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:

Ataxia-telangiectasia: from a rare disorder to a paradigm for cell signalling and cancer

A Corrigendum to this article was published on 01 December 2008

Key Points

  • Ataxia-telangiectasia (A-T) is a rare human genetic disorder that is characterized by cancer predisposition, neurodegeneration, immunodeficiency, radiosensitivity, chromosomal instability and cell-cycle checkpoint defects.

  • A-T mutated (ATM) protein is a member of the phosphoinositide 3-kinase (PI3K)-related protein kinase (PIKK) family of protein kinases and has a central role in the response to DNA double-strand breaks (DSBs).

  • ATM responds to breaks introduced by DNA-damaging agents and to physiologically induced breaks.

  • ATM is activated by dissociation of an inactive dimer to form an active monomeric form of the protein.

  • Autophosphorylation and acetylation contribute to ATM activation.

  • Partial activation of ATM occurs in response to agents that alter chromatin structure, and full activation is achieved when it is recruited to the MRE11–RAD50–NBS1 (MRN) complex at the break site.

  • Recruitment of ATM to a DNA DSB occurs as part of an assembly complex with other proteins, including H2AX, MDC1, RNF8 and BRCA1.

  • The MRN complex assists in the activation of ATM and in time becomes a substrate (through NBS1) for active ATM.

  • Once activated, ATM phosphorylates a multitude of substrates that are involved in cell-cycle checkpoint activation and DNA repair. The MRN complex has an adaptor role, at least in the case of some of these substrates.

Abstract

First described over 80 years ago, ataxia-telangiectasia (A-T) was defined as a clinical entity 50 years ago. Although not encountered by most clinicians, it is a paradigm for cancer predisposition and neurodegenerative disorders and has a central role in our understanding of the DNA-damage response, signal transduction and cell-cycle control. The discovery of the protein A-T mutated (ATM) that is deficient in A-T paved the way for rapid progress on understanding how ATM functions with a host of other proteins to protect against genome instability and reduce the risk of cancer and other pathologies.

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: Activation of ATM.
Figure 2: The MRN complex tethers DNA broken ends.
Figure 3: Assembly of DNA-damage response proteins at a double-strand break.
Figure 4: ATM activation and signalling to downstream substrates in response to DNA double-strand breaks.

Similar content being viewed by others

References

  1. Boder, E. Ataxia-telangiectasia: an overview. Kroc. Found. Ser. 19, 1–63 (1985). An excellent description of the clinical phenotype in A-T particularly for the neurological defects.

    CAS  PubMed  Google Scholar 

  2. Syllaba, K. & Henner, K. Contribution a l'independence de l'athetose double idiopathique et congenitale. Atteinte familiale, syndrome dystrophique, signe de reseau vasculaire conjonctival, integrite psychique. Rev. Neurol. 1, 541–562 (1926) (in French).

    Google Scholar 

  3. Boder, E. & Sedgwick, R. P. Ataxia-telangiectasia. A familial syndrome of progressive cerebellar ataxia, oculocutaneous telangiectasia and frequent pulmonary infection. A preliminary report on 7 children, an autopsy, and a case history. Univ. S. Calif. Med. Bull. 9, 15–28 (1957).

    Google Scholar 

  4. Lavin, M. F. & Shiloh, Y. The genetic defect in ataxia-telangiectasia. Annu. Rev. Immunol. 15, 177–202 (1997). A comprehensive early review on A-T and the cloning and characteristics of the ATM gene.

    Article  CAS  PubMed  Google Scholar 

  5. Gatti, R. A. et al. Localization of an ataxia-telangiectasia gene to chromosome 11q22–23. Nature 336, 577–580 (1988).

    Article  CAS  PubMed  Google Scholar 

  6. Savitsky, K. et al. A single ataxia-telangiectasia gene with a product similar to PI-3 kinase kinase. Science. 268, 1749–1753 (1995). Describes the identification of the ATM gene, which is defective in A-T.

    Article  CAS  PubMed  Google Scholar 

  7. Brumbaugh, K. M. et al. The mRNA surveillance protein hSMG-1 functions in genotoxic stress response pathways in mammalian cells. Mol. Cell 14, 585–598 (2004).

    Article  CAS  PubMed  Google Scholar 

  8. Yamashita, A. et al. Human SMG-1, a novel phosphatidylinositol 3-kinase-related protein kinase, associates with components of the mRNA surveillance complex and is involved in the regulation of nonsense-mediated mRNA decay. Genes Dev. 15, 2215–2228 (2001).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  9. Banin, S. et al. Enhanced phosphorylation of p53 by ATM in response to DNA damage. Science 281, 1674–1677 (1998).

    Article  CAS  PubMed  Google Scholar 

  10. Canman, C. E. et al. Activation of the ATM kinase by ionizing radiation and phosphorylation of p53. Science 281, 1677–1679 (1998).

    Article  CAS  PubMed  Google Scholar 

  11. Khanna K. K. et al. ATM associates with and phosphorylates p53: mapping the region of interaction. Nature Genet. 20, 398–400 (1998). References 9–11 identified the first substrate for ATM, p53.

    Article  CAS  PubMed  Google Scholar 

  12. Kastan, M. B. et al. A mammalian cell cycle checkpoint pathway utilising p53 and GADD45 is defective in ataxia-telangiectasia. Cell 71, 587–597 (1992). Description of a radiation-induced pathway that connects p53 with the A-T gene product.

    Article  CAS  PubMed  Google Scholar 

  13. Khanna, K. K. & Lavin, M. F. Ionizing radiation and UV induction of p53 protein by different pathways in ataxia-telangiectasia cells. Oncogene 8, 3307–3312 (1993).

    CAS  PubMed  Google Scholar 

  14. Matsuoka, S. et al. ATM and ATR substrate analysis reveals extensive protein networks responsive to DNA damage. Science 316, 1160–1166 (2007).

    Article  CAS  PubMed  Google Scholar 

  15. Linding, R. et al. Systematic discovery of in vivo phosphorylation networks. Cell 129, 1415–1426 (2007).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  16. Bosotti, R., Isacchi, A. & Sonnhammer, E. L. FAT: a novel domain in PIK-related kinases. Trends Biochem. Sci. 25, 225–227 (2000).

    Article  CAS  PubMed  Google Scholar 

  17. Shafman, T. et al. Interaction between ATM protein and c-Abl in response to DNA damage. Nature 387, 520–523 (1997). ABL kinase is identified as an ATM interacting protein. See also reference 116.

    Article  CAS  PubMed  Google Scholar 

  18. Carney, J. P. et al. The hMre11/hRad50 protein complex and Nijmegen breakage syndrome: linkage of double-strand break repair to the cellular DNA damage response. Cell 93, 477–486 (1998).

    Article  CAS  PubMed  Google Scholar 

  19. Matsuura, S. et al. Positional cloning of the gene for Nijmegen breakage syndrome. Nature Genet. 19, 179–181 (1998).

    Article  CAS  PubMed  Google Scholar 

  20. Varon, R. et al. Nibrin, a novel DNA double-strand break repair protein, is mutated in Nijmegen breakage syndrome. Cell 93, 467–476 (1998).

    Article  CAS  PubMed  Google Scholar 

  21. Stewart, G. S. et al. The DNA double-strand break repair gene hMRE11 is mutated in individuals with an ataxia-telangiectasia-like disorder. Cell 99, 577–587 (1999).

    Article  CAS  PubMed  Google Scholar 

  22. Houldsworth, J. & Lavin, M. F. Effect of ionizing radiation on DNA synthesis in ataxia telangiectasia cells. Nucleic Acids Res. 8, 3709–3720 (1980). First description of a cell-cycle checkpoint defect (radioresistant DNA synthesis) in A-T cells. See also reference 23.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  23. Painter, R. B. & Young, B. R. Radiosensitivity in ataxia-telangiectasia: a new explanation. Proc. Natl Acad. Sci. USA 77, 7315–7317 (1980).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  24. Beamish, H. & Lavin, M. F. Radiosensitivity in ataxia-telangiectasia: anomalies in radiation-induced cell cycle delay. Int. J. Radiat. Biol. 65, 175–184 (1994).

    Article  CAS  PubMed  Google Scholar 

  25. Scott, D. & Zampetti-Bosseler, F. Cell cycle dependence of mitotic delay in X-irradiated normal and ataxia-telangiectasia fibroblasts. Int. J. Radiat. Biol. 42, 679–683 (1982).

    CAS  Google Scholar 

  26. Jaspers, N. G. & Bootsma, D. Abnormal levels of UV-induced unscheduled DNA synthesis in ataxia telangiectasia cells after exposure to ionizing radiation. Mutat. Res. 1–2, 439–446 (1982).

    Article  Google Scholar 

  27. Lange, E. et al. Localization of an ataxia-telangiectasia gene to an approximately 500-kb interval on chromosome 11q23.1: linkage analysis of 176 families by an international consortium. Am. J. Hum. Genet. 1, 112–119 (1995).

    Google Scholar 

  28. Savitsky, K. et al. The complete sequence of the coding region of the ATM gene reveals similarity to cell cycle regulators in different species. Hum. Mol. Genet. 4, 2025–2032 (1995).

    Article  CAS  PubMed  Google Scholar 

  29. Uziel, T. et al. Genomic organization of the ATM gene. Genomics 2, 317–320 (1996).

    Article  Google Scholar 

  30. Chun, H. H. & Gatti, R. A. Ataxia-telangiectasia, an evolving phenotype. DNA Repair 3, 1187–1196 (2004).

    Article  CAS  PubMed  Google Scholar 

  31. Gotoff, S. P., Amirmokri, E. & Liebner, E. J. Ataxia-telangiectasia. Neoplasia, untoward response to X-irradiation, and tuberous sclerosis. Am. J. Dis. Child. 114, 617–625 (1967).

    Article  CAS  PubMed  Google Scholar 

  32. Taylor, A. M. et al. Ataxia-telangiectasia: a human mutation with abnormal radiation sensitivity. Nature 4, 427–429 (1975).

    Article  Google Scholar 

  33. Chen, P. C., Lavin, M. F., Kidson, C. & Moss, D. Identification of ataxia telangiectasia heterozygotes, a cancer prone population. Nature 274, 484–486 (1978).

    Article  CAS  PubMed  Google Scholar 

  34. Riballo, E. et al. A pathway of double-strand break rejoining dependent upon ATM, Artemis, and proteins locating to γ-H2AX foci. Mol. Cell 16, 715–724 (2004).

    Article  CAS  PubMed  Google Scholar 

  35. Foray, N. et al. Hypersensitivity of ataxia telangiectasia fibroblasts to ionizing radiation is associated with a repair deficiency of DNA double-strand breaks. Int. J. Radiat. Biol. 72, 271–283 (1997).

    Article  CAS  PubMed  Google Scholar 

  36. Kuhne, M., Rothkamm, K. & Lobrich, M. No dose-dependence of DNA double-strand break misrejoining following alpha-particle irradiation. Int. J. Radiat. Biol. 76, 891–900 (2000).

    Article  CAS  PubMed  Google Scholar 

  37. Goodarzi, A. A. et al. ATM signalling facilitates repair of DNA double-strand breaks associated with heterochromatin. Mol. Cell 31, 167–177 (2008).

    Article  CAS  PubMed  Google Scholar 

  38. Berkovich, E., Monnat, R. J. Jr, & Kastan, M. B. Roles of ATM and NBS1 in chromatin structure modulation and DNA double-strand break repair. Nature Cell Biol. 9, 683–690 (2007). A model system to investigate the series of events associated with ATM activation.

    Article  CAS  PubMed  Google Scholar 

  39. Yazdi, P. T. et al. SMC1 is a downstream effector in the ATM/NBS1 branch of the human S-phase checkpoint. Genes Dev. 16, 571–582 (2002).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  40. Kitagawa, R., Bakkenist, C. J., McKinnon, P. J. & Kastan, M. B. Phosphorylation of SMC1 is a critical downstream event in the ATM–NBS1–BRCA1 pathway. Genes Dev. 18, 1423–1438 (2004).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  41. Lumsden, J. M. et al. Immunoglobulin class switch recombination is impaired in Atm-deficient mice. J. Exp. Med. 200, 1111–1121 (2004).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  42. Bredemeyer, A. L. et al. ATM stabilizes DNA double-strand-break complexes during V(D)J recombination. Nature 442, 466–470 (2006).

    Article  CAS  PubMed  Google Scholar 

  43. Vecchio, M. S., Olaru, A., Livak, F. & Hodes, R. J. ATM deficiency impairs thymocyte maturation because of defective resolution of T cell receptor α locus coding end breaks. Proc. Natl Acad. Sci. USA 104, 6323–6328 (2007).

    Article  CAS  Google Scholar 

  44. Perkins, E. J. et al. Sensing of intermediates V(D)J recombination by ATM. Genes Dev. 16, 159–164 (2002).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  45. Matei, I. R. et al. ATM deficiency disrupts TCRα locus integrity and the maturation of CD4+CD8+ thymocytes. Blood 109, 1887–1896 (2007).

    Article  CAS  PubMed  Google Scholar 

  46. Lim, D. S. et al. ATM binds to β-adaptin in cytoplasmic vesicles. Proc. Natl Acad. Sci. USA. 95, 10146–10151 (1998).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  47. Watters, D. et al. Localization of a portion of extranuclear ATM to peroxisomes. J. Biol. Chem. 274, 34277–34282 (1999).

    Article  CAS  PubMed  Google Scholar 

  48. Lavin, M. F. et al. in Handbook of Cell Signalling (eds Bradshaw R. A. and Dennis, E. A.) 225–236 (Academic, London, 2003).

    Book  Google Scholar 

  49. Yang, D.Q. & Kastan, M. B. Participation of ATM in insulin signalling through phosphorylation of eIF-4E-binding protein 1. Nature Cell Biol. 2, 893–898 (2000).

    Article  CAS  PubMed  Google Scholar 

  50. Schlach, D. S., McFarlin, D. & Barlow, M. H. An unusual form of diabetes mellitus in ataxia telangiectasia. N. Engl. J. Med. 282, 1396–1402 (1970).

    Article  Google Scholar 

  51. Bar, R. S. et al. Extreme insulin resistance in ataxia telangiectasia: defect in affinity of insulin receptors. N. Engl. J. Med. 298, 1164–1171 (1978).

    Article  CAS  PubMed  Google Scholar 

  52. Schneider, J. G. et al. ATM-dependent suppression of stress signaling reduces vascular disease in metabolic syndrome. Cell. Metab. 4, 377–389 (2006).

    Article  CAS  PubMed  Google Scholar 

  53. Bakkenist, C. J. & Kastan, M. B. DNA damage activates ATM through intermolecular autophosphorylation and dimer dissociation. Nature 421, 499–506 (2003). A major contribution to understanding the mechanism of ATM activation.

    Article  CAS  PubMed  Google Scholar 

  54. Barlow, C. et al. ATM is a cytoplasmic protein in mouse brain required to prevent lysosomal accumulation. Proc. Natl Acad. Sci. USA 97, 871–876 (2000).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  55. Biton, S. et al. ATM-mediated response to DNA double strand breaks in human neurons derived from stem cells. DNA Repair 6, 128–134 (2006).

    Article  PubMed  CAS  Google Scholar 

  56. Dar, I. et al. Analysis of the ataxia telangiectasia mutated-mediated DNA damage response in murine cerebellar neurons. J. Neurosci. 26, 7767–7674 (2006).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  57. Taylor, A. M. R., Groom, A. & Byrd, P. J. Ataxia-telangiectasia-like disorder (ATLD) — its clinical presentation and molecular basis. DNA Repair. 3, 1219–1225 (2004).

    Article  CAS  PubMed  Google Scholar 

  58. Cimprich, K. A. & Cortez, D. ATR: an essential regulator of genome integrity. Nature Rev. Mol. Cell Biol. 9, 616–627 (2008).

    CAS  Google Scholar 

  59. Huen, M. S. Y. & Chen, J. The DNA damage response pathways: at the crossroad of protein modifications. Cell Res. 18, 8–16 (2008).

    Article  CAS  PubMed  Google Scholar 

  60. You, Z. et al. Rapid activation of ATM on DNA flanking double-strand breaks. Nature Cell Biol. 9, 1311–1318 (2007).

    Article  CAS  PubMed  Google Scholar 

  61. de Jager, M. et al. Human Rad50/Mre11 is a flexible complex that can tether DNA ends. Mol. Cell 8, 1129–1135 (2001).

    Article  CAS  PubMed  Google Scholar 

  62. van den Bosch, M., Bree, R. T. & Lowndes, N. F. The MRN complex: coordinating and mediating the response to broken chromosomes. EMBO Rep. 4, 844–849 (2003).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  63. Hopfner, K. P., Craig, L., Moncalian, G., Zinkel, R. A. et al. The Rad50 zinc-hook is a structure joining Mre11 complexes in DNA recombination repair. Nature 418, 562–566 (2002).

    Article  CAS  PubMed  Google Scholar 

  64. Moreno-Herrero, F. et al. Mesoscale conformational changes in the DNA-repair complex Rad50/Mre11/Nbs1 upon binding DNA. Nature 437, 440–443 (2005).

    Article  CAS  PubMed  Google Scholar 

  65. Paull, T. T. & Gellert, M. The 3′ to 5′ exonuclease activity of Mre 11 facilitates repair of DNA double-strand breaks. Mol. Cell 1, 969–979 (1998).

    Article  CAS  PubMed  Google Scholar 

  66. Paull, T. T. & Gellert, M. Nbs1 potentiates ATP-driven DNA unwinding and endonuclease cleavage by the Mre11/Rad50 complex. Genes Dev. 13, 1276–1288 (1999).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  67. Mirzoeva, O. K. & Petrini, J. H. DNA replication-dependent nuclear dynamics of the Mre11 complex. Mol. Cancer Res. 1, 207–218 (2003).

    CAS  PubMed  Google Scholar 

  68. Uziel, T. et al. Requirement of the MRN complex for ATM activation by DNA damage. EMBO J. 22, 5612–5621 (2003). Shows the importance of the MRN complex as an upstream activator of ATM.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  69. Cerosaletti, K. & Concannon, P. Independent roles for nibrin and Mre11–Rad50 in the activation and function of Atm. J. Biol. Chem. 279, 38813–38819 (2004).

    Article  CAS  PubMed  Google Scholar 

  70. Digweed, M., Reis, A. & Sperling, K. Nijmegen breakage syndrome: consequences of defective DNA double strand break repair. Bioessays 21, 649–656 (1999).

    Article  CAS  PubMed  Google Scholar 

  71. Falck, J., Coates, J. & Jackson, S. P. Conserved modes of recruitment of ATM, ATR and DNA-PKcs to sites of DNA damage. Nature 434, 605–611 (2005). Describes a role for NBS1 in the recruitment of ATM to sites of DNA damage.

    Article  CAS  PubMed  Google Scholar 

  72. Cerosaletti, K., Wright, J. & Concannon, P. Active role for nibrin in the kinetics of ATM activation. Mol. Cell. Biol. 26, 1691–1699 (2005).

    Article  CAS  Google Scholar 

  73. Kozlov, S. V. et al. Involvement of novel autophosphorylation sites in ATM activation. EMBO J. 25, 3504–3514 (2006).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  74. Goodarzi, A. A., et al. Autophosphorylation of ataxia-telangiectasia mutated is regulated by protein phosphatase 2A. EMBO J. 23, 4451–4461 (2004).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  75. Shreeram, S. et al. Wip1 phosphatase modulates ATM-dependent signalling pathways. Mol. Cell 23, 757–764 (2006).

    Article  CAS  PubMed  Google Scholar 

  76. Ali, A., et al. Requirement of protein phosphatase 5 in DNA-damage-induced ATM activation. Genes Dev. 18, 249–254 (2004).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  77. Sun, Y. et al. A role for the Tip60 histone acetyltransferase in the acetylation and activation of ATM. Proc. Natl Acad. Sci. USA 102, 13182–13187 (2005).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  78. Sun, Y. et al. DNA damage-induced acetylation of lysine 3016 of ATM activates ATM kinase activity. Mol. Cell. Biol. 27, 8502–8509 (2007).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  79. Pellegrini, M. et al. Autophosphorylation at serine 1987 is dispensable for murine Atm activation in vivo. Nature 443, 222–225 (2006). Describes a mouse model that questions the importance of autophosphorylation in ATM activation.

    Article  CAS  PubMed  Google Scholar 

  80. Hamer, G. et al. Ataxia telangiectasia mutated expression and activation of the testis. Biol. Reprod. 70, 1206–1212 (2004).

    Article  CAS  PubMed  Google Scholar 

  81. Powers, J. T. et al. E2F1 uses the ATM signalling pathway to induce p53 and Chk2 phosphorylation and apoptosis. Mol. Cancer Res. 2, 203–214 (2004).

    CAS  PubMed  Google Scholar 

  82. Kurz, E. U., Douglas, P. & Lees-Miller, S. Doxorubicin activates ATM-dependent phosphorylation of multiple downstream targets in part through the generation of reactive oxygen species. J. Biol. Chem. 279, 53272–53281 (2004).

    Article  CAS  PubMed  Google Scholar 

  83. Shiloh, Y. The ATM-mediated DNA-damage response: taking shape. Trends Biochem. Sci. 31, 402–410 (2006).

    Article  CAS  PubMed  Google Scholar 

  84. Beausoleil, S. A. et al. Large-scale characterisation of HeLa cell nuclear phosphoproteins. Proc. Natl Acad. Sci. USA 101, 12130–12135 (2004).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  85. Lukas, C. et al. Mdc1 couples DNA double-strand break recognition by Nbs1 with its H2AX-dependent chromatin retention. EMBO J. 23, 2674–2683 (2004).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  86. Jacob, N. K., Stout, A. R. & Price, C. M. Modulation of telomere length dynamics by the subtelomeric region of tetrahymena telomeres. Mol. Biol. Cell 15, 3719–3728 (2004).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  87. Bassing, C. H. et al. Histone H2AX: a dosage-dependent suppressor of oncogenic translocations and tumors. Cell 114, 359–370 (2003).

    Article  CAS  PubMed  Google Scholar 

  88. Celeste, A. et al. H2AX haploinsufficiency modifies genomic stability and tumor susceptibility. Cell 114, 371–383 (2003).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  89. Stucki, M. & Jackson, S. P. MDC1/NFBD1: A key regulator of the DNA damage response in higher eukaryotes. DNA Repair 3, 953–957 (2004).

    Article  CAS  PubMed  Google Scholar 

  90. Chapman, J. R. & Jackson, S. P. Phospho-dependent interactions between Nbs1 and MDC1 mediate chromatin retention of the MRN complex at sites of DNA damage. EMBO Rep. 9, 795–801 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  91. Melander, F. et al. Phosphorylation of SDT repeats in the MDC1 N terminus triggers retention of NBS1 at the DNA damage-modified chromatin. J. Cell Biol. 181, 213–226 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  92. Huen, M. S. et al. RNF8 transduces the DNA-damage signal via histone ubiquitylation and checkpoint protein assembly. Cell 131, 901–914 (2007).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  93. Kolas, N. K. et al. Orchestration of the DNA-damage response by the RNF8 ubiquitin ligase. Science 318, 1637–1640 (2007).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  94. Mailand, N. et al. RNF8 ubiquitylates histones at DNA double-strand breaks and promotes assembly of repair proteins. Cell 131, 887–900 (2007).

    Article  CAS  PubMed  Google Scholar 

  95. Lavin, M. F. ATM and the Mre11 complex combine to recognize and signal DNA double-strand breaks. Oncogene 26, 7749–7758 (2007).

    Article  CAS  PubMed  Google Scholar 

  96. Buscemi, G. et al. Chk2 activation dependence on Nbs1 after DNA damage. Mol. Cell. Biol. 21, 5214–5222 (2001).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  97. Nakanishi, K. et al. Interaction of FANCD2 and NBS1 in the DNA damage response. Nature Cell Biol. 4, 913–920 (2002).

    Article  CAS  PubMed  Google Scholar 

  98. Gatei, M. et al. Ataxia-telangiectasia-mutated (ATM) and NBS1-dependent phosphorylation of Chk1 on Ser-317 in response to ionizing radiation. J. Biol. Chem. 278, 14806–14811 (2003).

    Article  CAS  PubMed  Google Scholar 

  99. Spycher, C. et al. Constitutive phosphorylation of MDC1 physically links the Mre11–Rad50–Nbs1 complex to damaged chromatin. J. Cell Biol. 181, 227–240 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  100. Lou, Z., et al. MDC1 maintains genomic stability by participating in the amplification of ATM-dependent DNA damage signals. Mol. Cell 21, 187–200 (2006).

    Article  CAS  PubMed  Google Scholar 

  101. Lukas, C. et al. Distinct spatiotemporal dynamics of mammalian checkpoint regulators induced by DNA damage. Nature Cell Biol. 5, 255–260 (2003). Describes the use of microlasers to determine the order at which DNA-damage response proteins are recruited to chromatin.

    Article  CAS  PubMed  Google Scholar 

  102. Bekker-Jensen, S. et al. Spatial organization of the mammalian genome surveillance machinery in response to DNA strand breaks. J. Cell Biol. 173, 195–206 (2006).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  103. Zhao, S. et al. Functional link between ataxia-telangiectasia and Nijmegen breakage syndrome gene products. Nature 405, 473–477 (2000).

    Article  CAS  PubMed  Google Scholar 

  104. Wu, X. et al. ATM phosphorylation of Nijmegen breakage syndrome protein is required in a DNA damage response. Nature 405, 477–482 (2000).

    Article  CAS  PubMed  Google Scholar 

  105. Gatei, M., Young, D., Cerosaletti, K. M., Desai-Mehta, A. et al. ATM-dependent phosphorylation of nibrin in response to radiation exposure. Nature Genet. 25, 115–119 (2000).

    Article  CAS  PubMed  Google Scholar 

  106. Lim, D. S. et al. ATM phosphorylates p95/nbs1 in an S-phase checkpoint pathway. Nature 404, 613–617 (2000).

    Article  CAS  PubMed  Google Scholar 

  107. Lee, J. H. et al. Regulation of Mre11/Rad50 by Nbs1: effects on nucleotide-dependent DNA binding and association with ataxia-telangiectasia-like disorder mutant complexes. J. Biol. Chem. 278, 45171–45178 (2003).

    Article  CAS  PubMed  Google Scholar 

  108. Falck, J., Petrini, J. H., Williams, B. R., Lukas, J. The DNA damage-dependent intra-S phase checkpoint is regulated by parallel pathways. Nature Genet. 30, 290–294 (2002).

    Article  PubMed  Google Scholar 

  109. Lee, J. H. & Paull, T. T. ATM activation by DNA double-strand breaks through the Mre11–Rad50–Nbs1 complex. Science 308, 551–554 (2005).

    Article  CAS  PubMed  Google Scholar 

  110. Soutoglou, E. & Misteli, T. Activation of the cellular DNA damage response in the absence of DNA lesions. Science 320, 1507–1510 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  111. Halazonetis, T. D., Gorgoulis, V. G. & Bartek, J. An oncogene-induced DNA damage model for cancer development. Science 319, 1352–1355 (2008).

    Article  CAS  PubMed  Google Scholar 

  112. Swift, M., Morrell, D., Massey, R. B. & Chase, C. L. Incidence of cancer in 161 families affected by ataxia-telangiectasia. N. Engl. J. Med. 325, 1831–1836 (1991). First comprehensive report on increased cancer incidence in A-T heterozygotes in particular breast cancers.

    Article  CAS  PubMed  Google Scholar 

  113. Renwick, A. et al. ATM mutations that cause ataxia-telangiectasia are breast cancer susceptibility alleles. Nature Genet. 38, 873–875 (2008).

    Article  CAS  Google Scholar 

  114. Kim, S. T., Lim, D. S., Canman, C. E. & Kastan, M. B. Substrate specificities and identification of putative substrates of ATM kinase family members. J. Biol. Chem. 274, 37538–37543 (1999).

    Article  CAS  PubMed  Google Scholar 

  115. O'Neill, T. et al. Utilisation of oriented peptide libraries to identify substrate motifs selected by ATM. J. Biol. Chem. 275, 22719–22727 (2000).

    Article  CAS  PubMed  Google Scholar 

  116. Petersen, R., Kelly D. & Good, R. A. Ataxia-telangiectasia: its association with a defective thymus, immunological deficiency disease and malignancy. Lancet 1, 1189–1193 (1964).

    Article  Google Scholar 

  117. Spector, B. D., Filipovich, A. H., Perry, G.S. & Kersey, H. in Ataxia-Telangiectasia (eds Bridges, B. A. & Harnden, D. G.) 103–141 (John Wiley & Sons, 1982).

    Google Scholar 

  118. Cornforth, M. N. & Bedford, J. S. On the nature of a defect in cells from individuals with ataxia-telangiectasia. Science 1, 1589–1591 (1985). Describes the defective gene in A-T which is localized to chromosome 11q.

    Article  Google Scholar 

  119. Baskaran, R. et al. Ataxia telangiectasia mutant protein activates c-Abl tyrosine kinase in response to ionizing radiation. Nature 387, 516–519 (1997).

    Article  CAS  PubMed  Google Scholar 

  120. Smith, G. C. et al. Purification and DNA binding properties of the ataxia-telangiectasia gene product ATM. Proc. Natl Acad. Sci. USA. 96, 11134–11139 (1999).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  121. Chan, D. W. et al. Purification and characterisation of ATM from human placenta. A manganese-dependent, wortmannin-sensitive serine/threonine protein kinase. J. Biol. Chem. 275, 7803–7810 (2000).

    Article  CAS  PubMed  Google Scholar 

  122. Bartkova, J., Rezaei, N., Liontos, M., Karakaidos, P. et al. Oncogene-induced senescence is part of the tumorigenesis barrier imposed by DNA damage checkpoints. Nature 444, 633–637 (2006).

    Article  CAS  PubMed  Google Scholar 

  123. Gougoulis, V. G., Vassiliou, L. V., Karakaidos, P., Zacharatos, P. et al. Activation of the DNA damage checkpoint and genomic instability in human precancerous lesions. Nature 434, 907–913 (2005).

    Article  CAS  Google Scholar 

Download references

Author information

Authors and Affiliations

Authors

Related links

Related links

FURTHER INFORMATION

Leiden Open Variation Database

Martin Lavin's homepage

Glossary

Checkpoint

In this example, a cell-cycle checkpoint is a control mechanism to ensure that chromosomes are intact for cell division.

Adaptor

A protein that assists in the process of downstream signalling. In this case, NBS1 has an adaptor role in ATM signalling.

Radioresistant DNA synthesis

The absence of a steep component of inhibition of DNA synthesis in a dose–response curve when the rate of DNA synthesis is plotted against radiation doses.

Complementation group

This refers to previous studies on A-T cells in which cell fusion was used to determine whether more than one protein was involved in the defect.

Non-homologous end joining (NHEJ) repair

The repair of DNA double-strand breaks (free ends) by DNA-PK and other cofactors without the requirement for DNA recombination.

Heterochromatin

A cytologically distinct, tightly packed form of chromatin in which transcription is repressed.

V(D)J recombination

A mechanism that involves gene rearrangement in the maturation of immunoglobulin and T-cell receptor genes.

Peroxisomes

Subcellular organelles that metabolize fatty acids and clear the cell of toxic peroxides.

Endosomes

Vesicles derived from the plasma membrane to transport proteins and other substances into the cell.

Purkinje cell

Neuronal cells found on the cerebellum between the molecular and granular layers.

Autophosphorylation

The process by which a protein kinase uses itself as a substrate for phosphorylation in many cases for self-activation.

Downstream signalling

Part of a signal transduction pathway in which a protein, such as ATM, phosphorylates a series of substrates that assist in controlling various cellular processes.

FHA domain

(Forkhead-associated domain). A short sequence of amino acids that binds to phosphothreonine residues on various proteins.

ATPase domain

The region of a protein that is responsible for the hydrolysis of ATP and is coupled to protein phosphorylation.

Exonuclease

The enzyme that degrades nucleic acids from 3′ to 5′ free ends.

Endonuclease

An enzyme that is capable of hydrolysing phosphodiester bonds in nucleic acids that are away from the free ends.

Ubiquitylation

Post-translational modification of a protein by the covalent attachment of a ubiquitin protein to enable degradation or other forms of regulation.

Rights and permissions

Reprints and permissions

About this article

Cite this article

Lavin, M. Ataxia-telangiectasia: from a rare disorder to a paradigm for cell signalling and cancer. Nat Rev Mol Cell Biol 9, 759–769 (2008). https://doi.org/10.1038/nrm2514

Download citation

  • Issue Date:

  • DOI: https://doi.org/10.1038/nrm2514

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