.
. Protein–Protein Interactions: A Molecular Cloning Manual, Second EditionCSHL Press .
. . . . .
.
.
.
. .
 
.
. . .
.  FIRST VISIT?
 TRY THESE LINKS
.
. . .
.
.   Enroll for Updates
  Privacy Policy
  Purchase the book
.
.
 
 
.
. . .
.  BOOK COVER .
. . .
.
. Protein–Protein Interactions: A Molecular Cloning Manual, Second Edition cover .
.
CLICK TO ENLARGE
 
Buy the Book
 
   
 

Chapter 2: Building a Better Web: Progress in the Concept and Methodology of Protein Interaction Studies—References

Agalioti T., Chen G., and Thanos D. 2002. Deciphering the transcriptional histone acetylation code for a human gene. Cell 111: 381–392.

Aloni-Grinstein R., Zan-Bar I., Alboum I., Goldfinger N., and Rotter V. 1993. Wild type p53 functions as a control protein in the differentiation pathway of the B-cell lineage. Oncogene 8: 3297–3305.

An W., Kim J., and Roeder RG. 2004. Ordered cooperative functions of PRMT1, p300, and CARM1 in transcriptional activation by p53. Cell 117: 735–748.

Arrowsmith C.H. 1999. Structure and function in the p53 family. Cell Death Differ. 6: 1169–1173.

Badciong J.C. and Haas A.L. 2002. MdmX is a RING finger ubiquitin ligase capable of synergistically enhancing Mdm2 ubiquitination. J. Biol. Chem. 277: 49668–49675.

Baker S.J., Fearon E.R., Nigro J.M., Hamilton S.R., Preisinger A.C., Jessup J.M., vanTuinen P., Ledbetter D.H., Barker D.F., Nakamura Y., et al. 1989. Chromosome 17 deletions and p53 gene mutations in colorectal carcinomas. Science 244: 217–221.

Baltimore D. 1970. RNA-dependent DNA polymerase in virions of RNA tumour viruses. Nature 226: 1209–1211.

Barak Y., Juven T., Haffner R., and Oren M. 1993. mdm2 expression is induced by wild type p53 activity. EMBO J. 12: 461–468.

Bargonetti J., Friedman P.N., Kern S.E., Vogelstein B., and Prives C. 1991. Wild-type but not mutant p53 immunopurified proteins bind to sequences adjacent to the SV40 origin of replication. Cell 65: 1083–1091.

Bargonetti J., Manfredi J.J., Chen X., Marshak D.R., and Prives C. 1993. A proteolytic fragment from the central region of p53 has marked sequence-specific DNA-binding activity when generated from wild-type but not from oncogenic mutant p53 protein. Genes Dev. 7: 2565–2574.

Barlev N.A., Liu L., Chehab N.H., Mansfield K., Harris K.G., Halazonetis T.D., and Berger S.L. 2001. Acetylation of p53 activates transcription through recruitment of coactivators/histone acetyltransferases. Mol. Cell 8: 1243–1254.

Bernardi R., Scaglioni P.P., Bergmann S., Horn H.F., Vousden K.H., and Pandolfi P.P. 2004. PML regulates p53 stability by sequestering Mdm2 to the nucleolus. Nat. Cell Biol. 6: 665–672.

Bottger V., Bottger A., Howard S.F., Picksley S.M., Chene P., Garcia-Echeverria C., Hochkeppel H.K., and Lane D.P. 1996. Identification of novel mdm2 binding peptides by phage display. Oncogene 13: 2141–2147.

Brent R. and Ptashne M. 1985. A eukaryotic transcriptional activator bearing the DNA specificity of a prokaryotic repressor. Cell 43: 729–736.

Brodsky M.H., Nordstrom W., Tsang G., Kwan E., Rubin G.M., and Abrams J.M. 2000. Drosophila p53 binds a damage response element at the reaper locus. Cell 101: 103–113.

Bykov V.J., Issaeva N., Shilov A., Hultcrantz M., Pugacheva E., Chumakov P., Bergman J., Wiman K.G., and Selivanova G. 2002. Restoration of the tumor suppressor function to mutant p53 by a low-molecular-weight compound. Nat. Med. 8: 282–288.

Caelles C., Helmberg A., and Karin M. 1994. p53-dependent apoptosis in the absence of transcriptional activation of p53-target genes. Nature 370: 220–223.

Chang J., Kim D.H., Lee S.W., Choi K.Y., and Sung Y.C. 1995. Transactivation ability of p53 transcriptional activation domain is directly related to the binding affinity to TATA-binding protein. J. Biol. Chem. 270: 25014–25019.

Chen D., Ma H., Hong H., Koh S.S., Huang S.M., Schurter B.T., Aswad D.W., and Stallcup M.R. 1999. Regulation of transcription by a protein methyltransferase. Science 284: 2174–2177.

Chen J., Marechal V., and Levine A.J. 1993. Mapping of the p53 and mdm-2 interaction domains. Mol. Cell. Biol. 13: 4107–4114.

Chene P., Fuchs J., Bohn J., Garcia-Echeverria C., Furet P., and Fabbro D. 2000. A small synthetic peptide, which inhibits the p53-hdm2 interaction, stimulates the p53 pathway in tumour cell lines. J. Mol. Biol. 299: 245–253.

Chipuk J.E., Kuwana T., Bouchier-Hayes L., Droin N.M., Newmeyer D.D., Schuler M., and Green D.R. 2004. Direct activation of Bax by p53 mediates mitochondrial membrane permeabilization and apoptosis. Science 303: 1010–1014.

Cho Y., Gorina S., Jeffrey P.D., and Pavletich N.P. 1994. Crystal structure of a p53 tumor suppressor-DNA complex: Understanding tumorigenic mutations. Science 265: 346–355.

Ciciarello M., Mangiacasale R., Casenghi M., Zaira Limongi M., D'Angelo M., Soddu S., Lavia P., and Cundari E. 2001. p53 displacement from centrosomes and p53-mediated G1 arrest following transient inhibition of the mitotic spindle. J. Biol. Chem. 276: 19205–19213.

Clarke A.R., Purdie C.A., Harrison D.J., Morris R.G., Bird C.C., Hooper M.L., and Wyllie A.H. 1993. Thymocyte apoptosis induced by p53-dependent and independent pathways. Nature 362: 849–852.

Clore G.M., Omichinski J.G., Sakaguchi K., Zambrano N., Sakamoto H., Appella E., and Gronenborn A.M. 1994. High-resolution structure of the oligomerization domain of p53 by multidimensional NMR. Science 265: 386–391.

Cohen S.N., Chang A.C., Boyer H.W., and Helling R.B. 1973. Construction of biologically functional bacterial plasmids in vitro. Proc. Natl. Acad. Sci. 70: 3240–3244.

Courtois S., de Fromentel C.C., and Hainaut P. 2004. p53 protein variants: Structural and functional similarities with p63 and p73 isoforms. Oncogene 23: 631–638.

Courtois S., Verhaegh G., North S., Luciani M.G., Lassus P., Hibner U., Oren M., and Hainaut P. 2002. ΔN-p53, a natural isoform of p53 lacking the first transactivation domain, counteracts growth suppression by wild-type p53. Oncogene 21: 6722–6728.

Crook T., Marston N.J., Sara E.A., and Vousden K.H. 1994. Transcriptional activation by p53 correlates with suppression of growth but not transformation. Cell 79: 817–827.

Dameron K.M., Volpert O.V., Tainsky M.A., and Bouck N. 1994. Control of angiogenesis in fibroblasts by p53 regulation of thrombospondin-1. Science 265: 1582–1584.

Danial N.N. and Korsmeyer S.J. 2004. Cell death: Critical control points. Cell 116: 205–219.

Dash B.C. and El-Deiry W.S. 2004. Cell cycle checkpoint control mechanisms that can be disrupted in cancer. Methods Mol. Biol. 280: 99–161.

DeLeo A.B., Jay G., Appella E., Dubois G.C., Law L.W., and Old L.J. 1979. Detection of a transformation-related antigen in chemically induced sarcomas and other transformed cells of the mouse. Proc. Natl. Acad. Sci. 76: 2420–2424.

Derry W.B., Putzke A.P., and Rothman J.H. 2001. Caenorhabditis elegans p53: Role in apoptosis, meiosis, and stress resistance. Science 294: 591–595.

Donehower L.A., Harvey M., Slagle B.L., McArthur M.J., Montgomery Jr., C.A., Butel J.S., and Bradley A. 1992. Mice deficient for p53 are developmentally normal but susceptible to spontaneous tumours. Nature 356: 215–221.

Dumont P., Leu J.I., Della Pietra III, A.C., George D.L., and Murphy M. 2003. The codon 72 polymorphic variants of p53 have markedly different apoptotic potential. Nat. Genet. 33: 357–365.

Dutta A., Ruppert J.M., Aster J.C., and Winchester E. 1993. Inhibition of DNA replication factor RPA by p53. Nature 365: 79–82.

Edman P. 1959. Chemistry of amino acids and peptides. Annu. Rev. Biochem. 28: 69–96.

Eliyahu D., Michalovitz D., and Oren M. 1985. Overproduction of p53 antigen makes established cells highly tumorigenic. Nature 316: 158–160.

Eliyahu D., Raz A., Gruss P., Givol D., and Oren M. 1984. Participation of p53 cellular tumour antigen in transformation of normal embryonic cells. Nature 312: 646–649.

Erster S., Mihara M., Kim R.H., Petrenko O., and Moll U.M. 2004. 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. Mol. Cell. Biol. 24: 6728–6741.

Espinosa J.M. and Emerson B.M. 2001. Transcriptional regulation by p53 through intrinsic DNA/chromatin binding and site-directed cofactor recruitment. Mol. Cell 8: 57–69.

Fang S., Jensen J.P., Ludwig R.L., Vousden K.H., and Weissman A.M. 2000. Mdm2 is a RING finger-dependent ubiquitin protein ligase for itself and p53. J. Biol. Chem. 275: 8945–8951.

Fields S. and Jang S.K. 1990. Presence of a potent transcription activating sequence in the p53 protein. Science 249: 1046–1049.

Fields S. and Song O. 1989. A novel genetic system to detect protein-protein interaction. Nature 340: 245–246.

Finlay C.A. 1993. The mdm-2 oncogene can overcome wild-type p53 suppression of transformed cell growth. Mol. Cell. Biol. 13: 301–306.

Finlay C.A., Hinds P.W., and Levine A.J. 1989. The p53 proto-oncogene can act as a suppressor of transformation. Cell 57: 1083–1093.

Fiscella M., Ullrich S.J., Zambrano N., Shields M.T., Lin D., Lees-Miller S.P., Anderson C.W., Mercer W.E., and Appella E. 1993. Mutation of the serine 15 phosphorylation site of human p53 reduces the ability of p53 to inhibit cell cycle progression. Oncogene 8: 1519–1528.

Fischer P.M. and Lane D.P. 2004. Small-molecule inhibitors of the p53 suppressor HDM2: Have protein-protein interactions come of age as drug targets? Trends Pharmacol. Sci. 25: 343–346.

Fridman J.S. and Lowe S.W. 2003. Control of apoptosis by p53. Oncogene 22: 9030–9040.

Friedman P.N., Chen X., Bargonetti J., and Prives C. 1993. The p53 protein is an unusually shaped tetramer that binds directly to DNA. Proc. Natl. Acad. Sci. 90: 3319–3323.

Garcia-Echeverria C., Chene P., Blommers M.J., and Furet P. 2000. Discovery of potent antagonists of the interaction between human double minute 2 and tumor suppressor p53. J. Med. Chem. 43: 3205–3208.

Gasco M. and Crook T. 2003. p53 family members and chemoresistance in cancer: What we know and what we need to know. Drug Resist. Updates 6: 323–328.

Gilbert W. and Maxam A. 1973. The nucleotide sequence of the lac operator. Proc. Natl. Acad. Sci. 70: 3581–3584.

Gostissa M., Hofmann T.G., Will H., and Del Sal G. 2003. Regulation of p53 functions: Let's meet at the nuclear bodies. Curr. Opin. Cell Biol. 15: 351–357.

Graeber T.G., Peterson J.F., Tsai M., Monica K., Fornace Jr., A.J., and Giaccia A.J. 1994. Hypoxia induces accumulation of p53 protein, but activation of a G1-phase checkpoint by low-oxygen conditions is independent of p53 status. Mol. Cell. Biol. 14: 6264–6277.

Graeber T.G., Osmanian C., Jacks T., Housman D.E., Koch C.J., Lowe S.W., and Giaccia A.J. 1996. Hypoxia-mediated selection of cells with diminished apoptotic potential in solid tumours. Nature 379: 88–91.

Grossman S.R., Deato M.E., Brignone C., Chan H.M., Kung A.L., Tagami H., Nakatani Y., and Livingston D.M. 2003. Polyubiquitination of p53 by a ubiquitin ligase activity of p300. Science 300: 342–344.

Gu W. and Roeder R.G. 1997. Activation of p53 sequence-specific DNA binding by acetylation of the p53 C-terminal domain. Cell 90: 595–606.

Gu W., Shi X.L., and Roeder R.G. 1997. Synergistic activation of transcription by CBP and p53. Nature 387: 819–823.

Halazonetis T.D. and Kandil A.N. 1993. Conformational shifts propagate from the oligomerization domain of p53 to its tetrameric DNA binding domain and restore DNA binding to select p53 mutants. EMBO J. 12: 5057–5064.

Halazonetis T.D., Davis L.J., and Kandil A.N. 1993. Wild-type p53 adopts a `mutant'-like conformation when bound to DNA. EMBO J. 12: 1021–1028.

Hanawalt P.C. 2001. Controlling the efficiency of excision repair. Mutat. Res. 485: 3–13.

Haupt Y., Barak Y., and Oren M. 1996. Cell type-specific inhibition of p53-mediated apoptosis by mdm2. EMBO J. 15: 1596–1606.

Haupt Y., Maya R., Kazaz A., and Oren M. 1997. Mdm2 promotes the rapid degradation of p53. Nature 387: 296–299.

He Z., Brinton B.T., Greenblatt J., Hassell J.A., and Ingles C.J. 1993. The transactivator proteins VP16 and GAL4 bind replication factor A. Cell 73: 1223–1232.

Herskowitz I. 1987. Functional inactivation of genes by dominant negative mutations. Nature 329: 219–222.

Hollstein M., Rice K., Greenblatt M.S., Soussi T., Fuchs R., Sorlie T., Hovig E., Smith-Sorensen B., Montesano R., and Harris C.C. 1994. Database of p53 gene somatic mutations in human tumors and cell lines. Nucleic Acids Res. 22: 3551–3555.

Honda R., Tanaka H., and Yasuda H. 1997. Oncoprotein MDM2 is a ubiquitin ligase E3 for tumor suppressor p53. FEBS Lett. 420: 25–27.

Horikoshi N., Usheva A., Chen J., Levine A.J., Weinmann R., and Shenk T. 1995. Two domains of p53 interact with the TATA-binding protein, and the adenovirus 13S E1A protein disrupts the association, relieving p53-mediated transcriptional repression. Mol. Cell. Biol. 15: 227–234.

Hunter T. 1987. A thousand and one protein kinases. Cell 50: 823–829.

Hupp T.R., Sparks A., and Lane D.P. 1995. Small peptides activate the latent sequence-specific DNA binding function of p53. Cell 83: 237–245.

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.

Jackson P., Bos E., and Braithwaite A.W. 1993. Wild-type mouse p53 down-regulates transcription from different virus enhancer/promoters. Oncogene 8: 589–597.

Jay G., Khoury G., DeLeo A.B., Dippold W.G., and Old L.J. 1981. p53 transformation-related protein: Detection of an associated phosphotransferase activity. Proc. Natl. Acad. Sci. 78: 2932–2936.

Jenkins J.R., Rudge K., and Currie G.A. 1984. Cellular immortalization by a cDNA clone encoding the transformation-associated phosphoprotein p53. Nature 312: 651–654.

Jimenez G.S., Nister M., Stommel J.M., Beeche M., Barcarse E.A., Zhang X.Q., O'Gorman S., and Wahl G.M. 2000. A transactivation-deficient mouse model provides insights into Trp53 regulation and function. Nat. Genet. 26: 37–43.

Jin S., Martinek S., Joo W.S., Wortman J.R., Mirkovic N., Sali A., Yandell M.D., Pavletich N.P., Young M.W., and Levine A.J. 2000. Identification and characterization of a p53 homologue in Drosophila melanogaster. Proc. Natl. Acad. Sci. 97: 7301–7306.

Jost C.A., Marin M.C., and Kaelin Jr., W.G. 1997. p73 is a simian [correction of human] p53-related protein that can induce apoptosis. Nature 389: 191–194.

Kaghad M., Bonnet H., Yang A., Creancier L., Biscan J.C., Valent A., Minty A., Chalon P., Lelias J.M., Dumont X., et al. 1997. Monoallelically expressed gene related to p53 at 1p36, a region frequently deleted in neuroblastoma and other human cancers. Cell 90: 809–819.

Kamijo T., Weber J.D., Zambetti G., Zindy F., Roussel M.F., and Sherr C.J. 1998. Functional and physical interactions of the ARF tumor suppressor with p53 and Mdm2. Proc. Natl. Acad. Sci. 95: 8292–8297.

Kanda T., Segawa K., Ohuchi N., Mori S., and Ito Y. 1994. Stimulation of polyomavirus DNA replication by wild-type p53 through the DNA-binding site. Mol. Cell. Biol. 14: 2651–2663.

Kastan M.B., Onyekwere O., Sidransky D., Vogelstein B., and Craig R.W. 1991. Participation of p53 protein in the cellular response to DNA damage. Cancer Res. 51: 6304–6311.

Kern S.E., Kinzler K.W., Bruskin A., Jarosz D., Friedman P., Prives C., and Vogelstein B. 1991. Identification of p53 as a sequence-specific DNA-binding protein. Science 252: 1708–1711.

Kieser A., Weich H.A., Brandner G., Marme D., and Kolch W. 1994. Mutant p53 potentiates protein kinase C induction of vascular endothelial growth factor expression. Oncogene 9: 963–969.

Knudson Jr., A.G. 1971. Mutation and cancer: Statistical study of retinoblastoma. Proc. Natl. Acad. Sci. 68: 820–823.

Ko L.J. and Prives C. 1996. p53: Puzzle and paradigm. Genes Dev. 10: 1054–1072.

Komarova E.A. and Gudkov A.V. 2001. Chemoprotection from p53-dependent apoptosis: Potential clinical applications of the p53 inhibitors. Biochem. Pharmacol. 62: 657–667.

Kubbutat M.H. and Vousden K.H. 1997. Proteolytic cleavage of human p53 by calpain: A potential regulator of protein stability. Mol. Cell. Biol. 17: 460–468.

Lagger G., Doetzlhofer A., Schuettengruber B., Haidweger E., Simboeck E., Tischler J., Chiocca S., Suske G., Rotheneder H., Wintersberger E., et al. 2003. The tumor suppressor p53 and histone deacetylase 1 are antagonistic regulators of the cyclin-dependent kinase inhibitor p21/WAF1/CIP1 gene. Mol. Cell. Biol. 23: 2669–2679.

Lai Z., Yang T., Kim Y.B., Sielecki T.M., Diamond M.A., Strack P., Rolfe M., Caligiuri M., Benfield P.A., Auger K.R. et al. 2002. Differentiation of Hdm2-mediated p53 ubiquitination and Hdm2 autoubiquitination activity by small molecular weight inhibitors. Proc. Natl. Acad. Sci. 99: 14734–14739.

Lander E.S., Linton L.M., Birren B., Nusbaum C., Zody M.C., Baldwin J., Devon K., Dewar K., Doyle M., FitzHugh W., et al. 2001. Initial sequencing and analysis of the human genome. Nature 409: 860–921.

Lane D. 1992. Cancer. p53, guardian of the genome. Nature 358: 15–16.

Lane D. 2004. Curing cancer with p53. N. Engl. J. Med. 350: 2711–2712.

Lane D.P. and Crawford L.V. 1979. T antigen is bound to a host protein in SV40-transformed cells. Nature 278: 261–263.

Lane D.P. and Fischer P.M. 2004. Turning the key on p53. Nature 427: 789–790.

Langheinrich U., Hennen E., Stott G., and Vacun G. 2002. Zebrafish as a model organism for the identification and characterization of drugs and genes affecting p53 signaling. Curr. Biol. 12: 2023–2028.

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.

Lechner M.S., Makc D.H., Finicle A.B., Crook T., Vousden K.H., and Laimins L.A. 1992. Human papillomavirus E6 proteins bind p53 in vivo and abrogate p53-mediated repression of transcription. EMBO J. 11: 3045–3052.

Lees-Miller S.P., Sakaguchi K., Ullrich S.J., Appella E., and Anderson C.W. 1992. Human DNA-activated protein kinase phosphorylates serines 15 and 37 in the amino-terminal transactivation domain of human p53. Mol. Cell. Biol. 12: 5041–5049.

Leng R.P., Lin Y., Ma W., Wu H., Lemmers B., Chung S., Parant J.M., Lozano G., Hakem R., and Benchimol S. 2003. Pirh2, a p53-induced ubiquitin-protein ligase, promotes p53 degradation. Cell 112: 779–791.

Leu J.I., Dumont P., Hafey M., Murphy M.E., and George D.L. 2004. Mitochondrial p53 activates Bak and causes disruption of a Bak-Mcl1 complex. Nat. Cell Biol. 6: 443–450.

Leung K.M., Po L.S., Tsang F.C., Siu W.Y., Lau A., Ho H.T., and Poon R.Y. 2002. The candidate tumor suppressor ING1b can stabilize p53 by disrupting the regulation of p53 by MDM2. Cancer Res. 62: 4890–4893.

Leveillard T., Andera L., Bissonnette N., Schaeffer L., Bracco L., Egly J.M., and Wasylyk B. 1996. Functional interactions between p53 and the TFIIH complex are affected by tumour-associated mutations. EMBO J. 15: 1615–1624.

Levine A.J. 1997. p53, the cellular gatekeeper for growth and division. Cell 88: 323–331.

Li M., Luo J., Brooks C.L., and Gu W. 2002a. Acetylation of p53 inhibits its ubiquitination by Mdm2. J. Biol. Chem. 277: 50607–50611.

Li M., Chen D., Shiloh A., Luo J., Nikolaev A.Y., Qin J., and Gu W. 2002b. Deubiquitination of p53 by HAUSP is an important pathway for p53 stabilization. Nature 416: 648–653.

Li R. and Botchan M.R. 1993. The acidic transcriptional activation domains of VP16 and p53 bind the cellular replication protein A and stimulate in vitro BPV-1 DNA replication. Cell 73: 1207–1221.

Linzer D.I. and Levine A.J. 1979. Characterization of a 54K dalton cellular SV40 tumor antigen present in SV40-transformed cells and uninfected embryonal carcinoma cells. Cell 17: 43–52.

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.

Lohrum M.A., Ludwig R.L., Kubbutat M.H., Hanlon M., and Vousden K.H. 2003. Regulation of HDM2 activity by the ribosomal protein L11. Cancer Cell 3: 577–587.

Louria-Hayon I., Grossman T., Sionov R.V., Alsheich O., Pandolfi P.P., and Haupt Y. 2003. The promyelocytic leukemia protein protects p53 from Mdm2-mediated inhibition and degradation. J. Biol. Chem. 278: 33134–33141.

Lowe S.W. and Sherr C.J. 2003. Tumor suppression by Ink4a-Arf: Progress and puzzles. Curr. Opin. Genet. Dev. 13: 77–83.

Lowe S.W., Schmitt E.M., Smith S.W., Osborne B.A., and Jacks T. 1993. p53 is required for radiation-induced apoptosis in mouse thymocytes. Nature 362: 847–849.

Luo J., Nikolaev A.Y., Imai S., Chen D., Su F., Shiloh A., Guarente L., and Gu W. 2001. Negative control of p53 by Sir2α promotes cell survival under stress. Cell 107: 137–148.

MacLachlan T.K., Takimoto R., and El-Deiry W.S. 2002. BRCA1 directs a selective p53-dependent transcriptional response towards growth arrest and DNA repair targets. Mol. Cell. Biol. 22: 4280–4292.

Malkin D., Li F.P., Strong L.C., Fraumeni Jr., J.F., Nelson C.E., Kim D.H., Kassel J., Gryka M.A., Bischoff F.Z., Tainsky M.A., et al. 1990. Germ line p53 mutations in a familial syndrome of breast cancer, sarcomas, and other neoplasms. Science 250: 1233–1238.

Marchenko N.D., Zaika A., and Moll U.M. 2000. Death signal-induced localization of p53 protein to mitochondria. A potential role in apoptotic signaling. J. Biol. Chem. 275: 16202–16212.

Marotta C.A., Forget B.G., Weissman S.M., Verma I.M., McCaffrey R.P., and Baltimore D. 1974. Nucleotide sequences of human globin messenger RNA. Proc. Natl. Acad. Sci. 71: 2300–2304.

Marston N.J., Crook T., and Vousden K.H. 1994. Interaction of p53 with MDM2 is independent of E6 and does not mediate wild type transformation suppressor function. Oncogene 9: 2707–2716.

Marx J. 2004. Cancer research. Drug candidate bolsters cell's tumor defenses. Science 303: 23–25.

Maya R., Balass M., Kim S.T., Shkedy D., Leal J.F., Shifman O., Moas M., Buschmann T., Ronai Z., Shiloh Y., et al. 2001. ATM-dependent phosphorylation of Mdm2 on serine 395: Role in p53 activation by DNA damage. Genes Dev. 15: 1067–1077.

Meek D.W. 1994. Post-translational modification of p53. Semin. Cancer Biol. 5: 203–210.

Mendoza L., Orozco E., Rodriguez M.A., Garcia-Rivera G., Sanchez T., Garcia E., and Gariglio P. 2003. Ehp53, an Entamoeba histolytica protein, ancestor of the mammalian tumour suppressor p53. Microbiology 149: 885–893.

Midgley C.A., Owens B., Briscoe C.V., Thomas D.B., Lane D.P., and Hall P.A. 1995. Coupling between gamma irradiation, p53 induction and the apoptotic response depends upon cell type in vivo. J. Cell Sci. 108: 1843–1848.

Mihara M., Erster S., Zaika A., Petrenko O., Chittenden T., Pancoska P., and Moll U.M. 2003. p53 has a direct apoptogenic role at the mitochondria. Mol. Cell 11: 577–590.

Miller S.D., Farmer G., and Prives C. 1995. p53 inhibits DNA replication in vitro in a DNA-binding-dependent manner. Mol. Cell. Biol. 15: 6554–6560.

Miyashita T., Krajewski S., Krajewska M., Wang H.G., Lin H.K., Liebermann D.A., Hoffman B., and Reed J.C. 1994. Tumor suppressor p53 is a regulator of bcl-2 and bax gene expression in vitro and in vivo. Oncogene 9: 1799–1805.

Momand J., Zambetti G.P., Olson D.C., George D., and Levine A.J. 1992. The mdm-2 oncogene product forms a complex with the p53 protein and inhibits p53-mediated transactivation. Cell 69: 1237–1245.

Morris V.B., Brammall J., Noble J., and Reddel R. 2000. p53 localizes to the centrosomes and spindles of mitotic cells in the embryonic chick epiblast, human cell lines, and a human primary culture: An immunofluorescence study. Exp. Cell Res. 256: 122–130.

Mowat M., Cheng A., Kimura N., Bernstein A., and Benchimol S. 1985. Rearrangements of the cellular p53 gene in erythro-leukaemic cells transformed by Friend virus. Nature 314: 633–636.

Murphy M., Ahn J., Walker K.K., Hoffman W.H., Evans R.M., Levine A.J., and George D.L. 1999. Transcriptional repression by wild-type p53 utilizes histone deacetylases, mediated by interaction with mSin3a. Genes Dev. 13: 2490–2501.

Nigro J.M., Baker S.J., Preisinger A.C., Jessup J.M., Hostetter R., Cleary K., Bigner S.H., Davidson N., Baylin S., Devilee P., et al. 1989. Mutations in the p53 gene occur in diverse human tumour types. Nature 342: 705–708.

O'Brien K.P., Westerlund I., and Sonnhammer E.L. 2004. OrthoDisease: A database of human disease orthologs. Hum. Mutat. 24: 112–119.

O'Rourke R.W., Miller C.W., Kato G.J., Simon K.J., Chen D.L., Dang C.V., and Koeffler H.P. 1990. A potential transcriptional activation element in the p53 protein. Oncogene 5: 1829–1832.

Okorokov A.L., Ponchel F., and Milner J. 1997. Induced N- and C-terminal cleavage of p53: A core fragment of p53, generated by interaction with damaged DNA, promotes cleavage of the N-terminus of full-length p53, whereas ssDNA induces C-terminal cleavage of p53. EMBO J. 16: 6008–6017.

Palazzo R.E., Vogel J.M., Schnackenberg B.J., Hull D.R., and Wu X. 2000. Centrosome maturation. Curr. Top. Dev. Biol. 49: 449–470.

Pan Y. and Chen J. 2003. MDM2 promotes ubiquitination and degradation of MDMX. Mol. Cell. Biol. 23: 5113–5121.

Parada L.F., Land H., Weinberg R.A., Wolf D., and Rotter V. 1984. Cooperation between gene encoding p53 tumour antigen and ras in cellular transformation. Nature 312: 649–651.

Pariat M., Carillo S., Molinari M., Salvat C., Debussche L., Bracco L., Milner J., and Piechaczyk M. 1997. Proteolysis by calpains: A possible contribution to degradation of p53. Mol. Cell. Biol. 17: 2806–2815.

Pavletich N.P., Chambers K.A., and Pabo C.O. 1993. The DNA-binding domain of p53 contains the four conserved regions and the major mutation hot spots. Genes Dev. 7: 2556–2564.

Perry M.E., Piette J., Zawadzki J.A., Harvey D., and Levine A.J. 1993. The mdm-2 gene is induced in response to UV light in a p53-dependent manner. Proc. Natl. Acad. Sci. 90: 11623–11627.

Pickart C.M. 2001. Mechanisms underlying ubiquitination. Annu. Rev. Biochem. 70: 503–533.

Pietenpol J.A., Tokino T., Thiagalingam S., El-Deiry W.S., Kinzler K.W., and Vogelstein B. 1994. Sequence-specific transcriptional activation is essential for growth suppression by p53. Proc. Natl. Acad. Sci. 91: 1998–2002.

Querido E., Blanchette P., Yan Q., Kamura T., Morrison M., Boivin D., Kaelin W.G., Conaway R.C., Conaway J.W., and Branton P.E. 2001. Degradation of p53 by adenovirus E4orf6 and E1B55K proteins occurs via a novel mechanism involving a Cullin-containing complex. Genes Dev. 15: 3104–3117.

Raycroft L., Wu H.Y., and Lozano G. 1990. Transcriptional activation by wild-type but not transforming mutants of the p53 anti-oncogene. Science 249: 1049–1051.

Rice J.C. and Allis C.D. 2001. Histone methylation versus histone acetylation: New insights into epigenetic regulation. Curr. Opin. Cell Biol. 13: 263–273.

Rubbi C.P. and Milner J. 2003. p53 is a chromatin accessibility factor for nucleotide excision repair of DNA damage. EMBO J. 22: 975–986.

Sah V.P., Attardi L.D., Mulligan G.J., Williams B.O., Bronson R.T., and Jacks T. 1995. A subset of p53-deficient embryos exhibit exencephaly. Nat. Genet. 10: 175–180.

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.

Schumacher B., Hofmann K., Boulton S., and Gartner A. 2001. The C. elegans homolog of the p53 tumor suppressor is required for DNA damage-induced apoptosis. Curr. Biol. 11: 1722–1727.

Selivanova G., Ryabchenko L., Jansson E., Iotsova V., and Wiman K.G. 1999. Reactivation of mutant p53 through interaction of a C-terminal peptide with the core domain. Mol. Cell. Biol. 19: 3395–3402.

Seto E., Usheva A., Zambetti G.P., Momand J., Horikoshi N., Weinmann R., Levine A.J., and Shenk T. 1992. Wild-type p53 binds to the TATA-binding protein and represses transcription. Proc. Natl. Acad. Sci. 89: 12028–12032.

Shaulian E., Haviv I., Shaul Y., and Oren M. 1995. Transcriptional repression by the C-terminal domain of p53. Oncogene 10: 671–680.

Shaulian E., Zauberman A., Ginsberg D., and Oren M. 1992. Identification of a minimal transforming domain of p53: Negative dominance through abrogation of sequence-specific DNA binding. Mol. Cell. Biol. 12: 5581–5592.

Shaulian E., Zauberman A., Milner J., Davies E.A., and Oren M. 1993. Tight DNA binding and oligomerization are dispensable for the ability of p53 to transactivate target genes and suppress transformation. EMBO J. 12: 2789–2797.

Smith H., Edman P., and Owen J.A. 1962. N-Terminal amino-acids of human haptoglobins. Nature 193: 286–287.

Smith M.L., Chen I.T., Zhan Q., Bae I., Chen C.Y., Gilmer T.M., Kastan M.B., O'Connor P.M., and Fornace A.J., Jr. 1994. Interaction of the p53-regulated protein Gadd45 with proliferating cell nuclear antigen. Science 266: 1376–1380.

Snyder E.L., Meade B.R., Saenz C.C., and Dowdy S.F. 2004. Treatment of terminal peritoneal carcinomatosis by a transducible p53-activating peptide. PLoS Biol. 2: E36.

Sogame N., Kim M., and Abrams J.M. 2003. Drosophila p53 preserves genomic stability by regulating cell death. Proc. Natl. Acad. Sci. 100: 4696–4701.

Southan C. 2004. Has the yo-yo stopped? An assessment of human protein-coding gene number. Proteomics 4: 1712–1726.

Srivastava S., Zou Z.Q., Pirollo K., Blattner W., and Chang E.H. 1990. Germ-line transmission of a mutated p53 gene in a cancer-prone family with Li-Fraumeni syndrome. Nature 348: 747–749.

Stanbridge E.J. 1985. A case for human tumor-suppressor genes. Bioessays 3: 252–255.

Stukenberg P.T. 2004. Triggering p53 after cytokinesis failure. J. Cell Biol. 165: 607–608.

Sui G., Affar E.B., Shi Y., Brignone C., Wall N.R., Yin P., Donohoe M., Luke M.P., Calvo D., and Grossman S.R. 2004. Yin Yang 1 is a negative regulator of p53. Cell 117: 859–872.

Sutcliffe J.E. and Brehm A. 2004. Of flies and men; p53, a tumour suppressor. FEBS Lett. 567: 86–91.

Tritarelli A., Oricchio E., Ciciarello M., Mangiacasale R., Palena A., Lavia P., Soddu S., and Cundari E. 2004. p53 localization at centrosomes during mitosis and postmitotic checkpoint are ATM-dependent and require serine 15 phosphorylation. Mol. Biol. Cell 15: 3751–3757.

Tyner S.D., Venkatachalam S., Choi J., Jones S., Ghebranious N., Igelmann H., Lu X., Soron G., Cooper B., Brayton C., et al. 2002. p53 mutant mice that display early ageing-associated phenotypes. Nature 415: 45–53.

Ullrich S.J., Sakaguchi K., Lees-Miller S.P., Fiscella M., Mercer W.E., Anderson C.W., and Appella E. 1993. Phosphorylation at Ser-15 and Ser-392 in mutant p53 molecules from human tumors is altered compared to wild-type p53. Proc. Natl. Acad. Sci. 90: 5954–5958.

Unger T., Nau M.M., Segal S., and Minna J.D. 1992. p53: A transdominant regulator of transcription whose function is ablated by mutations occurring in human cancer. EMBO J. 11: 1383–1390.

Van Meir E.G., Polverini P.J., Chazin V.R., Su Huang H.J., de Tribolet N., and Cavenee W.K. 1994. Release of an inhibitor of angiogenesis upon induction of wild type p53 expression in glioblastoma cells. Nat. Genet. 8: 171–176.

Vassilev L.T., Vu B.T., Graves B., Carvajal D., Podlaski F., Filipovic Z., Kong N., Kammlott U., Lukacs C., Klein C., et al. 2004. In vivo activation of the p53 pathway by small-molecule antagonists of MDM2. Science 303: 844–848.

Vaziri H., Dessain S.K., Ng Eaton E., Imai S.I., Frye R.A., Pandita T.K., Guarente L., and Weinberg R.A. 2001. hSIR2(SIRT1) functions as an NAD-dependent p53 deacetylase. Cell 107: 149–159.

Venter J.C., Adams M.D., Myers E.W., Li P.W., Mural R.J., Sutton G.G., Smith H.O., Yandell M., Evans C.A., Holt R.A., et al. 2001. The sequence of the human genome. Science 291: 1304–1351.

Wang P., Reed M., Wang Y., Mayr G., Stenger J.E., Anderson M.E., Schwedes J.F., and Tegtmeyer P. 1994. p53 domains: Structure, oligomerization, and transformation. Mol. Cell. Biol. 14: 5182–5191.

Wang X.W., Yeh H., Schaeffer L., Roy R., Moncollin V., Egly J.M., Wang Z., Freidberg E.C., Evans M.K., Taffe B.G., et al. 1995. p53 modulation of TFIIH-associated nucleotide excision repair activity. Nat. Genet. 10: 188–195.

Wang Y., Reed M., Wang P., Stenger J.E., Mayr G., Anderson M.E., Schwedes J.F., and Tegtmeyer P. 1993. p53 domains: Identification and characterization of two autonomous DNA-binding regions. Genes Dev. 7: 2575–2586.

Weber J.D., Taylor L.J., Roussel M.F., Sherr C.J., and Bar-Sagi D. 1999. Nucleolar Arf sequesters Mdm2 and activates p53. Nat. Cell Biol. 1: 20–26.

Weintraub H., Hauschka S., and Tapscott S.J. 1991. The MCK enhancer contains a p53 responsive element. Proc. Natl. Acad. Sci. 88: 4570–4571.

Wosik K., Antel J., Kuhlmann T., Bruck W., Massie B., and Nalbantoglu J. 2003. Oligodendrocyte injury in multiple sclerosis: A role for p53. J. Neurochem. 85: 635–644.

Wu X., Bayle J.H., Olson D., and Levine A.J. 1993. The p53-mdm-2 autoregulatory feedback loop. Genes Dev. 7: 1126–1132.

Xiao H., Pearson A., Coulombe B., Truant R., Zhang S., Regier J.L., Triezenberg S.J., Reinberg D., Flores O., Ingles C.J., et al. 1994. Binding of basal transcription factor TFIIH to the acidic activation domains of VP16 and p53. Mol. Cell. Biol. 14: 7013–7024.

Yang A., Kaghad M., Wang Y., Gillett E., Fleming M.D., Dotsch V., Andrews N.C., Caput D., and McKeon F. 1998. p63, a p53 homolog at 3q27-29, encodes multiple products with transactivating, death-inducing, and dominant-negative activities. Mol. Cell 2: 305–316.

Zhang H., Somasundaram K., Peng Y., Tian H., Bi D., Weber B.L., and El-Deiry W.S. 1998. BRCA1 physically associates with p53 and stimulates its transcriptional activity. Oncogene 16: 1713–1721.

Zhang W., McClain C., Gau J.P., Guo X.Y., and Deisseroth A.B. 1994. Hyperphosphorylation of p53 induced by okadaic acid attenuates its transcriptional activation function. Cancer Res. 54: 4448–4453.

<<< Chapter 1 References            Chapter 3 References >>>

 
 
 

 
   
. .