, C. B., Yi, K. S., Matsuzaki, K., Kim, M. S., and Kim, S. C. (2000). Structure-activity analysis of buforin II, a histone H2A-derived antimicrobial peptide: the proline hinge is responsible for the cell-penetrating ability of buforin II. Proc Natl Acad Sci U S A 97, 8245-8250.

Pavletich, N. P., and al., e. (1993). The DNA-binding domain of p53 contains the four conserved regions and the

major mutation hot spots. Genes Dev 7, 2556-2564.

Phelan, A., and al., e. (1998). Intercellular delivery of functional p53 by herpesvirus protein VP22. Nat Biotechnol 16, 440-443.

Pooga, M., and al., e. (1998a). Cell penetrating PNA constructs regulate galanin receptor levels and modify pain transmission in vivo. Nat Biotechnol 16, 857-861.

Pooga, M., and al., e. (1998b). Cell penetration by transportan. FASEB J 12, 67-77.

Pooga, M., Lindgren, M., Hallbrink, M., Brakenhielm, E., and Langel, U. (1998). Galanin-based peptides, galparan and transportan, with receptor-dependent and independent activities. Ann N Y Acad Sci 863, 450-453.

Richard, J. P., Melikov, K., Vives, E., Ramos, C., Verbeure, B., Gait, M. J., Chernomordik, L. V., and Lebleu, B. (2003). Cell-penetrating peptides. A reevaluation of the mechanism of cellular uptake. J Biol Chem 278, 585-590.

Rojas, M., and al., e. (1996). Controlling epidermal growth factor (EGF)-stimulated Ras activation in intact cells by a cell permeable peptide mimicking phosphorylated EGF-receptor. J Biol Chem 271.

Saar, K., Lindgren, M., Hansen, M., Eiriksdottir, E., Jiang, Y., Rosenthal-Aizman, K., Sassian, M., and Langel, U. (2005).

Cell-penetrating peptides: a comparative membrane toxicity study. Anal Biochem 345, 55-65.

Schagger, H., and von Jagow, G. (1987). Trcicine-sodium dodecyl sulfate-polyacrylamide gel electrophoresis for the separation of proteins in the range from 1 to 100 kDa. Anal Biochem 166, 368-379.

Scheffner, M., Huibregtse, J. M., Vierstra, R. D., and Howley, P. M. (1993). The HPV-16 E6 and E6-AP complex functions as a ubiquitin-protein ligase in the ubiquitination of p53. Cell 75, 495-505.

Schmitt, C. A., and al., e. (2002). A senescence program controlled by p53 and p16INK4a contributes to the outcome of cancer therapy. Cell 109, 335-346.

Selivanova, G., and al., e. (1998). Reactivation of mutant p53: a new strategy for cancer therapy? Semin Cancer Biol 8, 369-378.

Smythe, E., and Warren, G. (1991). The mechanism of receptor-mediated endocytosis. Eur J Biochem 202, 689-699.

Steger, J. E., and al., e. (1994). p53 oligomerization and DNA looping are linked with transcriptional activation. EMBO J 13, 6011-6020.

Stoeckl, L., Funk, A., Kopitzki, A., Brandenburg, B., Oess, S., Will, H., Sirma, H., and Hildt, E. (2006). Identification of a structural motif crucial for infectivity of hepatitis B viruses. Proc Natl Acad Sci U S A.

Stommel, J. M., and al., e. (1999). A leucine-rich nuclear export signal in the p53 tetramerization domain: regulation of subcellular localization and p53 activity by NES masking. EMBO J 18.

Sturzbecher, H. W., and al., e. (1992). A C-terminal alpha-helix plus basic region motif is the major structural determinant of p53 tetramerization. Oncogene 7, 1513-1523.

Takai, H., Naka, K., Okada, Y., Watanabe, M., Harada, N., Saito, S., Anderson, C. W., Appella, E., Nakanishi, M., Suzuki,

H., et al. (2002). Chk2-deficient mice exhibit radioresistance and defective p53-mediated transcription. Embo J 21, 5195-5205.

Takimoto, R., and al., e. (2000). Wild-type p53 transactivates the KILLER/DR5 gene throug an intronic sequence specific DNA-binding site. Oncogene 19, 1735-1743.

Tanaka, H., and al., e. (2000). A ribonucleotide reductase gene involved in a p53 dependent cell-cycle checkpoint for DNA damage. Nature 404, 24-25.

Vivés, E., and al., e. (1997). A truncated HIV-1 Tat protein basic domaine rapidly translocates throug the plasma membrane and accumulates in the cell nucleus. J Biol Chem 272, 16010-16017.

Vives, E., Richard, J. P., Rispal, C., and Lebleu, B. (2003). TAT peptide internalization: seeking the mechanism of entry.

Curr Protein Pept Sci 4, 125-132.

Vogelstein, B., Lane, D., and Levine, A. (2000). Surfing the p53 network. Nature 408, 307-310.

Vousden, H. (2000). The ins and outs of p53. Nat Cell Biol 2, E178-E180.

Wagner, P., Fuchs, A., Gotz, C., Nastainczyk, W., and Montenarh, M. (1998). Fine mapping and regulation of the association of p53 with p34cdc2. Oncogene 16, 105-111.

Wang, P., and al., e. (1994). p53 domains: structure, oligomerization and transformation. Mol Cell Biol 14, 5182-5191.

Wang, Y., and al., e. (1993). p53 domains: identification and characterization of two autonomous DNA-binding regions. Genes Dev 7, 2575-2586.

Wang, Y., Debatin, K. M., and Hug, H. (2001). HIPK2 overexpression leads to stabilization of p53 protein and increased p53 transcriptional activity by decreasing Mdm2 protein levels. BMC Mol Biol 2, 8.

Weizman, T., and al., e. (1995). p21 is necessary for p53-mediated G1-arrest in human cancer cells. Cancer Res 55, 5187-5190.

Wu, X., and Deng, Y. (2002). Bax and BH3-domain-only proteins in p53-mediated apoptosis. Front Biosci 7, d151-156.

Xiao, G., Chicas, A., Olivier, M., Taya, Y., Tyagi, S., Kramer, F. R., and Bargonetti, J. (2000). A DNA damage signal is required for p53 to activate gadd45. Cancer Res 60, 1711-1719.

Xie, S., Wang, Q., Wu, H., Cogswell, J., Lu, L., Jhanwar-Uniyal, M., and Dai, W. (2001). Reactive oxygen species-induced phosphorylation of p53 on serine 20 is mediated in part by polo-like kinase-3. J Biol Chem 276, 36194-36199.

Xing, J., Sheppard, H. M., Corneillie, S. I., and Liu, X. (2001). p53 Stimulates TFIID-TFIIA-promoter complex assembly, and p53-T antigen complex inhibits TATA binding protein-TATA interaction. Mol Cell Biol 21, 3652-3661.

Zhang, Y., and al., e. (2001). A p53 amino-terminal nuclear export signal inhibited by DNA damage-induced phosphorylation. Science 292, 1910-1915.