Opinion | Published:

Unravelling mechanisms of p53-mediated tumour suppression

Nature Reviews Cancer volume 14, pages 359370 (2014) | Download Citation

Abstract

p53 is a crucial tumour suppressor that responds to diverse stress signals by orchestrating specific cellular responses, including transient cell cycle arrest, cellular senescence and apoptosis, which are all processes associated with tumour suppression. However, recent studies have challenged the relative importance of these canonical cellular responses for p53-mediated tumour suppression and have highlighted roles for p53 in modulating other cellular processes, including metabolism, stem cell maintenance, invasion and metastasis, as well as communication within the tumour microenvironment. In this Opinion article, we discuss the roles of classical p53 functions, as well as emerging p53-regulated processes, in tumour suppression.

Access optionsAccess options

Rent or Buy article

Get time limited or full article access on ReadCube.

from$8.99

All prices are NET prices.

References

  1. 1.

    & Blinded by the light: the growing complexity of p53. Cell 137, 413–431 (2009).

  2. 2.

    & p53 at a glance. J. Cell Sci. 123, 2527–2532 (2010).

  3. 3.

    et al. Gain of function of a p53 hot spot mutation in a mouse model of Li-Fraumeni syndrome. Cell 119, 861–872 (2004).

  4. 4.

    & p53 gain-of-function cancer mutants induce genetic instability by inactivating ATM. Nature Cell Biol. 9, 573–580 (2007).

  5. 5.

    & A genetic model for colorectal tumorigenesis. Cell 61, 759–767 (1990).

  6. 6.

    et al. Germ line p53 mutations in a familial syndrome of breast cancer, sarcomas, and other neoplasms. Science 250, 1233–1238 (1990).

  7. 7.

    et al. An expression signature for p53 status in human breast cancer predicts mutation status, transcriptional effects, and patient survival. Proc. Natl Acad. Sci. USA 102, 13550–13555 (2005).

  8. 8.

    & Somatic mutations in cancer prognosis and prediction: lessons from TP53 and EGFR genes. Curr. Opin. Oncol. 23, 88–92 (2011).

  9. 9.

    , , & Inactivation of p53 in breast cancers correlates with stem cell transcriptional signatures. Proc. Natl Acad. Sci. USA 107, 22745–22750 (2010).

  10. 10.

    , , & Mutations in the p53 tumor suppressor gene: important milestones at the various steps of tumorigenesis. Genes Cancer 2, 466–474 (2011).

  11. 11.

    & The complexity of p53 modulation: emerging patterns from divergent signals. Genes Dev. 12, 2973–2983 (1998).

  12. 12.

    , & The regulation of multiple p53 stress responses is mediated through MDM2. Genes Cancer 3, 199–208 (2012).

  13. 13.

    et al. Distinct p53 transcriptional programs dictate acute DNA-damage responses and tumor suppression. Cell 145, 571–583 (2011).

  14. 14.

    et al. Tumor suppression in the absence of p53-mediated cell-cycle arrest, apoptosis, and senescence. Cell 149, 1269–1283 (2012).

  15. 15.

    et al. p53 efficiently suppresses tumor development in the complete absence of its cell-cycle inhibitory and proapoptotic effectors p21, Puma, and Noxa. Cell Rep. 3, 1339–1345 (2013).

  16. 16.

    , , & Transcriptional control of human p53-regulated genes. Nat. Rev. Mol. Cell Biol. 9, 402–412 (2008).

  17. 17.

    et al. Mice deficient for p53 are developmentally normal but susceptible to spontaneous tumours. Nature 356, 215–221 (1992).

  18. 18.

    et al. Tumor spectrum analysis in p53-mutant mice. Curr. Biol. 4, 1–7 (1994).

  19. 19.

    et al. Tumour incidence, spectrum and ploidy in mice with a large deletion in the p53 gene. Oncogene 9, 603–609 (1994).

  20. 20.

    et al. Genetic background alters the spectrum of tumors that develop in p53-deficient mice. FASEB J. 7, 938–943 (1993).

  21. 21.

    & Probing p53 biological functions through the use of genetically engineered mouse models. Mutat. Res. 576, 4–21 (2005).

  22. 22.

    & Transcriptional regulation by p53. Cold Spring Harb. Perspect Biol. 2, a000935 (2010).

  23. 23.

    , & TP53 mutations in human cancers: origins, consequences, and clinical use. Cold Spring Harb. Perspect Biol. 2, a001008 (2010).

  24. 24.

    , , & The p53 protein is an unusually shaped tetramer that binds directly to DNA. Proc. Natl Acad. Sci. USA 90, 3319–3323 (1993).

  25. 25.

    & Deconstructing p53 transcriptional networks in tumor suppression. Trends Cell Biol. 22, 97–106 (2012).

  26. 26.

    et al. Role of p53 serine 46 in p53 target gene regulation. PLoS ONE 6, e17574 (2011).

  27. 27.

    et al. Global genomic profiling reveals an extensive p53-regulated autophagy program contributing to key p53 responses. Genes Dev. 27, 1016–1031 (2013).

  28. 28.

    et al. Insights into p53 transcriptional function via genome-wide chromatin occupancy and gene expression analysis. Cell Death Differ. 19, 1992–2002 (2012).

  29. 29.

    , , & The p53QS transactivation-deficient mutant shows stress-specific apoptotic activity and induces embryonic lethality. Nature Genet. 37, 145–152 (2005).

  30. 30.

    et al. Full p53 transcriptional activation potential is dispensable for tumor suppression in diverse lineages. Proc. Natl Acad. Sci. USA 108, 17123–17128 (2011).

  31. 31.

    Cancer. p53, guardian of the genome. Nature 358, 15–16 (1992).

  32. 32.

    et al. WAF1, a potential mediator of p53 tumor suppression. Cell 75, 817–825 (1993).

  33. 33.

    , , , & Mice lacking p21CIP1/WAF1 undergo normal development, but are defective in G1 checkpoint control. Cell 82, 675–684 (1995).

  34. 34.

    et al. Radiation-induced cell cycle arrest compromised by p21 deficiency. Nature 377, 552–557 (1995).

  35. 35.

    , , & Tumor susceptibility of p21(Waf1/Cip1)-deficient mice. Cancer Res. 61, 6234–6238 (2001).

  36. 36.

    et al. Genomic instability in Gadd45a-deficient mice. Nature Genet. 23, 176–184 (1999).

  37. 37.

    et al. G1 checkpoint failure and increased tumor susceptibility in mice lacking the novel p53 target Ptprv. EMBO J. 24, 3093–3103 (2005).

  38. 38.

    et al. Role of PML in cell growth and the retinoic acid pathway. Science 279, 1547–1551 (1998).

  39. 39.

    et al. Role of promyelocytic leukemia (PML) protein in tumor suppression. J. Exp. Med. 193, 521–529 (2001).

  40. 40.

    , & Gadd45a suppresses Ras-driven mammary tumorigenesis by activation of c-Jun NH2-terminal kinase and p38 stress signaling resulting in apoptosis and senescence. Cancer Res. 66, 8448–8454 (2006).

  41. 41.

    et al. Gadd45a functions as a promoter or suppressor of breast cancer dependent on the oncogenic stress. Cancer Res. 70, 9671–9681 (2010).

  42. 42.

    et al. Gadd45a protects against UV irradiation-induced skin tumors, and promotes apoptosis and stress signaling via MAPK and p53. Cancer Res. 62, 7305–7315 (2002).

  43. 43.

    et al. Chromosome stability, in the absence of apoptosis, is critical for suppression of tumorigenesis in Trp53 mutant mice. Nature Genet. 36, 63–68 (2004).

  44. 44.

    , & Differential activation of target cellular promoters by p53 mutants with impaired apoptotic function. Mol. Cell. Biol. 16, 4952–4960 (1996).

  45. 45.

    et al. Specific loss of apoptotic but not cell-cycle arrest function in a human tumor derived p53 mutant. EMBO J. 15, 827–838 (1996).

  46. 46.

    & Transcriptional control of the proliferation cluster by the tumor suppressor p53. Mol. Biosyst. 6, 17–29 (2010).

  47. 47.

    , , , & p21 delays tumor onset by preservation of chromosomal stability. Proc. Natl Acad. Sci. USA 103, 19842–19847 (2006).

  48. 48.

    et al. p53 DNA binding cooperativity is essential for apoptosis and tumor suppression in vivo. Cell Rep. 3, 1512–1525 (2013).

  49. 49.

    et al. DNA binding cooperativity of p53 modulates the decision between cell-cycle arrest and apoptosis. Mol. Cell 38, 356–368 (2010).

  50. 50.

    & The DNA-damage response in human biology and disease. Nature 461, 1071–1078 (2009).

  51. 51.

    , , , & p53-dependent regulation of nucleotide excision repair in murine epidermis in vivo. J. Cutan. Med. Surg. 3, 16–20 (1998).

  52. 52.

    , & p53 directly enhances rejoining of DNA double-strand breaks with cohesive ends in gamma-irradiated mouse fibroblasts. Cancer Res. 59, 2562–2565 (1999).

  53. 53.

    , , , & Implication of p53 in base excision DNA repair: in vivo evidence. Oncogene 21, 731–737 (2002).

  54. 54.

    et al. Deletion of XPC leads to lung tumors in mice and is associated with early events in human lung carcinogenesis. Proc. Natl Acad. Sci. USA 102, 13200–13205 (2005).

  55. 55.

    et al. Mouse models for xeroderma pigmentosum group A and group C show divergent cancer phenotypes. Cancer Res. 68, 1347–1353 (2008).

  56. 56.

    et al. Defective nucleotide excision repair in xpc mutant mice and its association with cancer predisposition. Mutat. Res. 459, 99–108 (2000).

  57. 57.

    et al. Tumor-prone phenotype of the DDB2-deficient mice. Oncogene 24, 469–478 (2005).

  58. 58.

    et al. Telomere dysfunction suppresses spontaneous tumorigenesis in vivo by initiating p53-dependent cellular senescence. EMBO Rep. 8, 497–503 (2007).

  59. 59.

    et al. A new mouse model to explore the initiation, progression, and therapy of BRAFV600E-induced lung tumors. Genes Dev. 21, 379–384 (2007).

  60. 60.

    et al. Tumour biology: senescence in premalignant tumours. Nature 436, 642 (2005).

  61. 61.

    et al. Crucial role of p53-dependent cellular senescence in suppression of Pten-deficient tumorigenesis. Nature 436, 725–730 (2005).

  62. 62.

    et al. Restoration of p53 function leads to tumour regression in vivo. Nature 445, 661–665 (2007).

  63. 63.

    et al. Senescence and tumour clearance is triggered by p53 restoration in murine liver carcinomas. Nature 445, 656–660 (2007).

  64. 64.

    et al. Stage-specific sensitivity to p53 restoration during lung cancer progression. Nature 468, 572–575 (2010).

  65. 65.

    et al. p53-dependent apoptosis suppresses tumor growth and progression in vivo. Cell 78, 703–711 (1994).

  66. 66.

    , , & Bax suppresses tumorigenesis and stimulates apoptosis in vivo. Nature 385, 637–640 (1997).

  67. 67.

    , , , & INK4a/ARF mutations accelerate lymphomagenesis and promote chemoresistance by disabling p53. Genes Dev. 13, 2670–2677 (1999).

  68. 68.

    , , , & Disruption of the ARF-Mdm2-p53 tumor suppressor pathway in Myc-induced lymphomagenesis. Genes Dev. 13, 2658–2669 (1999).

  69. 69.

    et al. Dissecting p53 tumor suppressor functions in vivo. Cancer Cell 1, 289–298 (2002).

  70. 70.

    et al. Puma is an essential mediator of p53-dependent and -independent apoptotic pathways. Cancer Cell 4, 321–328 (2003).

  71. 71.

    , , & Bax accelerates tumorigenesis in p53-deficient mice. Cancer Res. 61, 659–665 (2001).

  72. 72.

    et al. p53- and drug-induced apoptotic responses mediated by BH3-only proteins puma and noxa. Science 302, 1036–1038 (2003).

  73. 73.

    , & Adult mice lacking the p53/p63 target gene Perp are not predisposed to spontaneous tumorigenesis but display features of ectodermal dysplasia syndromes. Cell Death Differ. 13, 1614–1618 (2006).

  74. 74.

    , , & Bax loss impairs Myc-induced apoptosis and circumvents the selection of p53 mutations during Myc-mediated lymphomagenesis. Mol. Cell. Biol. 21, 7653–7662 (2001).

  75. 75.

    et al. Puma and to a lesser extent Noxa are suppressors of Myc-induced lymphomagenesis. Cell Death Differ. 16, 684–696 (2009).

  76. 76.

    , , , & Specific requirement for Bax, not Bak, in Myc-induced apoptosis and tumor suppression in vivo. J. Biol. Chem. 281, 10890–10895 (2006).

  77. 77.

    et al. Suppression of tumorigenesis by the p53 target PUMA. Proc. Natl Acad. Sci. USA 101, 9333–9338 (2004).

  78. 78.

    et al. Loss of the p53/p63 regulated desmosomal protein Perp promotes tumorigenesis. PLoS Genet. 6, e1001168 (2010).

  79. 79.

    , & An oncogene-induced DNA damage model for cancer development. Science 319, 1352–1355 (2008).

  80. 80.

    et al. DNA damage response as a candidate anti-cancer barrier in early human tumorigenesis. Nature 434, 864–870 (2005).

  81. 81.

    et al. Activation of the DNA damage checkpoint and genomic instability in human precancerous lesions. Nature 434, 907–913 (2005).

  82. 82.

    , , , & The pathological response to DNA damage does not contribute to p53-mediated tumour suppression. Nature 443, 214–217 (2006).

  83. 83.

    , , , & Tumour biology: policing of oncogene activity by p53. Nature 443, 159 (2006).

  84. 84.

    , & Timed somatic deletion of p53 in mice reveals age-associated differences in tumor progression. PLoS ONE 4, e6654 (2009).

  85. 85.

    et al. Temporal dissection of p53 function in vitro and in vivo. Nature Genet. 37, 718–726 (2005).

  86. 86.

    et al. Selective activation of p53-mediated tumour suppression in high-grade tumours. Nature 468, 567–571 (2010).

  87. 87.

    et al. Genome-wide analysis of p53 under hypoxic conditions. Mol. Cell. Biol. 26, 3492–3504 (2006).

  88. 88.

    et al. The antioxidant function of the p53 tumor suppressor. Nature Med. 11, 1306–1313 (2005).

  89. 89.

    & Metabolic reprogramming: a cancer hallmark even warburg did not anticipate. Cancer Cell 21, 297–308 (2012).

  90. 90.

    & Metabolic regulation by p53. J. Mol. Med. 89, 237–245 (2011).

  91. 91.

    , & p53 regulates glucose metabolism through an IKK-NF-κB pathway and inhibits cell transformation. Nature Cell Biol. 10, 611–618 (2008).

  92. 92.

    , & The tumor suppressor p53 down-regulates glucose transporters GLUT1 and GLUT4 gene expression. Cancer Res. 64, 2627–2633 (2004).

  93. 93.

    et al. p53 regulates mitochondrial respiration. Science 312, 1650–1653 (2006).

  94. 94.

    et al. TIGAR, a p53-inducible regulator of glycolysis and apoptosis. Cell 126, 107–120 (2006).

  95. 95.

    , , , & Regeneration of peroxiredoxins by p53-regulated sestrins, homologs of bacterial AhpD. Science 304, 596–600 (2004).

  96. 96.

    et al. Glutaminase 2, a novel p53 target gene regulating energy metabolism and antioxidant function. Proc. Natl Acad. Sci. USA 107, 7455–7460 (2010).

  97. 97.

    Stress-responsive sestrins link p53 with redox regulation and mammalian target of rapamycin signaling. Antioxid. Redox Signal 15, 1679–1690 (2011).

  98. 98.

    & p53 regulation of metabolic pathways. Cold Spring Harb. Perspect Biol. 2, a001040 (2010).

  99. 99.

    & An overview of the molecular mechanism of autophagy. Curr. Top. Microbiol. Immunol. 335, 1–32 (2009).

  100. 100.

    & Autophagy, stress, and cancer metabolism: what doesn't kill you makes you stronger. Cold Spring Harb. Symp. Quant. Biol. 76, 389–396 (2011).

  101. 101.

    et al. DRAM, a p53-induced modulator of autophagy, is critical for apoptosis. Cell 126, 121–134 (2006).

  102. 102.

    , , & Upregulation of human autophagy-initiation kinase ULK1 by tumor suppressor p53 contributes to DNA-damage-induced cell death. Cell Death Differ. 18, 1598–1607 (2011).

  103. 103.

    et al. Extracellular adenosine sensing-a metabolic cell death priming mechanism downstream of p53. Mol. Cell 50, 394–406 (2013).

  104. 104.

    et al. Direct cell reprogramming is a stochastic process amenable to acceleration. Nature 462, 595–601 (2009).

  105. 105.

    et al. The Ink4/Arf locus is a barrier for iPS cell reprogramming. Nature 460, 1136–1139 (2009).

  106. 106.

    et al. A p53-mediated DNA damage response limits reprogramming to ensure iPS cell genomic integrity. Nature 460, 1149–1153 (2009).

  107. 107.

    et al. Immortalization eliminates a roadblock during cellular reprogramming into iPS cells. Nature 460, 1145–1148 (2009).

  108. 108.

    et al. Suppression of induced pluripotent stem cell generation by the p53-p21 pathway. Nature 460, 1132–1135 (2009).

  109. 109.

    et al. Linking the p53 tumour suppressor pathway to somatic cell reprogramming. Nature 460, 1140–1144 (2009).

  110. 110.

    et al. Mutant p53 facilitates somatic cell reprogramming and augments the malignant potential of reprogrammed cells. J. Exp. Med. 207, 2127–2140 (2010).

  111. 111.

    , , , & Multiple roles of p53-related pathways in somatic cell reprogramming and stem cell differentiation. Cancer Res. 72, 5635–5645 (2012).

  112. 112.

    et al. miR-34 miRNAs provide a barrier for somatic cell reprogramming. Nature Cell Biol. 13, 1353–1360 (2011).

  113. 113.

    , , & The emerging functions of the p53-miRNA network in stem cell biology. Cell Cycle 11, 2063–2072 (2012).

  114. 114.

    , & The p53 pathway in hematopoiesis: lessons from mouse models, implications for humans. Blood 120, 5118–5127 (2012).

  115. 115.

    et al. p53 loss promotes acute myeloid leukemia by enabling aberrant self-renewal. Genes Dev. 24, 1389–1402 (2010).

  116. 116.

    et al. p53 and Pten control neural and glioma stem/progenitor cell renewal and differentiation. Nature 455, 1129–1133 (2008).

  117. 117.

    et al. Selective inactivation of p53 facilitates mouse epithelial tumor progression without chromosomal instability. Mol. Cell. Biol. 21, 6017–6030 (2001).

  118. 118.

    et al. CKIα ablation highlights a critical role for p53 in invasiveness control. Nature 470, 409–413 (2011).

  119. 119.

    , , , & Changes in p53 expression in mouse fibroblasts can modify motility and extracellular matrix organization. Oncogene 19, 5826–5830 (2000).

  120. 120.

    , & p19Arf-p53 tumor suppressor pathway regulates cell motility by suppression of phosphoinositide 3-kinase and Rac1 GTPase activities. J. Biol. Chem. 278, 14414–14419 (2003).

  121. 121.

    & Rho family GTPases cooperate with p53 deletion to promote primary mouse embryonic fibroblast cell invasion. Oncogene 23, 5577–5585 (2004).

  122. 122.

    , , & Regulation of Cdc42-mediated morphological effects: a novel function for p53. EMBO J. 21, 2373–2382 (2002).

  123. 123.

    , , & Loss of p53 promotes RhoA-ROCK-dependent cell migration and invasion in 3D matrices. J. Cell Biol. 178, 23–30 (2007).

  124. 124.

    & SnapShot: The epithelial-mesenchymal transition. Cell 145, 162 (2011).

  125. 125.

    et al. A p53/miRNA-34 axis regulates Snail1-dependent cancer cell epithelial-mesenchymal transition. J. Cell Biol. 195, 417–433 (2011).

  126. 126.

    et al. p53 regulates epithelial-mesenchymal transition and stem cell properties through modulating miRNAs. Nature Cell Biol. 13, 317–323 (2011).

  127. 127.

    et al. p53 regulates epithelial-mesenchymal transition through microRNAs targeting ZEB1 and ZEB2. J. Exp. Med. 208, 875–883 (2011).

  128. 128.

    et al. Rb deletion in mouse mammary progenitors induces luminal-B or basal-like/EMT tumor subtypes depending on p53 status. J. Clin. Invest. 120, 3296–3309 (2010).

  129. 129.

    et al. Loss of p53 in enterocytes generates an inflammatory microenvironment enabling invasion and lymph node metastasis of carcinogen-induced colorectal tumors. Cancer Cell 23, 93–106 (2013).

  130. 130.

    et al. EMT and induction of miR-21 mediate metastasis development in Trp53-deficient tumours. Sci. Rep. 2, 434 (2012).

  131. 131.

    & Accessories to the crime: functions of cells recruited to the tumor microenvironment. Cancer Cell 21, 309–322 (2012).

  132. 132.

    , , & Control of angiogenesis in fibroblasts by p53 regulation of thrombospondin-1. Science 265, 1582–1584 (1994).

  133. 133.

    , & Interactions between the tumor suppressor p53 and immune responses. Curr. Opin. Oncol. 25, 85–92 (2013).

  134. 134.

    et al. Non-cell-autonomous tumor suppression by p53. Cell 153, 449–460 (2013).

  135. 135.

    et al. Senescence-associated secretory phenotypes reveal cell-nonautonomous functions of oncogenic RAS and the p53 tumor suppressor. PLoS Biol. 6, 2853–2868 (2008).

  136. 136.

    , , & Selective evolution of stromal mesenchyme with p53 loss in response to epithelial tumorigenesis. Cell 123, 1001–1011 (2005).

  137. 137.

    et al. Frequent somatic mutations in PTEN and TP53 are mutually exclusive in the stroma of breast carcinomas. Nature Genet. 32, 355–357 (2002).

  138. 138.

    , , , & Possible alternative carcinogenesis pathway featuring microsatellite instability in colorectal cancer stroma. Br. J. Cancer 89, 707–712 (2003).

  139. 139.

    et al. Molecular genetic alterations in the laser-capture-microdissected stroma adjacent to bladder carcinoma. Cancer 98, 1830–1836 (2003).

  140. 140.

    et al. Genetic alterations in the peritumoral stromal cells of malignant and borderline epithelial ovarian tumors as indicated by allelic imbalance on chromosome 3p. Int. J. Cancer 109, 247–252 (2004).

  141. 141.

    et al. Evidence for nonautonomous effect of p53 tumor suppressor in carcinogenesis. Cancer Res. 65, 1627–1630 (2005).

  142. 142.

    & Live or let die: the cell's response to p53. Nature Rev. Cancer 2, 594–604 (2002).

  143. 143.

    , , & Crystal structure of a p53 tumor suppressor-DNA complex: understanding tumorigenic mutations. Science 265, 346–355 (1994).

  144. 144.

    et al. Gain of function mutations in p53. Nature Genet. 4, 42–46 (1993).

  145. 145.

    & When mutants gain new powers: news from the mutant p53 field. Nature Rev. Cancer 9, 701–713 (2009).

  146. 146.

    , & A common gain of function of p53 cancer mutants in inducing genetic instability. Oncogene 29, 949–956 (2010).

  147. 147.

    et al. Two hot spot mutant p53 mouse models display differential gain of function in tumorigenesis. Cell Death Differ. 20, 898–909 (2013).

  148. 148.

    et al. Mutant p53 gain of function in two mouse models of Li-Fraumeni syndrome. Cell 119, 847–860 (2004).

  149. 149.

    et al. Mutant p53(R270H) gain of function phenotype in a mouse model for oncogene-induced mammary carcinogenesis. Int. J. Cancer 122, 1701–1709 (2008).

  150. 150.

    & p53 mutations in cancer. Nature Cell Biol. 15, 2–8 (2013).

  151. 151.

    et al. Gain of function of mutant p53: the mutant p53/NF-Y protein complex reveals an aberrant transcriptional mechanism of cell cycle regulation. Cancer Cell 10, 191–202 (2006).

  152. 152.

    , & TopBP1 mediates mutant p53 gain of function through NF-Y and p63/p73. Mol. Cell. Biol. 31, 4464–4481 (2011).

  153. 153.

    et al. Modulation of the vitamin D3 response by cancer-associated mutant p53. Cancer Cell 17, 273–285 (2010).

  154. 154.

    et al. Physical interaction with human tumor-derived p53 mutants inhibits p63 activities. J. Biol. Chem. 277, 18817–18826 (2002).

  155. 155.

    , , , & A subset of tumor-derived mutant forms of p53 down-regulate p63 and p73 through a direct interaction with the p53 core domain. Mol. Cell. Biol. 21, 1874–1887 (2001).

  156. 156.

    et al. Gain-of-function p53 mutants have widespread genomic locations partially overlapping with p63. Oncotarget 3, 132–143 (2012).

  157. 157.

    & Cell-cycle checkpoints and cancer. Nature 432, 316–323 (2004).

  158. 158.

    & Posttranslational modification of p53: cooperative integrators of function. Cold Spring Harb. Perspect Biol. 1, a000950 (2009).

  159. 159.

    , , & DNA damage-induced phosphorylation of p53 alleviates inhibition by MDM2. Cell 91, 325–334 (1997).

  160. 160.

    , , & Phosphorylation of Ser-20 mediates stabilization of human p53 in response to DNA damage. Proc. Natl Acad. Sci. USA 96, 13777–13782 (1999).

  161. 161.

    , & The p53 orchestra: Mdm2 and Mdmx set the tone. Trends Cell Biol. 20, 299–309 (2010).

  162. 162.

    , & Rescue of early embryonic lethality in mdm2-deficient mice by deletion of p53. Nature 378, 203–206 (1995).

  163. 163.

    et al. Rescue of embryonic lethality in Mdm4-null mice by loss of Trp53 suggests a nonoverlapping pathway with MDM2 to regulate p53. Nature Genet. 29, 92–95 (2001).

  164. 164.

    et al. Oncoprotein MDM2 conceals the activation domain of tumour suppressor p53. Nature 362, 857–860 (1993).

  165. 165.

    , , , & The mdm-2 oncogene product forms a complex with the p53 protein and inhibits p53-mediated transactivation. Cell 69, 1237–1245 (1992).

  166. 166.

    et al. Amplification of Mdmx (or Mdm4) directly contributes to tumor formation by inhibiting p53 tumor suppressor activity. Mol. Cell. Biol. 24, 5835–5843 (2004).

  167. 167.

    , & Regulation of p53 stability by Mdm2. Nature 387, 299–303 (1997).

  168. 168.

    , , & Mdm2 promotes the rapid degradation of p53. Nature 387, 296–299 (1997).

  169. 169.

    , , & p53 N-terminal phosphorylation: a defining layer of complex regulation. Carcinogenesis 33, 1441–1449 (2012).

  170. 170.

    et al. p14ARF links the tumour suppressors RB and p53. Nature 395, 124–125 (1998).

  171. 171.

    , , , & Distinct roles for E2F proteins in cell growth control and apoptosis. Proc. Natl. Acad. Sci. USA 94, 7245–7250 (1997).

  172. 172.

    & Association of p19ARF with Mdm2 inhibits ubiquitin ligase activity of Mdm2 for tumor suppressor p53. EMBO J. 18, 22–27 (1999).

  173. 173.

    & Control of p53 ubiquitination and nuclear export by MDM2 and ARF. Cell Growth Differ. 12, 175–186 (2001).

  174. 174.

    & Mutations in human ARF exon 2 disrupt its nucleolar localization and impair its ability to block nuclear export of MDM2 and p53. Mol. Cell 3, 579–591 (1999).

  175. 175.

    , , , & Nucleolar Arf sequesters Mdm2 and activates p53. Nature Cell Biol. 1, 20–26 (1999).

  176. 176.

    & Tumor suppression by Ink4a-Arf: progress and puzzles. Curr. Opin. Genet. Dev. 13, 77–83 (2003).

  177. 177.

    et al. Regulation of MCP-1 chemokine transcription by p53. Mol. Cancer 9, 82 (2010).

  178. 178.

    et al. p53 activates ICAM-1 (CD54) expression in an NF-κB-independent manner. EMBO J. 22, 1567–1578 (2003).

  179. 179.

    , , & Identification of NCF2/p67phox as a novel p53 target gene. Cell Cycle 11, 4589–4596 (2012).

  180. 180.

    et al. Notch1 is a p53 target gene involved in human keratinocyte tumor suppression through negative regulation of ROCK1/2 and MRCKα kinases. Genes Dev. 21, 562–577 (2007).

  181. 181.

    & The regulation of energy metabolism and the IGF-1/mTOR pathways by the p53 protein. Trends Cell Biol. 20, 427–434 (2010).

  182. 182.

    , , , & Metabolic regulation by p53 family members. Cell. Metab. 18, 617–633 (2013).

  183. 183.

    , & Ku86 autoantigen related protein-1 transcription initiates from a CpG island and is induced by p53 through a nearby p53 response element. Nucleic Acids Res. 30, 1713–1724 (2002).

  184. 184.

    , & The regulation of cellular metabolism by tumor suppressor p53. Cell Biosci. 3, 9 (2013).

  185. 185.

    et al. Identification of fractalkine, a CX3C-type chemokine, as a direct target of p53. Cancer Res. 60, 3722–3726 (2000).

  186. 186.

    et al. p53-dependent senescence delays Eμ-myc-induced B-cell lymphomagenesis. Oncogene 29, 1260–1269 (2010).

Download references

Acknowledgements

The authors thank P. Garcia, N. Raj, and D. Jiang for critical reading of the manuscript. The authors apologize to those whose work was not cited owing to space constraints.

Author information

Affiliations

  1. Division of Radiation and Cancer Biology, Department of Radiation Oncology, Stanford University School of Medicine, CCSR-South, Room 1255, 269 Campus Drive, Stanford, California 94305, USA.

    • Kathryn T. Bieging
    • , Stephano Spano Mello
    •  & Laura D. Attardi
  2. Department of Genetics, Stanford University School of Medicine, CCSR-South, Room 1255, 269 Campus Drive, Stanford, California 94305, USA.

    • Laura D. Attardi

Authors

  1. Search for Kathryn T. Bieging in:

  2. Search for Stephano Spano Mello in:

  3. Search for Laura D. Attardi in:

Competing interests

The authors declare no competing financial interests.

Corresponding author

Correspondence to Laura D. Attardi.

About this article

Publication history

Published

DOI

https://doi.org/10.1038/nrc3711

Further reading