1932

Abstract

, encoding the p53 transcription factor, is the most frequently mutated tumor suppressor gene across all human cancer types. While p53 has long been appreciated to induce antiproliferative cell cycle arrest, apoptosis, and senescence programs in response to diverse stress signals, various studies in recent years have revealed additional important functions for p53 that likely also contribute to tumor suppression, including roles in regulating tumor metabolism, ferroptosis, signaling in the tumor microenvironment, and stem cell self-renewal/differentiation. Not only does loss or mutation cause cancer, but hyperactive p53 also drives various pathologies, including developmental phenotypes, premature aging, neurodegeneration, and side effects of cancer therapies. These findings underscore the importance of balanced p53 activity and influence our thinking of how to best develop cancer therapies based on modulating the p53 pathway.

Loading

Article metrics loading...

/content/journals/10.1146/annurev-pathol-042320-025840
2022-01-24
2024-05-02
Loading full text...

Full text loading...

/deliver/fulltext/pathmechdis/17/1/annurev-pathol-042320-025840.html?itemId=/content/journals/10.1146/annurev-pathol-042320-025840&mimeType=html&fmt=ahah

Literature Cited

  1. 1. 
    Baker SJ, Fearon ER, Nigro JM, Hamilton SR, Preisinger AC et al. 1989. Chromosome 17 deletions and p53 gene mutations in colorectal carcinomas. Science 244:4901217–21
    [Google Scholar]
  2. 2. 
    Hollstein M, Sidransky D, Vogelstein B, Harris CC. 1991. p53 mutations in human cancers. Science 253:501549–53
    [Google Scholar]
  3. 3. 
    Valdez JM, Nichols KE, Kesserwan C 2017. Li-Fraumeni syndrome: a paradigm for the understanding of hereditary cancer predisposition. Br. J. Haematol. 176:4539–52
    [Google Scholar]
  4. 4. 
    Donehower LA, Harvey M, Slagle BL, McArthur MJ, Montgomery CA Jr. et al. 1992. Mice deficient for p53 are developmentally normal but susceptible to spontaneous tumours. Nature 356:6366215–21
    [Google Scholar]
  5. 5. 
    Jacks T, Remington L, Williams BO, Schmitt EM, Halachmi S et al. 1994. Tumor spectrum analysis in p53-mutant mice. Curr. Biol. 4:11–7
    [Google Scholar]
  6. 6. 
    Vousden KH, Prives C. 2009. Blinded by the light: the growing complexity of p53. Cell 137:3413–31
    [Google Scholar]
  7. 7. 
    Hu W, Feng Z, Levine AJ. 2012. The regulation of multiple p53 stress responses is mediated through MDM2. Genes Cancer 3:3–4199–208
    [Google Scholar]
  8. 8. 
    Bieging KT, Mello SS, Attardi LD. 2014. Unravelling mechanisms of p53-mediated tumour suppression. Nat. Rev. Cancer 14:5359–70
    [Google Scholar]
  9. 9. 
    Kruiswijk F, Labuschagne CF, Vousden KH. 2015. p53 in survival, death and metabolic health: a lifeguard with a licence to kill. Nat. Rev. Mol. Cell Biol. 16:7393–405
    [Google Scholar]
  10. 10. 
    Bieging KT, Attardi LD. 2015. Cancer: a piece of the p53 puzzle. Nature 520:754537–38
    [Google Scholar]
  11. 11. 
    Munoz-Fontela C, Mandinova A, Aaronson SA, Lee SW. 2016. Emerging roles of p53 and other tumour-suppressor genes in immune regulation. Nat. Rev. Immunol. 16:12741–50
    [Google Scholar]
  12. 12. 
    Spike BT, Wahl GM. 2011. p53, stem cells, and reprogramming: tumor suppression beyond guarding the genome. Genes Cancer 2:4404–19
    [Google Scholar]
  13. 13. 
    Sabapathy K, Lane DP. 2018. Therapeutic targeting of p53: All mutants are equal, but some mutants are more equal than others. Nat. Rev. Clin. Oncol. 15:113–30
    [Google Scholar]
  14. 14. 
    Bowen ME, Attardi LD. 2019. The role of p53 in developmental syndromes. J. Mol. Cell Biol. 11:3200–11
    [Google Scholar]
  15. 15. 
    Donehower LA. 2009. Using mice to examine p53 functions in cancer, aging, and longevity. Cold Spring Harb. Perspect. Biol. 1:6a001081
    [Google Scholar]
  16. 16. 
    Szybinska A, Lesniak W. 2017. P53 dysfunction in neurodegenerative diseases—the cause or effect of pathological changes?. Aging Dis. 8:4506–18
    [Google Scholar]
  17. 17. 
    Gudkov AV, Komarova EA. 2003. The role of p53 in determining sensitivity to radiotherapy. Nat. Rev. Cancer 3:2117–29
    [Google Scholar]
  18. 18. 
    Brady CA, Jiang D, Mello SS, Johnson TM, Jarvis LA et al. 2011. Distinct p53 transcriptional programs dictate acute DNA-damage responses and tumor suppression. Cell 145:4571–83
    [Google Scholar]
  19. 19. 
    Jiang D, Brady CA, Johnson TM, Lee EY, Park EJ et al. 2011. Full p53 transcriptional activation potential is dispensable for tumor suppression in diverse lineages. PNAS 108:4117123–28
    [Google Scholar]
  20. 20. 
    Mello SS, Valente LJ, Raj N, Seoane JA, Flowers BM et al. 2017. A p53 super-tumor suppressor reveals a tumor suppressive p53-Ptpn14-Yap axis in pancreatic cancer. Cancer Cell 32:4460–73.e6
    [Google Scholar]
  21. 21. 
    Raj N, Attardi LD. 2017. The transactivation domains of the p53 protein. Cold Spring Harb. . Perspect. Med. 7:1a026047
    [Google Scholar]
  22. 22. 
    Yang A, Kaghad M, Caput D, McKeon F 2002. On the shoulders of giants: p63, p73 and the rise of p53. Trends Genet. 18:290–95
    [Google Scholar]
  23. 23. 
    Koster MI, Roop DR. 2004. The role of p63 in development and differentiation of the epidermis. J. Dermatol. Sci. 34:13–9
    [Google Scholar]
  24. 24. 
    Jackson PK, Attardi LD. 2016. p73 and FoxJ1: programming multiciliated epithelia. Trends Cell Biol. 26:4239–40
    [Google Scholar]
  25. 25. 
    Lane DP. 1992. p53, guardian of the genome. Nature 358:638115–16
    [Google Scholar]
  26. 26. 
    el-Deiry WS, Tokino T, Velculescu VE, Levy DB, Parsons R et al. 1993. WAF1, a potential mediator of p53 tumor suppression. Cell 75:4817–25
    [Google Scholar]
  27. 27. 
    Brugarolas J, Chandrasekaran C, Gordon JI, Beach D, Jacks T et al. 1995. Radiation-induced cell cycle arrest compromised by p21 deficiency. Nature 377:6549552–57
    [Google Scholar]
  28. 28. 
    Deng C, Zhang P, Harper JW, Elledge SJ, Leder P. 1995. Mice lacking p21CIP1/WAF1 undergo normal development, but are defective in G1 checkpoint control. Cell 82:4675–84
    [Google Scholar]
  29. 29. 
    Barboza JA, Liu G, Ju Z, El-Naggar AK, Lozano G. 2006. p21 delays tumor onset by preservation of chromosomal stability. PNAS 103:5219842–47
    [Google Scholar]
  30. 30. 
    Hollander MC, Sheikh MS, Bulavin DV, Lundgren K, Augeri-Henmueller L et al. 1999. Genomic instability in Gadd45a-deficient mice. Nat. Genet. 23:2176–84
    [Google Scholar]
  31. 31. 
    Doumont G, Martoriati A, Beekman C, Bogaerts S, Mee PJ et al. 2005. G1 checkpoint failure and increased tumor susceptibility in mice lacking the novel p53 target Ptprv. EMBO J. 24:173093–103
    [Google Scholar]
  32. 32. 
    Schmitt CA, Fridman JS, Yang M, Baranov E, Hoffman RM et al. 2002. Dissecting p53 tumor suppressor functions in vivo. Cancer Cell 1:3289–98
    [Google Scholar]
  33. 33. 
    Michalak EM, Jansen ES, Happo L, Cragg MS, Tai L et al. 2009. Puma and to a lesser extent Noxa are suppressors of Myc-induced lymphomagenesis. Cell Death Differ. 16:5684–96
    [Google Scholar]
  34. 34. 
    Garrison SP, Jeffers JR, Yang C, Nilsson JA, Hall MA et al. 2008. Selection against PUMA gene expression in Myc-driven B-cell lymphomagenesis. Mol. Cell. Biol. 28:175391–402
    [Google Scholar]
  35. 35. 
    Hemann MT, Zilfou JT, Zhao Z, Burgess DJ, Hannon GJ et al. 2004. Suppression of tumorigenesis by the p53 target PUMA. PNAS 101:259333–38
    [Google Scholar]
  36. 36. 
    Symonds H, Krall L, Remington L, Saenz-Robles M, Lowe S et al. 1994. p53-dependent apoptosis suppresses tumor growth and progression in vivo. Cell 78:4703–11
    [Google Scholar]
  37. 37. 
    Lu X, Yang C, Yin C, Van Dyke T, Simin K. 2011. Apoptosis is the essential target of selective pressure against p53, whereas loss of additional p53 functions facilitates carcinoma progression. Mol. Cancer Res. 9:4430–39
    [Google Scholar]
  38. 38. 
    Christophorou MA, Ringshausen I, Finch AJ, Swigart LB, Evan GI. 2006. The pathological response to DNA damage does not contribute to p53-mediated tumour suppression. Nature 443:7108214–17
    [Google Scholar]
  39. 39. 
    Hinkal G, Parikh N, Donehower LA. 2009. Timed somatic deletion of p53 in mice reveals age-associated differences in tumor progression. PLOS ONE 4:8e6654
    [Google Scholar]
  40. 40. 
    Li T, Kon N, Jiang L, Tan M, Ludwig T et al. 2012. Tumor suppression in the absence of p53-mediated cell-cycle arrest, apoptosis, and senescence. Cell 149:61269–83
    [Google Scholar]
  41. 41. 
    Valente LJ, Gray DH, Michalak EM, Pinon-Hofbauer J, Egle A et al. 2013. p53 efficiently suppresses tumor development in the complete absence of its cell-cycle inhibitory and proapoptotic effectors p21, Puma, and Noxa. Cell Rep. 3:51339–45
    [Google Scholar]
  42. 42. 
    Mello SS, Attardi LD. 2018. Deciphering p53 signaling in tumor suppression. Curr. Opin. Cell Biol. 51:65–72
    [Google Scholar]
  43. 43. 
    Vander Heiden MG, Cantley LC, Thompson CB. 2009. Understanding the Warburg effect: the metabolic requirements of cell proliferation. Science 324:59301029–3344.
    [Google Scholar]
  44. 44. 
    Moon SH, Huang CH, Houlihan SL, Regunath K, Freed-Pastor WA et al. 2019. p53 represses the mevalonate pathway to mediate tumor suppression. Cell 176:3564–80.e19
    [Google Scholar]
  45. 45. 
    Jiang L, Kon N, Li T, Wang SJ, Su T et al. 2015. Ferroptosis as a p53-mediated activity during tumour suppression. Nature 520:754557–62
    [Google Scholar]
  46. 46. 
    Wang SJ, Li D, Ou Y, Jiang L, Chen Y et al. 2016. Acetylation is crucial for p53-mediated ferroptosis and tumor suppression. Cell Rep. 17:2366–73
    [Google Scholar]
  47. 47. 
    Jennis M, Kung CP, Basu S, Budina-Kolomets A, Leu JI et al. 2016. An African-specific polymorphism in the TP53 gene impairs p53 tumor suppressor function in a mouse model. Genes Dev. 30:8918–30
    [Google Scholar]
  48. 48. 
    Tarangelo A, Magtanong L, Bieging-Rolett KT, Li Y, Ye J et al. 2018. p53 suppresses metabolic stress-induced ferroptosis in cancer cells. Cell Rep. 22:3569–75
    [Google Scholar]
  49. 49. 
    Valente LJ, Tarangelo A, Li AM, Naciri M, Raj N et al. 2020. p53 deficiency triggers dysregulation of diverse cellular processes in physiological oxygen. J. Cell Biol. 219:11e201908212
    [Google Scholar]
  50. 50. 
    Xie Y, Zhu S, Song X, Sun X, Fan Y et al. 2017. The tumor suppressor p53 limits ferroptosis by blocking DPP4 activity. Cell Rep. 20:71692–704
    [Google Scholar]
  51. 51. 
    Chu B, Kon N, Chen D, Li T, Liu T et al. 2019. ALOX12 is required for p53-mediated tumour suppression through a distinct ferroptosis pathway. Nat. Cell Biol. 21:5579–91
    [Google Scholar]
  52. 52. 
    Xue W, Zender L, Miething C, Dickins RA, Hernando E et al. 2007. Senescence and tumour clearance is triggered by p53 restoration in murine liver carcinomas. Nature 445:7128656–60
    [Google Scholar]
  53. 53. 
    Wellenstein MD, Coffelt SB, Duits DEM, van Miltenburg MH, Slagter M et al. 2019. Loss of p53 triggers WNT-dependent systemic inflammation to drive breast cancer metastasis. Nature 572:7770538–42
    [Google Scholar]
  54. 54. 
    Blagih J, Zani F, Chakravarty P, Hennequart M, Pilley S et al. 2020. Cancer-specific loss of p53 leads to a modulation of myeloid and T cell responses. Cell Rep. 30:2481–96.e6
    [Google Scholar]
  55. 55. 
    Krizhanovsky V, Lowe SW. 2009. Stem cells: the promises and perils of p53. Nature 460:72591085–86
    [Google Scholar]
  56. 56. 
    Choi YJ, Lin CP, Ho JJ, He X, Okada N et al. 2011. miR-34 miRNAs provide a barrier for somatic cell reprogramming. Nat. Cell Biol. 13:111353–60
    [Google Scholar]
  57. 57. 
    Jain AK, Allton K, Iacovino M, Mahen E, Milczarek RJ et al. 2012. p53 regulates cell cycle and micro-RNAs to promote differentiation of human embryonic stem cells. PLOS Biol. 10:2e1001268
    [Google Scholar]
  58. 58. 
    Mizuno H, Spike BT, Wahl GM, Levine AJ. 2010. Inactivation of p53 in breast cancers correlates with stem cell transcriptional signatures. PNAS 107:5222745–50
    [Google Scholar]
  59. 59. 
    Boutelle AM, Attardi LD. 2021. p53 and tumor suppression: It takes a network. Trends Cell Biol. 31:4298–310
    [Google Scholar]
  60. 60. 
    Bieging-Rolett KT, Kaiser AM, Morgens DW, Boutelle AM, Seoane JA et al. 2020. Zmat3 is a key splicing regulator in the p53 tumor suppression program. Mol. Cell 80:3452–69.e9
    [Google Scholar]
  61. 61. 
    Janic A, Valente LJ, Wakefield MJ, Di Stefano L, Milla L et al. 2018. DNA repair processes are critical mediators of p53-dependent tumor suppression. Nat. Med. 24:7947–53
    [Google Scholar]
  62. 62. 
    Muys BR, Anastasakis DG, Claypool D, Pongor L, Li XL et al. 2021. The p53-induced RNA-binding protein ZMAT3 is a splicing regulator that inhibits the splicing of oncogenic CD44 variants in colorectal carcinoma. Genes Dev. 35:1–2102–16
    [Google Scholar]
  63. 63. 
    Halevy O, Michalovitz D, Oren M 1990. Different tumor-derived p53 mutants exhibit distinct biological activities. Science 250:4977113–16
    [Google Scholar]
  64. 64. 
    Dittmer D, Pati S, Zambetti G, Chu S, Teresky AK et al. 1993. Gain of function mutations in p53. Nat. Genet. 4:142–46
    [Google Scholar]
  65. 65. 
    Olive KP, Tuveson DA, Ruhe ZC, Yin B, Willis NA et al. 2004. Mutant p53 gain of function in two mouse models of Li-Fraumeni syndrome. Cell 119:6847–60
    [Google Scholar]
  66. 66. 
    Lang GA, Iwakuma T, Suh YA, Liu G, Rao VA et al. 2004. Gain of function of a p53 hot spot mutation in a mouse model of Li-Fraumeni syndrome. Cell 119:6861–72
    [Google Scholar]
  67. 67. 
    Hanel W, Marchenko N, Xu S, Yu SX, Weng W et al. 2013. Two hot spot mutant p53 mouse models display differential gain of function in tumorigenesis. Cell Death Differ. 20:7898–909
    [Google Scholar]
  68. 68. 
    Bouaoun L, Sonkin D, Ardin M, Hollstein M, Byrnes G et al. 2016. TP53 variations in human cancers: new lessons from the IARC TP53 Database and genomics data. Hum. Mutat. 37:9865–76
    [Google Scholar]
  69. 69. 
    Sabapathy K. 2015. The contrived mutant p53 oncogene—beyond loss of functions. Front. Oncol. 5:276
    [Google Scholar]
  70. 70. 
    Kim MP, Lozano G. 2018. Mutant p53 partners in crime. Cell Death Differ. 25:1161–68
    [Google Scholar]
  71. 71. 
    Aubrey BJ, Janic A, Chen Y, Chang C, Lieschke EC et al. 2018. Mutant TRP53 exerts a target gene-selective dominant-negative effect to drive tumor development. Genes Dev. 32:21–221420–29
    [Google Scholar]
  72. 72. 
    Boettcher S, Miller PG, Sharma R, McConkey M, Leventhal M et al. 2019. A dominant-negative effect drives selection of TP53 missense mutations in myeloid malignancies. Science 365:6453599–604
    [Google Scholar]
  73. 73. 
    Brittany M, Flowers HX, Mulligan AS, Hanson KJ, Seoane JA et al. 2021. Cell of origin influences pancreatic cancer subtype. Cancer Discovery 11:660–77
    [Google Scholar]
  74. 74. 
    Montes de Oca Luna R, Wagner DS, Lozano G. 1995. Rescue of early embryonic lethality in mdm2-deficient mice by deletion of p53. Nature 378:6553203–6
    [Google Scholar]
  75. 75. 
    Parant J, Chavez-Reyes A, Little NA, Yan W, Reinke V et al. 2001. Rescue of embryonic lethality in Mdm4-null mice by loss of Trp53 suggests a nonoverlapping pathway with MDM2 to regulate p53. Nat. Genet. 29:192–95
    [Google Scholar]
  76. 76. 
    Migliorini D, Lazzerini Denchi E, Danovi D, Jochemsen A, Capillo M et al. 2002. Mdm4 (Mdmx) regulates p53-induced growth arrest and neuronal cell death during early embryonic mouse development. Mol. Cell. Biol. 22:155527–38
    [Google Scholar]
  77. 77. 
    Terzian T, Wang Y, Van Pelt CS, Box NF, Travis EL et al. 2007. Haploinsufficiency of Mdm2 and Mdm4 in tumorigenesis and development. Mol. Cell. Biol. 27:155479–85
    [Google Scholar]
  78. 78. 
    Hamard PJ, Barthelery N, Hogstad B, Mungamuri SK, Tonnessen CA et al. 2013. The C terminus of p53 regulates gene expression by multiple mechanisms in a target- and tissue-specific manner in vivo. Genes Dev. 27:171868–85
    [Google Scholar]
  79. 79. 
    Simeonova I, Jaber S, Draskovic I, Bardot B, Fang M et al. 2013. Mutant mice lacking the p53 C-terminal domain model telomere syndromes. Cell Rep. 3:62046–58
    [Google Scholar]
  80. 80. 
    Van Nostrand JL, Brady CA, Jung H, Fuentes DR, Kozak MM et al. 2014. Inappropriate p53 activation during development induces features of CHARGE syndrome. Nature 514:7521228–32
    [Google Scholar]
  81. 81. 
    Zentner GE, Layman WS, Martin DM, Scacheri PC 2010. Molecular and phenotypic aspects of CHD7 mutation in CHARGE syndrome. Am. J. Med. Genet. A 152A:674–8682
    [Google Scholar]
  82. 82. 
    Bowen ME, McClendon J, Long HK, Sorayya A, Van Nostrand JL et al. 2019. The spatiotemporal pattern and intensity of p53 activation dictates phenotypic diversity in p53-driven developmental syndromes. Dev. Cell 50:2212–28.e6
    [Google Scholar]
  83. 83. 
    Bowen ME, Mulligan AS, Sorayya A, Attardi LD. 2021. Puma- and Caspase9-mediated apoptosis is dispensable for p53-driven neural crest-based developmental defects. Cell Death Differ. 28:2083–94
    [Google Scholar]
  84. 84. 
    Pant V, Xiong S, Chau G, Tsai K, Shetty G et al. 2016. Distinct downstream targets manifest p53-dependent pathologies in mice. Oncogene 35:445713–21
    [Google Scholar]
  85. 85. 
    McGowan KA, Li JZ, Park CY, Beaudry V, Tabor HK et al. 2008. Ribosomal mutations cause p53-mediated dark skin and pleiotropic effects. Nat. Genet. 40:8963–70
    [Google Scholar]
  86. 86. 
    Toki T, Yoshida K, Wang R, Nakamura S, Maekawa T et al. 2018. De novo mutations activating germline TP53 in an inherited bone-marrow-failure syndrome. Am. J. Hum. Genet. 103:3440–47
    [Google Scholar]
  87. 87. 
    Draptchinskaia N, Gustavsson P, Andersson B, Pettersson M, Willig T-N et al. 1999. The gene encoding ribosomal protein S19 is mutated in Diamond-Blackfan anaemia. Nat. Genet. 21:169–75
    [Google Scholar]
  88. 88. 
    Barlow JL, Drynan LF, Hewett DR, Holmes LR, Lorenzo-Abalde S et al. 2010. A p53-dependent mechanism underlies macrocytic anemia in a mouse model of human 5q- syndrome. Nat. Med. 16:159–66
    [Google Scholar]
  89. 89. 
    Boultwood J, Pellagatti A, Wainscoat JS 2012. Haploinsufficiency of ribosomal proteins and p53 activation in anemia: Diamond-Blackfan anemia and the 5q− syndrome. Adv. Biol. Regul. 52:1196–203
    [Google Scholar]
  90. 90. 
    Tyner SD, Venkatachalam S, Choi J, Jones S, Ghebranious N et al. 2002. p53 mutant mice that display early ageing-associated phenotypes. Nature 415:686745–53
    [Google Scholar]
  91. 91. 
    Maier B, Gluba W, Bernier B, Turner T, Mohammad K et al. 2004. Modulation of mammalian life span by the short isoform of p53. Genes Dev. 18:3306–19
    [Google Scholar]
  92. 92. 
    Liu D, Ou L, Clemenson G, Chao C, Lutske ME et al. 2010. Puma is required for p53-induced depletion of adult stem cells. Nat. Cell Biol. 12:993–98
    [Google Scholar]
  93. 93. 
    Lessel D, Wu D, Trujillo C, Ramezani T, Lessel I et al. 2017. Dysfunction of the MDM2/p53 axis is linked to premature aging. J. Clin. Investig. 127:103598–608
    [Google Scholar]
  94. 94. 
    Maor-Nof M, Shipony Z, Lopez-Gonzalez R, Nakayama L, Zhang YJ et al. 2021. p53 is a central regulator driving neurodegeneration caused by C9orf72 poly(PR). Cell 184:3689–708.e20
    [Google Scholar]
  95. 95. 
    Komarova EA, Christov K, Faerman AI, Gudkov AV. 2000. Different impact of p53 and p21 on the radiation response of mouse tissues. Oncogene 19:333791–98
    [Google Scholar]
  96. 96. 
    Wade M, Li YC, Wahl GM. 2013. MDM2, MDMX and p53 in oncogenesis and cancer therapy. Nat. Rev. Cancer 13:283–96
    [Google Scholar]
  97. 97. 
    Kussie PH, Gorina S, Marechal V, Elenbaas B, Moreau J et al. 1996. Structure of the MDM2 oncoprotein bound to the p53 tumor suppressor transactivation domain. Science 274:5289948–53
    [Google Scholar]
  98. 98. 
    Vassilev LT, Vu BT, Graves B, Carvajal D, Podlaski F et al. 2004. In vivo activation of the p53 pathway by small-molecule antagonists of MDM2. Science 303:5659844–48
    [Google Scholar]
  99. 99. 
    Patton JT, Mayo LD, Singhi AD, Gudkov AV, Stark GR et al. 2006. Levels of HdmX expression dictate the sensitivity of normal and transformed cells to Nutlin-3. Cancer Res. 66:63169–76
    [Google Scholar]
  100. 100. 
    Hu B, Gilkes DM, Farooqi B, Sebti SM, Chen J. 2006. MDMX overexpression prevents p53 activation by the MDM2 inhibitor Nutlin. J. Biol. Chem. 281:4433030–35
    [Google Scholar]
  101. 101. 
    Burgess A, Chia KM, Haupt S, Thomas D, Haupt Y, Lim E 2016. Clinical overview of MDM2/X-targeted therapies. Front. Oncol. 6:7
    [Google Scholar]
  102. 102. 
    Mullard A. 2020. p53 programmes plough on. Nat. Rev. Drug Discov. 19:8497–500
    [Google Scholar]
  103. 103. 
    Herman AG, Hayano M, Poyurovsky MV, Shimada K, Skouta R et al. 2011. Discovery of Mdm2-MdmX E3 ligase inhibitors using a cell-based ubiquitination assay. Cancer Discov. 1:4312–25
    [Google Scholar]
  104. 104. 
    Pellegrino M, Mancini F, Luca R, Coletti A, Giacche N et al. 2015. Targeting the MDM2/MDM4 interaction interface as a promising approach for p53 reactivation therapy. Cancer Res. 75:214560–72
    [Google Scholar]
  105. 105. 
    Huang M, Zhang H, Liu T, Tian D, Gu L et al. 2013. Triptolide inhibits MDM2 and induces apoptosis in acute lymphoblastic leukemia cells through a p53-independent pathway. Mol. Cancer Ther. 12:2184–94
    [Google Scholar]
  106. 106. 
    Carvajal LA, Neriah DB, Senecal A, Benard L, Thiruthuvanathan V et al. 2018. Dual inhibition of MDMX and MDM2 as a therapeutic strategy in leukemia. Sci. Transl. Med. 10:436eaao3003
    [Google Scholar]
  107. 107. 
    Ringshausen I, O'Shea CC, Finch AJ, Swigart LB, Evan GI 2006. Mdm2 is critically and continuously required to suppress lethal p53 activity in vivo. Cancer Cell 10:6501–14
    [Google Scholar]
  108. 108. 
    Hupp TR, Sparks A, Lane DP. 1995. Small peptides activate the latent sequence-specific DNA binding function of p53. Cell 83:2237–45
    [Google Scholar]
  109. 109. 
    Selivanova G, Iotsova V, Okan I, Fritsche M, Strom M et al. 1997. Restoration of the growth suppression function of mutant p53 by a synthetic peptide derived from the p53 C-terminal domain. Nat. Med. 3:6632–38
    [Google Scholar]
  110. 110. 
    Bykov VJ, Issaeva N, Shilov A, Hultcrantz M, Pugacheva E et al. 2002. Restoration of the tumor suppressor function to mutant p53 by a low-molecular-weight compound. Nat. Med. 8:3282–88
    [Google Scholar]
  111. 111. 
    Lambert JM, Gorzov P, Veprintsev DB, Soderqvist M, Segerback D et al. 2009. PRIMA-1 reactivates mutant p53 by covalent binding to the core domain. Cancer Cell 15:5376–88
    [Google Scholar]
  112. 112. 
    Lehmann S, Bykov VJ, Ali D, Andren O, Cherif H et al. 2012. Targeting p53 in vivo: a first-in-human study with p53-targeting compound APR-246 in refractory hematologic malignancies and prostate cancer. J. Clin. Oncol. 30:293633–39
    [Google Scholar]
  113. 113. 
    Bykov VJ, Issaeva N, Selivanova G, Wiman KG. 2002. Mutant p53-dependent growth suppression distinguishes PRIMA-1 from known anticancer drugs: a statistical analysis of information in the National Cancer Institute database. Carcinogenesis 23:122011–18
    [Google Scholar]
  114. 114. 
    Menichini P, Monti P, Speciale A, Cutrona G, Matis S et al. 2021. Antitumor effects of PRIMA-1 and PRIMA-1Met (APR246) in hematological malignancies: still a mutant P53-dependent affair?. Cells 10:198
    [Google Scholar]
  115. 115. 
    Yu X, Vazquez A, Levine AJ, Carpizo DR. 2012. Allele-specific p53 mutant reactivation. Cancer Cell 21:5614–25
    [Google Scholar]
  116. 116. 
    Butler JS, Loh SN. 2003. Structure, function, and aggregation of the zinc-free form of the p53 DNA binding domain. Biochemistry 42:82396–403
    [Google Scholar]
  117. 117. 
    Loh SN. 2010. The missing zinc: p53 misfolding and cancer. Metallomics 2:7442–49
    [Google Scholar]
  118. 118. 
    Yu X, Kogan S, Chen Y, Tsang AT, Withers T et al. 2018. Zinc metallochaperones reactivate mutant p53 using an ON/OFF switch mechanism: a new paradigm in cancer therapeutics. Clin. Cancer Res. 24:184505–17
    [Google Scholar]
  119. 119. 
    Li D, Marchenko ND, Moll UM. 2011. SAHA shows preferential cytotoxicity in mutant p53 cancer cells by destabilizing mutant p53 through inhibition of the HDAC6-Hsp90 chaperone axis. Cell Death Differ. 18:121904–13
    [Google Scholar]
  120. 120. 
    Alexandrova EM, Yallowitz AR, Li D, Xu S, Schulz R et al. 2015. Improving survival by exploiting tumour dependence on stabilized mutant p53 for treatment. Nature 523:7560352–56
    [Google Scholar]
  121. 121. 
    Blagosklonny MV, Toretsky J, Neckers L. 1995. Geldanamycin selectively destabilizes and conformationally alters mutated p53. Oncogene 11:5933–39
    [Google Scholar]
  122. 122. 
    Blagosklonny MV, Toretsky J, Bohen S, Neckers L 1996. Mutant conformation of p53 translated in vitro or in vivo requires functional HSP90. PNAS 93:168379–83
    [Google Scholar]
  123. 123. 
    Whitesell L, Sutphin P, An WG, Schulte T, Blagosklonny MV et al. 1997. Geldanamycin-stimulated destabilization of mutated p53 is mediated by the proteasome in vivo. Oncogene 14:232809–16
    [Google Scholar]
  124. 124. 
    Li D, Marchenko ND, Schulz R, Fischer V, Velasco-Hernandez T et al. 2011. Functional inactivation of endogenous MDM2 and CHIP by HSP90 causes aberrant stabilization of mutant p53 in human cancer cells. Mol. Cancer Res. 9:5577–88
    [Google Scholar]
  125. 125. 
    Shrestha L, Bolaender A, Patel HJ, Taldone T. 2016. Heat shock protein (HSP) drug discovery and development: targeting heat shock proteins in disease. Curr. Top. Med. Chem. 16:252753–64
    [Google Scholar]
  126. 126. 
    Wang Q, Fan S, Eastman A, Worland PJ, Sausville EA et al. 1996. UCN-01: a potent abrogator of G2 checkpoint function in cancer cells with disrupted p53. J. Natl. Cancer Inst. 88:14956–65
    [Google Scholar]
  127. 127. 
    Ma CX, Cai S, Li S, Ryan CE, Guo Z et al. 2012. Targeting Chk1 in p53-deficient triple-negative breast cancer is therapeutically beneficial in human-in-mouse tumor models. J. Clin. Investig. 122:41541–52
    [Google Scholar]
  128. 128. 
    Sur S, Pagliarini R, Bunz F, Rago C, Diaz LA Jr. et al. 2009. A panel of isogenic human cancer cells suggests a therapeutic approach for cancers with inactivated p53. PNAS 106:103964–69
    [Google Scholar]
  129. 129. 
    Wang Y, Li J, Booher RN, Kraker A, Lawrence T et al. 2001. Radiosensitization of p53 mutant cells by PD0166285, a novel G2 checkpoint abrogator. Cancer Res. 61:228211–17
    [Google Scholar]
  130. 130. 
    Leijen S, van Geel RM, Sonke GS, de Jong D, Rosenberg EH et al. 2016. Phase II study of WEE1 inhibitor AZD1775 plus carboplatin in patients with TP53-mutated ovarian cancer refractory or resistant to first-line therapy within 3 months. J. Clin. Oncol. 34:364354–61
    [Google Scholar]
  131. 131. 
    Gutteridge RE, Ndiaye MA, Liu X, Ahmad N. 2016. Plk1 inhibitors in cancer therapy: from laboratory to clinics. Mol. Cancer Ther. 15:71427–35
    [Google Scholar]
  132. 132. 
    Angius G, Tomao S, Stati V, Vici P, Bianco V, Tomao F. 2020. Prexasertib, a checkpoint kinase inhibitor: from preclinical data to clinical development. Cancer Chemother. Pharmacol. 85:19–20
    [Google Scholar]
  133. 133. 
    Thompson JM, Nguyen QH, Singh M, Razorenova OV. 2015. Approaches to identifying synthetic lethal interactions in cancer. Yale J. Biol. Med. 88:2145–55
    [Google Scholar]
  134. 134. 
    Sobhani N, D'Angelo A, Wang X, Young KH, Generali D, Li Y 2020. Mutant p53 as an antigen in cancer immunotherapy. Int. J. Mol. Sci. 21:114087
    [Google Scholar]
  135. 135. 
    Low L, Goh A, Koh J, Lim S, Wang CI. 2019. Targeting mutant p53-expressing tumours with a T cell receptor-like antibody specific for a wild-type antigen. Nat. Commun. 10:15382
    [Google Scholar]
  136. 136. 
    Schuler PJ, Harasymczuk M, Visus C, Deleo A, Trivedi S et al. 2014. Phase I dendritic cell p53 peptide vaccine for head and neck cancer. Clin. Cancer Res. 20:92433–44
    [Google Scholar]
  137. 137. 
    Hsiue EH, Wright KM, Douglass J, Hwang MS, Mog BJ et al. 2021. Targeting a neoantigen derived from a common TP53 mutation. Science 371:6533eabc8697
    [Google Scholar]
  138. 138. 
    Levine AJ. 2020. p53: 800 million years of evolution and 40 years of discovery. Nat. Rev. Cancer 20:8471–80
    [Google Scholar]
/content/journals/10.1146/annurev-pathol-042320-025840
Loading
/content/journals/10.1146/annurev-pathol-042320-025840
Loading

Data & Media loading...

  • Article Type: Review Article
This is a required field
Please enter a valid email address
Approval was a Success
Invalid data
An Error Occurred
Approval was partially successful, following selected items could not be processed due to error