Skip to main content

Thank you for visiting nature.com. You are using a browser version with limited support for CSS. To obtain the best experience, we recommend you use a more up to date browser (or turn off compatibility mode in Internet Explorer). In the meantime, to ensure continued support, we are displaying the site without styles and JavaScript.

  • Review Article
  • Published:

Regulation of the INK4b–ARF–INK4a tumour suppressor locus: all for one or one for all

Key Points

  • The INK4b–ARF–INK4a locus encodes three proteins that are implicated in senescence and tumour suppression by virtue of their effect on the retinoblastoma protein (pRb) and p53 pathways.

  • Senescence is an important mechanism of tumour suppression that protects against unrestrained cell proliferation. There is growing evidence that senescence occurs in vivo in pre-malignant lesions.

  • The unusual evolution of the INK4b–ARF–INK4a locus would be consistent with a need for co-regulation in response to hyperproliferative signals.

  • An extensive list of factors is reported to activate expression of one or more of the genes encoded by the locus, in different contexts. Activators are likely to engage senescence, and therefore the oncogenic potential of these agents is only revealed in cells in which the defences have been breached.

  • Repressors of the INK4b–ARF–INK4a locus are likely to register as oncoproteins because they undermine the cell's capacity to become senescent.

  • The locus is regulated by members of the Polycomb group of transcriptional repressors that are implicated in cell-fate decisions and the maintenance of stem-cell identity.

Abstract

The INK4b–ARF–INK4a locus encodes two members of the INK4 family of cyclin-dependent kinase inhibitors, p15INK4b and p16INK4a, and a completely unrelated protein, known as ARF. All three products participate in major tumour suppressor networks that are disabled in human cancer and influence key physiological processes such as replicative senescence, apoptosis and stem-cell self-renewal. Transcription from the locus is therefore kept under strict control. Mounting evidence suggests that although the individual genes can respond independently to positive and negative signals in different contexts, the entire locus might be coordinately suppressed by a cis-acting regulatory domain or by the action of Polycomb group repressor complexes.

This is a preview of subscription content, access via your institution

Access options

Buy this article

Prices may be subject to local taxes which are calculated during checkout

Figure 1: Organization of the INK4b–ARF–INK4a locus in different species.
Figure 2: Activators and repressors of the INK4b–ARF–INK4a locus.

Similar content being viewed by others

References

  1. Gonzalez, S. et al. Oncogenic activity of Cdc6 through repression of the INK4/ARF locus. Nature 440, 702–706 (2006). The authors identify an origin of DNA replication adjacent to the INK4b gene and show that it also functions as a regulatory domain for transcription. Loading of CDC6 or siRNA directed against the regulatory domain suppresses transcription from the entire INK4b–ARF–INK4a locus.

    CAS  PubMed  Google Scholar 

  2. Valk-Lingbeek, M. E., Bruggeman, S. W. & van Lohuizen, M. Stem cells and cancer; the polycomb connection. Cell 118, 409–418 (2004).

    CAS  PubMed  Google Scholar 

  3. Ruas, M. & Peters, G. The p16INK4a/CDKN2A tumor suppressor and its relatives. Biochim. Biophys. Acta 1378, F115–F177 (1998).

    CAS  PubMed  Google Scholar 

  4. Sharpless, N. E. INK4a/ARF: a multifunctional tumor suppressor locus. Mutat. Res. 576, 22–38 (2005).

    CAS  PubMed  Google Scholar 

  5. Kamb, A. et al. A cell cycle regulator potentially involved in genesis of many tumor types. Science 264, 436–440 (1994).

    CAS  PubMed  Google Scholar 

  6. Nobori, T. et al. Deletions of the cyclin-dependent kinase-4 inhibitor gene in multiple human cancers. Nature 368, 753–756 (1994).

    CAS  PubMed  Google Scholar 

  7. Hannon, G. J. & Beach, D. p15INK4B is a potential effector of TGF-β-induced cell cycle arrest. Nature 371, 257–261 (1994).

    CAS  PubMed  Google Scholar 

  8. Jen, J. et al. Deletion of p16 and p15 genes in brain tumors. Cancer Res. 54, 6353–6358 (1994).

    CAS  PubMed  Google Scholar 

  9. Mao, L. et al. A novel p16INK4A transcript. Cancer Res. 55, 2995–2997 (1995).

    CAS  PubMed  Google Scholar 

  10. Quelle, D. E., Zindy, F., Ashmun, R. A. & Sherr, C. J. Alternative reading frames of the INK4a tumor suppressor gene encode two unrelated proteins capable of inducing cell cycle arrest. Cell 83, 993–1000 (1995).

    CAS  PubMed  Google Scholar 

  11. Stone, S. et al. Complex structure and regulation of the P16 (MTS1) locus. Cancer Res. 55, 2988–2994 (1995).

    CAS  PubMed  Google Scholar 

  12. Kazianis, S. et al. Comparative structure and characterization of a CDKN2 gene in a Xiphophorus fish melanoma model. Oncogene 18, 5088–5099 (1999).

    CAS  PubMed  Google Scholar 

  13. Gilley, J. & Fried, M. One INK4 gene and no ARF at the Fugu equivalent of the human INK4A/ARF/INK4B tumour suppressor locus. Oncogene 20, 7447–7452 (2001).

    CAS  PubMed  Google Scholar 

  14. Kim, S. H., Mitchell, M., Fujii, H., Llanos, S. & Peters, G. Absence of p16INK4a and truncation of ARF tumor suppressors in chickens. Proc. Natl Acad. Sci. USA 100, 211–216 (2003).

    CAS  PubMed  Google Scholar 

  15. Quelle, D. E., Cheng, M., Ashmun, R. A. & Sherr, C. J. Cancer-associated mutations at the INK4a locus cancel cell cycle arrest by p16INK4a but not by the alternative reading frame protein p19ARF. Proc. Natl Acad. Sci. USA 94, 669–673 (1997).

    CAS  PubMed  Google Scholar 

  16. Stott, F. J. et al. The alternative product from the human CDKN2A locus, p14ARF, participates in a regulatory feedback loop with p53 and MDM2. EMBO J. 17, 5001–5014 (1998).

    CAS  PubMed  PubMed Central  Google Scholar 

  17. Bothner, B. et al. Defining the molecular basis of Arf and Hdm2 interactions. J. Mol. Biol. 314, 263–277 (2001).

    CAS  PubMed  Google Scholar 

  18. Pavletich, N. P. Mechanisms of cyclin-dependent kinase regulation: structures of Cdks, their cyclin activators, and Cip and INK4 inhibitors. J. Mol. Biol. 287, 821–828 (1999).

    CAS  PubMed  Google Scholar 

  19. McKeller, R. N. et al. The Arf tumor suppressor gene promotes hyaloid vascular regression during mouse eye development. Proc. Natl Acad. Sci. USA 99, 3848–3853 (2002).

    CAS  PubMed  Google Scholar 

  20. Krishnamurthy, J. et al. Ink4a/Arf expression is a biomarker of aging. J. Clin. Invest. 114, 1299–1307 (2004).

    CAS  PubMed  PubMed Central  Google Scholar 

  21. Zindy, F., Quelle, D. E., Roussel, M. F. & Sherr, C. J. Expression of the p16INK4a tumor suppressor versus other INK4 family members during mouse development and aging. Oncogene 15, 203–211 (1997).

    CAS  PubMed  Google Scholar 

  22. Zindy, F. et al. Arf tumor suppressor promoter monitors latent oncogenic signals in vivo. Proc. Natl Acad. Sci. USA 100, 15930–15935 (2003).

    CAS  PubMed  Google Scholar 

  23. Kim, S. H. et al. Upregulation of chicken p15INK4b at senescence and in the developing brain. J. Cell Sci. 119, 2435–2443 (2006).

    CAS  PubMed  Google Scholar 

  24. Serrano, M., Lin, A. W., McCurrach, M. E., Beach, D. & Lowe, S. W. Oncogenic ras provokes premature cell senescence associated with accumulation of p53 and p16INK4a. Cell 88, 593–602 (1997). A seminal paper that describes oncogene-induced senescence in response to activated Ras and the role of INK4a in mediating the arrest.

    CAS  PubMed  Google Scholar 

  25. Wright, W. E. & Shay, J. W. Historical claims and current interpretations of replicative aging. Nature Biotechnol. 20, 682–688 (2002).

    CAS  Google Scholar 

  26. Rangarajan, A. & Weinberg, R. A. Comparative biology of mouse versus human cells: modelling human cancer in mice. Nature Rev. Cancer 3, 952–959 (2003).

    CAS  Google Scholar 

  27. Kamijo, T. et al. Tumor suppression at the mouse INK4a locus mediated by the alternative reading frame product p19ARF. Cell 91, 649–659 (1997).

    CAS  PubMed  Google Scholar 

  28. Latres, E. et al. Limited overlapping roles of p15INK4b and p18INK4c cell cycle inhibitors in proliferation and tumorigenesis. EMBO J. 19, 3496–3506 (2000).

    CAS  PubMed  PubMed Central  Google Scholar 

  29. Krimpenfort, P., Quon, K. C., Mooi, W. J., Loonstra, A. & Berns, A. Loss of p16Ink4a confers susceptibility to metastatic melanoma in mice. Nature 413, 83–86 (2001).

    CAS  PubMed  Google Scholar 

  30. Sharpless, N. E. et al. Loss of p16Ink4a with retention of p19Arf predisposes mice to tumorigenesis. Nature 413, 86–91 (2001).

    CAS  PubMed  Google Scholar 

  31. Sharpless, N. E., Ramsey, M. R., Balasubramanian, P., Castrillon, D. H. & DePinho, R. A. The differential impact of p16INK4a or p19ARF deficiency on cell growth and tumorigenesis. Oncogene 23, 379–385 (2004). References 27–31 report on the tumour susceptibility of mice that are specifically defective for Ink4b, Arf or Ink4a , as well as characterizing the behaviour of the respective fibroblasts.

    CAS  PubMed  Google Scholar 

  32. Randle, D. H., Zindy, F., Sherr, C. J. & Roussel, M. F. Differential effects of p19Arf and p16Ink4a loss on senescence of murine bone marrow-derived preB cells and macrophages. Proc. Natl Acad. Sci. USA 98, 9654–9659 (2001).

    CAS  PubMed  Google Scholar 

  33. Lin, A. W. et al. Premature senescence involving p53 and p16 is activated in response to constitutive MEK/MAPK mitogenic signaling. Genes Dev. 12, 3008–3019 (1998).

    CAS  PubMed  PubMed Central  Google Scholar 

  34. Zhu, J., Woods, D., McMahon, M. & Bishop, J. M. Senescence of human fibroblasts induced by oncogenic Raf. Genes Dev. 12, 2997–3007 (1998).

    CAS  PubMed  PubMed Central  Google Scholar 

  35. Ohtani, N. et al. Opposing effects of Ets and Id proteins on p16INK4a expression during cellular senescence. Nature 409, 1067–1070 (2001).

    CAS  PubMed  Google Scholar 

  36. Michaloglou, C. et al. BRAFE600-associated senescence-like cell cycle arrest of human naevi. Nature 436, 720–724 (2005). Along with reference 43, this represents one of the first descriptions of oncogene-induced senescence in vivo in pre-malignant tumours. These studies lay to rest concerns that senescence might be a tissue-culture artefact.

    CAS  PubMed  Google Scholar 

  37. Iwasa, H., Han, J. & Ishikawa, F. Mitogen-activated protein kinase p38 defines the common senescence-signalling pathway. Genes Cells 8, 131–144 (2003).

    CAS  PubMed  Google Scholar 

  38. Haq, R. et al. Constitutive p38HOG mitogen-activated protein kinase activation induces permanent cell cycle arrest and senescence. Cancer Res. 62, 5076–5082 (2002).

    CAS  PubMed  Google Scholar 

  39. Deng, Q., Liao, R., Wu, B. L. & Sun, P. High intensity ras signaling induces premature senescence by activating p38 pathway in primary human fibroblasts. J. Biol. Chem. 279, 1050–1059 (2004).

    CAS  PubMed  Google Scholar 

  40. Wang, W. et al. Sequential activation of the MEK-extracellular signal-regulated kinase and MKK3/6-p38 mitogen-activated protein kinase pathways mediates oncogenic ras-induced premature senescence. Mol. Cell. Biol. 22, 3389–3403 (2002).

    PubMed  PubMed Central  Google Scholar 

  41. Bulavin, D. V. et al. Inactivation of the Wip1 phosphatase inhibits mammary tumorigenesis through p38 MAPK-mediated activation of the p16Ink4a–p19Arf pathway. Nature Genet. 36, 343–350 (2004).

    CAS  PubMed  Google Scholar 

  42. Ito, K. et al. Reactive oxygen species act through p38 MAPK to limit the lifespan of hematopoietic stem cells. Nature Med. 12, 446–451 (2006).

    CAS  PubMed  Google Scholar 

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

    CAS  PubMed  Google Scholar 

  44. Malumbres, M. et al. Cellular response to oncogenic Ras involves induction of the Cdk4 and Cdk6 inhibitor p15INK4b. Mol. Cell. Biol. 20, 2915–2925 (2000).

    CAS  PubMed  PubMed Central  Google Scholar 

  45. Palmero, I., Pantoja, C. & Serrano, M. p19ARF links the tumour suppressor p53 to Ras. Nature 395, 125–126 (1998). Together with reference 64, this paper provides the first evidence that Arf is activated by oncogenic signalling.

    CAS  PubMed  Google Scholar 

  46. Groth, A., Weber, J. D., Willumsen, B. M., Sherr, C. J. & Roussel, M. F. Oncogenic Ras induces p19ARF and growth arrest in mouse embryo fibroblasts lacking p21Cip1 and p27Kip1 without activating cyclin D-dependent kinases. J. Biol. Chem. 275, 27473–27480 (2000).

    CAS  PubMed  Google Scholar 

  47. Lin, A. W. & Lowe, S. W. Oncogenic ras activates the ARF–p53 pathway to suppress epithelial cell transformation. Proc. Natl Acad. Sci. USA 98, 5025–5030 (2001).

    CAS  PubMed  Google Scholar 

  48. Ferbeyre, G. et al. PML is induced by oncogenic ras and promotes premature senescence. Genes Dev. 14, 2015–2027 (2000).

    CAS  PubMed  PubMed Central  Google Scholar 

  49. Wei, W., Hemmer, R. M. & Sedivy, J. M. Role of p14ARF in replicative and induced senescence of human fibroblasts. Mol. Cell. Biol. 21, 6748–6757 (2001).

    CAS  PubMed  PubMed Central  Google Scholar 

  50. Brookes, S. et al. INK4a-deficient human diploid fibroblasts are resistant to RAS-induced senescence. EMBO J. 21, 2936–2945 (2002).

    CAS  PubMed  PubMed Central  Google Scholar 

  51. Berkovich, E., Lamed, Y. & Ginsberg, D. E2F and Ras synergize in transcriptionally activating p14ARF expression. Cell Cycle 2, 127–133 (2003).

    CAS  PubMed  Google Scholar 

  52. Rangarajan, A., Hong, S. J., Gifford, A. & Weinberg, R. A. Species- and cell type-specific requirements for cellular transformation. Cancer Cell 6, 171–183 (2004).

    CAS  PubMed  Google Scholar 

  53. Inoue, K., Roussel, M. F. & Sherr, C. J. Induction of ARF tumor suppressor gene expression and cell cycle arrest by transcription factor DMP1. Proc. Natl Acad. Sci. USA 96, 3993–3998 (1999).

    CAS  PubMed  Google Scholar 

  54. Sreeramaneni, R., Chaudhry, A., McMahon, M., Sherr, C. J. & Inoue, K. Ras–Raf–Arf signaling critically depends on the Dmp1 transcription factor. Mol. Cell. Biol. 25, 220–232 (2005).

    CAS  PubMed  PubMed Central  Google Scholar 

  55. Evan, G. I. et al. Induction of apoptosis in fibroblasts by c-myc protein. Cell 69, 119–128 (1992).

    CAS  PubMed  Google Scholar 

  56. Alevizopoulos, K., Vlach, J., Hennecke, S. & Amati, B. Cyclin E and c-Myc promote cell proliferation in the presence of p16INK4a and of hypophosphorylated retinoblastoma family proteins. EMBO J. 16, 5322–5333 (1997).

    CAS  PubMed  PubMed Central  Google Scholar 

  57. Felsher, D. W., Zetterberg, A., Zhu, J., Tlsty, T. & Bishop, J. M. Overexpression of MYC causes p53-dependent G2 arrest of normal fibroblasts. Proc. Natl Acad. Sci. USA 97, 10544–10548 (2000).

    CAS  PubMed  Google Scholar 

  58. Hermeking, H. et al. Identification of CDK4 as a target of c-MYC. Proc. Natl Acad. Sci. USA 97, 2229–2234 (2000).

    CAS  PubMed  Google Scholar 

  59. Mateyak, M. K., Obaya, A. J. & Sedivy, J. M. c-Myc regulates cyclin D–Cdk4 and –Cdk6 activity but affects cell cycle progression at multiple independent points. Mol. Cell. Biol. 19, 4672–4683 (1999).

    CAS  PubMed  PubMed Central  Google Scholar 

  60. Vafa, O. et al. c-Myc can induce DNA damage, increase reactive oxygen species, and mitigate p53 function: a mechanism for oncogene-induced genetic instability. Mol. Cell 9, 1031–1044 (2002).

    CAS  PubMed  Google Scholar 

  61. Drayton, S. et al. Tumor suppressor p16INK4a determines sensitivity of human cells to transformation by cooperating cellular oncogenes. Cancer Cell 4, 301–310 (2003).

    CAS  PubMed  Google Scholar 

  62. Guney, I., Wu, S. & Sedivy, J. M. Reduced c-Myc signaling triggers telomere-independent senescence by regulating Bmi-1 and p16INK4a. Proc. Natl Acad. Sci. USA 103, 3645–3650 (2006).

    CAS  PubMed  Google Scholar 

  63. Grandori, C. et al. Werner syndrome protein limits MYC-induced cellular senescence. Genes Dev. 17, 1569–1574 (2003).

    CAS  PubMed  PubMed Central  Google Scholar 

  64. Zindy, F. et al. Myc signaling via the ARF tumor suppressor regulates p53-dependent apoptosis and immortalization. Genes Dev. 12, 2424–2433 (1998).

    CAS  PubMed  PubMed Central  Google Scholar 

  65. Haupt, Y., Alexander, W. S., Barri, G., Klinken, S. P. & Adams, J. M. Novel zinc finger gene implicated as myc collaborator by retrovirally accelerated lymphomagenesis in E mu–myc transgenic mice. Cell 65, 753–763 (1991).

    CAS  PubMed  Google Scholar 

  66. van Lohuizen, M. et al. Identification of cooperating oncogenes in E mu–myc transgenic mice by provirus tagging. Cell 65, 737–752 (1991).

    CAS  PubMed  Google Scholar 

  67. Eischen, C. M., Weber, J. D., Roussel, M. F., Sherr, C. J. & Cleveland, J. L. Disruption of the ARF–Mdm2–p53 tumor suppressor pathway in Myc-induced lymphomagenesis. Genes Dev. 13, 2658–2669 (1999).

    CAS  PubMed  PubMed Central  Google Scholar 

  68. Jacobs, J. J. et al. Bmi-1 collaborates with c-Myc in tumorigenesis by inhibiting c-Myc-induced apoptosis via INK4a/ARF. Genes Dev. 13, 2678–2690 (1999).

    CAS  PubMed  PubMed Central  Google Scholar 

  69. Qi, Y. et al. p19ARF directly and differentially controls the functions of c-Myc independently of p53. Nature 431, 712–717 (2004).

    CAS  PubMed  Google Scholar 

  70. Datta, A. et al. Myc–ARF (alternate reading frame) interaction inhibits the functions of Myc. J. Biol. Chem. 279, 36698–36707 (2004).

    CAS  PubMed  Google Scholar 

  71. Seoane, J. et al. TGFβ influences Myc, Miz-1 and Smad to control the CDK inhibitor p15INK4b. Nature Cell Biol. 3, 400–408 (2001).

    CAS  PubMed  Google Scholar 

  72. Staller, P. et al. Repression of p15INK4b expression by Myc through association with Miz-1. Nature Cell Biol. 3, 392–399 (2001).

    CAS  PubMed  Google Scholar 

  73. Warner, B. J., Blain, S. W., Seoane, J. & Massague, J. Myc downregulation by transforming growth factor beta required for activation of the p15Ink4b G1 arrest pathway. Mol. Cell. Biol. 19, 5913–5922 (1999).

    CAS  PubMed  PubMed Central  Google Scholar 

  74. Hara, E. et al. Regulation of p16CDKN2 expression and its implications for cell immortalization and senescence. Mol. Cell. Biol. 16, 859–867 (1996).

    CAS  PubMed  PubMed Central  Google Scholar 

  75. Palmero, I. et al. Accumulation of p16INK4a in mouse fibroblasts as a function of replicative senescence and not of retinoblastoma gene status. Oncogene 15, 495–503 (1997).

    CAS  PubMed  Google Scholar 

  76. de Stanchina, E. et al. E1A signaling to p53 involves the p19ARF tumor suppressor. Genes Dev. 12, 2434–2442 (1998).

    CAS  PubMed  PubMed Central  Google Scholar 

  77. Lomazzi, M., Moroni, M. C., Jensen, M. R., Frittoli, E. & Helin, K. Suppression of the p53- or pRB-mediated G1 checkpoint is required for E2F-induced S-phase entry. Nature Genet. 31, 190–194 (2002).

    CAS  PubMed  Google Scholar 

  78. Kel, A. E. et al. Computer-assisted identification of cell cycle-related genes: new targets for E2F transcription factors. J. Mol. Biol. 309, 99–120 (2001).

    CAS  PubMed  Google Scholar 

  79. Aslanian, A., Iaquinta, P. J., Verona, R. & Lees, J. A. Repression of the Arf tumor suppressor by E2F3 is required for normal cell cycle kinetics. Genes Dev. 18, 1413–1422 (2004).

    CAS  PubMed  PubMed Central  Google Scholar 

  80. Komori, H., Enomoto, M., Nakamura, M., Iwanaga, R. & Ohtani, K. Distinct E2F-mediated transcriptional program regulates p14ARF gene expression. EMBO J. 24, 3724–3736 (2005).

    CAS  PubMed  PubMed Central  Google Scholar 

  81. Lowe, S. W. & Sherr, C. J. Tumor suppression by INK4a–ARF: progress and puzzles. Curr. Opin. Genet. Dev. 13, 77–83 (2003).

    CAS  PubMed  Google Scholar 

  82. Passegue, E. & Wagner, E. F. JunB suppresses cell proliferation by transcriptional activation of p16INK4a expression. EMBO J. 19, 2969–2979 (2000).

    CAS  PubMed  PubMed Central  Google Scholar 

  83. Ameyar-Zazoua, M. et al. AP-1 dimers regulate transcription of the p14/p19 ARF tumor suppressor gene. Oncogene 24, 2298–2306 (2005).

    CAS  PubMed  Google Scholar 

  84. Weitzman, J. B., Fiette, L., Matsuo, K. & Yaniv, M. JunD protects cells from p53-dependent senescence and apoptosis. Mol. Cell 6, 1109–1119 (2000).

    CAS  PubMed  Google Scholar 

  85. Roberts, C. W., Leroux, M. M., Fleming, M. D. & Orkin, S. H. Highly penetrant, rapid tumorigenesis through conditional inversion of the tumor suppressor gene Snf5. Cancer Cell 2, 415–425 (2002).

    CAS  PubMed  Google Scholar 

  86. Betz, B. L., Strobeck, M. W., Reisman, D. N., Knudsen, E. S. & Weissman, B. E. Re-expression of hSNF5/INI1/BAF47 in pediatric tumor cells leads to G1 arrest associated with induction of p16ink4a and activation of RB. Oncogene 21, 5193–5203 (2002).

    CAS  PubMed  Google Scholar 

  87. Oruetxebarria, I. et al. p16INK4a is required for hSNF5 chromatin remodeler-induced cellular senescence in malignant rhabdoid tumor cells. J. Biol. Chem. 279, 3807–3816 (2004).

    CAS  PubMed  Google Scholar 

  88. Smogorzewska, A. & de Lange, T. Different telomere damage signaling pathways in human and mouse cells. EMBO J. 21, 4338–4348 (2002).

    CAS  PubMed  PubMed Central  Google Scholar 

  89. Herbig, U., Jobling, W. A., Chen, B. P., Chen, D. J. & Sedivy, J. M. Telomere shortening triggers senescence of human cells through a pathway involving ATM, p53, and p21CIP1, but not p16INK4a. Mol. Cell 14, 501–513 (2004).

    CAS  PubMed  Google Scholar 

  90. Jacobs, J. J. & de Lange, T. Significant role for p16INK4a in p53-independent telomere-directed senescence. Curr. Biol. 14, 2302–2308 (2004).

    CAS  PubMed  Google Scholar 

  91. Morris, M., Hepburn, P. & Wynford-Thomas, D. Sequential extension of proliferative lifespan in human fibroblasts induced by over-expression of CDK4 or 6 and loss of p53 function. Oncogene 21, 4277–4288 (2002).

    CAS  PubMed  Google Scholar 

  92. Loercher, A. E., Tank, E. M., Delston, R. B. & Harbour, J. W. MITF links differentiation with cell cycle arrest in melanocytes by transcriptional activation of INK4A. J. Cell Biol. 168, 35–40 (2005).

    CAS  PubMed  PubMed Central  Google Scholar 

  93. Carreira, S. et al. Mitf cooperates with Rb1 and activates p21 Cip1 expression to regulate cell cycle progression. Nature 433, 764–769 (2005).

    CAS  PubMed  Google Scholar 

  94. Cong, F., Zou, X., Hinrichs, K., Calame, K. & Goff, S. P. Inhibition of v-Abl transformation by p53 and p19ARF. Oncogene 18, 7731–7739 (1999).

    CAS  PubMed  Google Scholar 

  95. Raveh, T., Droguett, G., Horwitz, M. S., DePinho, R. A. & Kimchi, A. DAP kinase activates a p19ARF/p53-mediated apoptotic checkpoint to suppress oncogenic transformation. Nature Cell Biol. 3, 1–7 (2001).

    CAS  PubMed  Google Scholar 

  96. Damalas, A., Kahan, S., Shtutman, M., Ben-Ze'ev, A. & Oren, M. Deregulated β-catenin induces a p53- and ARF-dependent growth arrest and cooperates with Ras in transformation. EMBO J. 20, 4912–4922 (2001).

    CAS  PubMed  PubMed Central  Google Scholar 

  97. Hansson, A., Manetopoulos, C., Jonsson, J. I. & Axelson, H. The basic helix-loop-helix transcription factor TAL1/SCL inhibits the expression of the p16INK4A and pT α genes. Biochem. Biophys. Res. Commun. 312, 1073–1081 (2003).

    CAS  PubMed  Google Scholar 

  98. Yang, X., He, Z., Xin, B. & Cao, L. LMP1 of Epstein–Barr virus suppresses cellular senescence associated with the inhibition of p16INK4a expression. Oncogene 19, 2002–2013 (2000).

    CAS  PubMed  Google Scholar 

  99. Ohtani, N. et al. Epstein–Barr virus LMP1 blocks p16INK4a-RB pathway by promoting nuclear export of E2F4/5. J. Cell Biol. 162, 173–183 (2003).

    CAS  PubMed  PubMed Central  Google Scholar 

  100. Alani, R. M., Young, A. Z. & Shifflett, C. B. Id1 regulation of cellular senescence through transcriptional repression of p16/Ink4a. Proc. Natl Acad. Sci. USA 98, 7812–7816 (2001).

    CAS  PubMed  Google Scholar 

  101. Kamijo, T. et al. Loss of the ARF tumor suppressor reverses premature replicative arrest but not radiation hypersensitivity arising from disabled atm function. Cancer Res. 59, 2464–2469 (1999).

    CAS  PubMed  Google Scholar 

  102. Krones-Herzig, A., Adamson, E. & Mercola, D. Early growth response 1 protein, an upstream gatekeeper of the p53 tumor suppressor, controls replicative senescence. Proc. Natl Acad. Sci. USA 100, 3233–3238 (2003).

    CAS  PubMed  Google Scholar 

  103. Linggi, B. et al. The t(8;21) fusion protein, AML1 ETO, specifically represses the transcription of the p14ARF tumor suppressor in acute myeloid leukemia. Nature Med. 8, 743–750 (2002).

    CAS  PubMed  Google Scholar 

  104. Maeda, T. et al. Role of the proto-oncogene Pokemon in cellular transformation and ARF repression. Nature 433, 278–285 (2005).

    CAS  PubMed  Google Scholar 

  105. Kamijo, T. et al. Functional and physical interactions of the ARF tumor suppressor with p53 and Mdm2. Proc. Natl Acad. Sci. USA 95, 8292–8297 (1998).

    CAS  PubMed  Google Scholar 

  106. Robertson, K. D. & Jones, P. A. The human ARF cell cycle regulatory gene promoter is a CpG island which can be silenced by DNA methylation and down-regulated by wild-type p53. Mol. Cell. Biol. 18, 6457–6473 (1998).

    CAS  PubMed  PubMed Central  Google Scholar 

  107. Jacobs, J. J. et al. Senescence bypass screen identifies TBX2, which represses Cdkn2a (p19 ARF) and is amplified in a subset of human breast cancers. Nature Genet. 26, 291–299 (2000).

    CAS  PubMed  Google Scholar 

  108. Brummelkamp, T. R. et al. TBX-3, the gene mutated in ulnar-mammary syndrome, is a negative regulator of p19ARF and inhibits senescence. J. Biol. Chem. 277, 6567–6572 (2002).

    CAS  PubMed  Google Scholar 

  109. Lingbeek, M. E., Jacobs, J. J. & van Lohuizen, M. The T-box repressors TBX2 and TBX3 specifically regulate the tumor suppressor gene p14ARF via a variant T-site in the initiator. J. Biol. Chem. 277, 26120–26127 (2002).

    CAS  PubMed  Google Scholar 

  110. Maestro, R. et al. Twist is a potential oncogene that inhibits apoptosis. Genes Dev. 13, 2207–2217 (1999).

    CAS  PubMed  PubMed Central  Google Scholar 

  111. Otte, A. P. & Kwaks, T. H. Gene repression by Polycomb group protein complexes: a distinct complex for every occasion? Curr. Opin. Genet. Dev. 13, 448–454 (2003).

    CAS  PubMed  Google Scholar 

  112. Lund, A. H. & van Lohuizen, M. Polycomb complexes and silencing mechanisms. Curr. Opin. Cell Biol. 16, 239–246 (2004).

    CAS  PubMed  Google Scholar 

  113. Jacobs, J. J., Kieboom, K., Marino, S., DePinho, R. A. & van Lohuizen, M. The oncogene and Polycomb-group gene bmi-1 regulates cell proliferation and senescence through the ink4a locus. Nature 397, 164–168 (1999). The first evidence that the INK4b–ARF–INK4a locus is subject to transcriptional repression by BMI1, laying the foundation for the role of Polycomb group complexes in senescence.

    CAS  PubMed  Google Scholar 

  114. van der Lugt, N. M. et al. Posterior transformation, neurological abnormalities, and severe hematopoietic defects in mice with a targeted deletion of the bmi-1 proto-oncogene. Genes Dev. 8, 757–769 (1994).

    CAS  PubMed  Google Scholar 

  115. Molofsky, A. V. et al. Bmi-1 dependence distinguishes neural stem cell self-renewal from progenitor proliferation. Nature 425, 962–967 (2003). References 115–119 describe the relative roles of Ink4a and Arf in determining the effects of Bmi1 deficiency on the self-renewal capacity of particular stem cells.

    CAS  PubMed  PubMed Central  Google Scholar 

  116. Lessard, J. & Sauvageau, G. Bmi-1 determines the proliferative capacity of normal and leukaemic stem cells. Nature 423, 255–260 (2003).

    CAS  PubMed  Google Scholar 

  117. Park, I. K. et al. Bmi-1 is required for maintenance of adult self-renewing haematopoietic stem cells. Nature 423, 302–305 (2003).

    CAS  PubMed  Google Scholar 

  118. Bruggeman, S. W. et al. Ink4a and Arf differentially affect cell proliferation and neural stem cell self-renewal in Bmi1-deficient mice. Genes Dev. 19, 1438–1443 (2005).

    CAS  PubMed  PubMed Central  Google Scholar 

  119. Molofsky, A. V., He, S., Bydon, M., Morrison, S. J. & Pardal, R. Bmi-1 promotes neural stem cell self-renewal and neural development but not mouse growth and survival by repressing the p16Ink4a and p19Arf senescence pathways. Genes Dev. 19, 1432–1437 (2005).

    CAS  PubMed  PubMed Central  Google Scholar 

  120. Core, N., Joly, F., Boned, A. & Djabali, M. Disruption of E2F signaling suppresses the INK4a-induced proliferative defect in M33-deficient mice. Oncogene 23, 7660–7668 (2004).

    CAS  PubMed  Google Scholar 

  121. Isono, K. et al. Mammalian polyhomeotic homologues Phc2 and Phc1 act in synergy to mediate polycomb repression of Hox genes. Mol. Cell. Biol. 25, 6694–6706 (2005).

    CAS  PubMed  PubMed Central  Google Scholar 

  122. Voncken, J. W. et al. Rnf2 (Ring1b) deficiency causes gastrulation arrest and cell cycle inhibition. Proc. Natl Acad. Sci. USA 100, 2468–2473 (2003).

    CAS  PubMed  Google Scholar 

  123. Gil, J., Bernard, D., Martinez, D. & Beach, D. Polycomb CBX7 has a unifying role in cellular lifespan. Nature Cell Biol. 6, 67–72 (2004).

    CAS  PubMed  Google Scholar 

  124. Bernstein, B. E. et al. A bivalent chromatin structure marks key developmental genes in embryonic stem cells. Cell 125, 315–326 (2006).

    CAS  PubMed  Google Scholar 

  125. Boyer, L. A. et al. Polycomb complexes repress developmental regulators in murine embryonic stem cells. Nature 441, 349–353 (2006).

    CAS  PubMed  Google Scholar 

  126. Bracken, A. P., Dietrich, N., Pasini, D., Hansen, K. H. & Helin, K. Genome-wide mapping of Polycomb target genes unravels their roles in cell fate transitions. Genes Dev. 20, 1123–1136 (2006).

    CAS  PubMed  PubMed Central  Google Scholar 

  127. Lee, T. I. et al. Control of developmental regulators by Polycomb in human embryonic stem cells. Cell 125, 301–313 (2006).

    CAS  PubMed  PubMed Central  Google Scholar 

  128. Negre, N. et al. Chromosomal distribution of PcG proteins during Drosophila development. PLoS Biol. 4, e170 (2006).

    PubMed  PubMed Central  Google Scholar 

  129. Tolhuis, B. et al. Genome-wide profiling of PRC1 and PRC2 Polycomb chromatin binding in Drosophila melanogaster. Nature Genet. 38, 694–699 (2006).

    CAS  PubMed  Google Scholar 

  130. Itahana, K. et al. Control of the replicative life span of human fibroblasts by p16 and the polycomb protein Bmi-1. Mol. Cell. Biol. 23, 389–401 (2003).

    CAS  PubMed  PubMed Central  Google Scholar 

  131. Leung, C. et al. Bmi1 is essential for cerebellar development and is overexpressed in human medulloblastomas. Nature 428, 337–341 (2004).

    CAS  PubMed  Google Scholar 

  132. Kranc, K. R. et al. Transcriptional coactivator Cited2 induces Bmi1 and Mel18 and controls fibroblast proliferation via Ink4a/ARF. Mol. Cell. Biol. 23, 7658–7666 (2003).

    CAS  PubMed  PubMed Central  Google Scholar 

  133. Sun, L. Q. et al. Growth retardation and premature aging phenotypes in mice with disruption of the SNF2-like gene, PASG. Genes Dev. 18, 1035–1046 (2004).

    CAS  PubMed  PubMed Central  Google Scholar 

  134. Smith, K. S. et al. Bmi-1 regulation of INK4A–ARF is a downstream requirement for transformation of hematopoietic progenitors by E2a–Pbx1. Mol. Cell 12, 393–400 (2003). In addition to altered expression of the BMI1 gene in human cancers, this paper shows how regulators of the Polycomb proteins might be implicated in tumorigenesis, through downstream effects on INK4b–ARF–INK4a.

    CAS  PubMed  Google Scholar 

  135. Yannoni, Y. M., Gaestel, M. & Lin, L. L. P66ShcA interacts with MAPKAP kinase 2 and regulates its activity. FEBS Lett. 564, 205–211 (2004).

    CAS  PubMed  Google Scholar 

  136. Voncken, J. W. et al. MAPKAP kinase 3pK phosphorylates and regulates chromatin association of the polycomb group protein Bmi1. J. Biol. Chem. 280, 5178–5187 (2005).

    CAS  PubMed  Google Scholar 

  137. Cheung, P., Allis, C. D. & Sassone-Corsi, P. Signaling to chromatin through histone modifications. Cell 103, 263–271 (2000).

    CAS  PubMed  Google Scholar 

  138. Fischle, W. et al. Regulation of HP1-chromatin binding by histone H3 methylation and phosphorylation. Nature 438, 1116–1122 (2005).

    CAS  PubMed  Google Scholar 

  139. Hirota, T., Lipp, J. J., Toh, B. H. & Peters, J. M. Histone H3 serine 10 phosphorylation by Aurora B causes HP1 dissociation from heterochromatin. Nature 438, 1176–1180 (2005).

    CAS  PubMed  Google Scholar 

  140. Fischle, W., Wang, Y. & Allis, C. D. Binary switches and modification cassettes in histone biology and beyond. Nature 425, 475–479 (2003).

    CAS  PubMed  Google Scholar 

  141. Brookes, S., Rowe, J., Gutierrez Del Arroyo, A., Bond, J. & Peters, G. Contribution of p16INK4a to replicative senescence of human fibroblasts. Exp. Cell Res. 298, 549–559 (2004).

    CAS  PubMed  Google Scholar 

  142. Hernandez-Munoz, I., Taghavi, P., Kuijl, C., Neefjes, J. & van Lohuizen, M. Association of BMI1 with polycomb bodies is dynamic and requires PRC2/EZH2 and the maintenance DNA methyltransferase DNMT1. Mol. Cell. Biol. 25, 11047–11058 (2005).

    CAS  PubMed  PubMed Central  Google Scholar 

  143. Vire, E. et al. The Polycomb group protein EZH2 directly controls DNA methylation. Nature 439, 871–874 (2006).

    CAS  PubMed  Google Scholar 

  144. Sherr, C. J. The Pezcoller lecture: cancer cell cycles revisited. Cancer Res. 60, 3689–3695 (2000).

    CAS  PubMed  Google Scholar 

  145. Ortega, S., Malumbres, M. & Barbacid, M. Cyclin D-dependent kinases, INK4 inhibitors and cancer. Biochim. Biophys. Acta 1602, 73–87 (2002).

    CAS  PubMed  Google Scholar 

  146. Giaccia, A. J. & Kastan, M. B. The complexity of p53 modulation: emerging patterns from divergent signals. Genes Dev. 12, 2973–2983 (1998).

    CAS  PubMed  Google Scholar 

  147. Michael, D. & Oren, M. The p53–Mdm2 module and the ubiquitin system. Semin. Cancer Biol. 13, 49–58 (2003).

    CAS  PubMed  Google Scholar 

  148. Evan, G. I. & Vousden, K. H. Proliferation, cell cycle and apoptosis in cancer. Nature 411, 342–348 (2001).

    CAS  PubMed  Google Scholar 

  149. Campisi, J. Senescent cells, tumor suppression, and organismal aging: good citizens, bad neighbors. Cell 120, 513–522 (2005).

    CAS  PubMed  Google Scholar 

  150. Collado, M. & Serrano, M. The power and the promise of oncogene-induced senescence markers. Nature Rev. Cancer 6, 472–476 (2006).

    CAS  Google Scholar 

Download references

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Gordon Peters.

Ethics declarations

Competing interests

The authors declare no competing financial interests.

Related links

Related links

FURTHER INFORMATION

Gordon Peters's homepage

Jesús Gil's homepage

Glossary

Senescence

A state of apparently permanent cell-cycle arrest that a cell adopts in the face of aberrant proliferative signals, intracellular stress or DNA damage.

Tumour suppressor

A gene whose loss of function contributes to cancer progression. Mutations in tumour suppressor genes are generally recessive and tumours arise when both copies of the gene are inactivated; for example, by deletion of the normal allele.

Epigenetic

Effects on the patterns of gene expression that are heritable through cell division without altering the genetic information carried in the cellular DNA.

Oncogene

A gene that contributes to cancer progression as a consequence of overexpression or of dominantly acting mutations that alter the activity or specificity of the gene product.

Polycomb group complexes

A group of multi-component protein complexes, first described in Drosophila melanogaster, that establish regions of chromatin in which gene expression is repressed.

Cyclin-dependent kinases

A family of serine–threonine protein kinases whose catalytic activity depends on their association with a cyclin subunit. They are involved in regulating cell-cycle progression.

Ankyrin repeat

A structural motif that comprises repeated modules of 33 amino acids that form a series of anti-parallel helix–loop–helix domains linked by β-hairpin loops. The motif was first identified in the protein ankyrin but is found in many other proteins with diverse functions.

INK4 proteins

A group of small (15–19 kDa) ankyrin-repeat proteins that bind specifically to the cyclin-dependent kinases CDK4 and CDK6, and block their association with regulatory D cyclins.

CIP/KIP family

A family of cyclin-dependent kinase inhibitors (designated p21CIP1, p27KIP1 and p57KIP2 in human cells) that bind to cyclin D-, cyclin E and cyclin A-dependent kinase complexes.

RAS

First identified in retroviruses that induce rat sarcomas (hence the name), the RAS genes encode G proteins that function in diverse signal transduction pathways and sustain activating point mutations in a high proportion of human cancers.

RAF–MEK–ERK

A kinase cascade that represents one of the major signal transduction pathways downstream of RAS, culminating in the phosphorylation of ERK (extracellular signal-regulated kinase). ERKs are classed as MAP kinases (mitogen-activated protein kinases), MEK is a MAPK kinase (MAPKK) and RAF is a MAPKK kinase.

Naevi

Raised, red/brown areas of skin, commonly referred to as moles, that represent benign precursors of melanomas.

p38 family

The four members of the p38 family, which are also known as stress-activated protein kinases (SAPKs), are a group of MAP kinases that are activated following exposure to cytokines, heat shock, high osmolarity and other cellular stresses.

MYC

MYC is a transcription factor that binds to DNA (at Ebox elements) as part of a heterodimeric complex with MAX. The gene is frequently overexpressed in human cancers as a consequence of DNA amplification, chromosomal translocation or mutations in the coding domain.

E-box element

A hexanucleotide sequence (CACGTG) that forms the core of promoter elements that are recognized by various basic helix–loop–helix (bHLH) transcription factors, including MYC and E2A.

BMI1

Initially identified as a common site of proviral integration in Moloney-virus-induced Bcell lymphomas, the BMI1 gene product was subsequently recognized as a member of the Polycomb group and a repressor of INK4a–ARF.

MDM2

Located on a mouse double minute (hence the name MDM) chromosome, the oncogene MDM2 encodes an E3 ubiquitin ligase that promotes the proteasome-mediated destruction of p53 and is transcriptionally activated by p53.

E2F family

A group of transcription factors that control the expression of genes that are involved in cell-cycle progression, by acting either as transcriptional repressors, in conjunction with members of the retinoblastoma family or as transcriptional activators.

AP1 family

Originally defined as 'activator protein-1', the generic term AP1 refers to a family of heterodimeric transcription factors that comprises members of the FOS (FOS, FRA1, FRA2), JUN (c-JUN, JUNB, JUNC, JUND) and ATF families.

ID proteins

A family of transcriptional regulators that lack the basic domain that is present in basic helix–loop–helix (bHLH) proteins and are therefore unable to bind to DNA. They function as dominant-negative inhibitors of certain bHLH proteins.

Basic helix–loop–helix domain

A protein structure motif that is found in various transcription factors, and that consists of a basic domain, which is responsible for sequence-specific DNA binding, followed by a helix–loop–helix domain that functions as a dimerization interface.

POZ/BTB domain

A protein–protein interaction domain that is found in various Pox virus and zinc-finger proteins. POZ domains from several zinc-finger proteins have been shown to mediate transcriptional repression and to interact with components of histone deacetylase corepressor complexes.

Chromodomain

A protein domain that is conserved in a wide number of chromatin-interacting proteins. The chromodomain is involved in binding to specific methylated lysine residues in the histone tails.

T-box proteins

The T-box family of transcription factors are involved in cell-fate decisions during development and contain a region (the T box) that is required for binding to the DNA consensus sequence TCACACCT.

CDC6

A cell-cycle-regulated protein that binds to origins of DNA replication and is involved in ensuring that DNA replication occurs only once during cell division.

Histone deacetylase

An enzyme that removes the acetyl groups from core histones that are generally associated with transcriptionally active chromatin. Deacetylation of the histone tails can allow access to histone methyltransferases that establish repressive chromatin.

Heterochromatin

A highly condensed and transcriptionally less active form of chromatin that occurs at defined sites, such as centromeres, silencer DNA elements or telomeres.

Rights and permissions

Reprints and permissions

About this article

Cite this article

Gil, J., Peters, G. Regulation of the INK4b–ARF–INK4a tumour suppressor locus: all for one or one for all. Nat Rev Mol Cell Biol 7, 667–677 (2006). https://doi.org/10.1038/nrm1987

Download citation

  • Issue Date:

  • DOI: https://doi.org/10.1038/nrm1987

This article is cited by

Search

Quick links

Nature Briefing

Sign up for the Nature Briefing newsletter — what matters in science, free to your inbox daily.

Get the most important science stories of the day, free in your inbox. Sign up for Nature Briefing