1932

Abstract

The ataxia-telangiectasia mutated (ATM) protein kinase is a master regulator of the DNA damage response, and it coordinates checkpoint activation, DNA repair, and metabolic changes in eukaryotic cells in response to DNA double-strand breaks and oxidative stress. Loss of ATM activity in humans results in the pleiotropic neurodegeneration disorder ataxia-telangiectasia. ATM exists in an inactive state in resting cells but can be activated by the Mre11–Rad50–Nbs1 (MRN) complex and other factors at sites of DNA breaks. In addition, oxidation of ATM activates the kinase independently of the MRN complex. This review discusses these mechanisms of activation, as well as the posttranslational modifications that affect this process and the cellular factors that affect the efficiency and specificity of ATM activation and substrate phosphorylation. I highlight functional similarities between the activation mechanisms of ATM, phosphatidylinositol 3-kinases (PI3Ks), and the other PI3K-like kinases, as well as recent structural insights into their regulation.

Loading

Article metrics loading...

/content/journals/10.1146/annurev-biochem-060614-034335
2015-06-02
2024-04-19
Loading full text...

Full text loading...

/deliver/fulltext/biochem/84/1/annurev-biochem-060614-034335.html?itemId=/content/journals/10.1146/annurev-biochem-060614-034335&mimeType=html&fmt=ahah

Literature Cited

  1. Anheim M, Tranchant C, Koenig M. 1.  2012. The autosomal recessive cerebellar ataxias. N. Engl. J. Med. 366:636–46 [Google Scholar]
  2. Spacey SD, Gatti RA, Bebb G. 2.  2000. The molecular basis and clinical management of ataxia telangiectasia. Can. J. Neurol. Sci. 27:184–91 [Google Scholar]
  3. Perlman SL, Boder E, Sedgewick RP, Gatti RA. 3.  2012. Ataxia-telangiectasia. Handb. Clin. Neurol. 103:307–32 [Google Scholar]
  4. Lavin MF. 4.  2008. Ataxia-telangiectasia: from a rare disorder to a paradigm for cell signalling and cancer. Nat. Rev. Mol. Cell Biol. 9:759–69 [Google Scholar]
  5. Uhrhammer N, Bay J-O, Perlman S, Gatti RA. 5.  2002. Ataxia-telangiectasia and variants. Atlas Genet. Cytogenet. Oncol. Haematol. 6:316–25 [Google Scholar]
  6. Savitsky K, Bar-Shira A, Gilad S, Rotman G, Ziv Y. 6.  et al. 1995. A single ataxia telangiectasia gene with a product similar to PI-3 kinase. Science 268:1749–53 [Google Scholar]
  7. Gatti RA, Berkel I, Boder E, Braedt G, Charmley P. 7.  et al. 1988. Localization of an ataxia-telangiectasia gene to chromosome 11q22–23. Nature 336:577–80 [Google Scholar]
  8. Savitsky K, Sfez S, Tagle DA, Ziv Y, Sartiel A. 8.  et al. 1995. The complete sequence of the coding region of the ATM gene reveals similarity to cell cycle regulators in different species. Hum. Mol. Genet. 4:2025–32 [Google Scholar]
  9. Jung M, Kondratyev A, Lee SA, Dimtchev A, Dritschilo A. 9.  1997. ATM gene product phosphorylates IκB-α. Cancer Res. 57:24–27 [Google Scholar]
  10. Lovejoy CA, Cortez D. 10.  2009. Common mechanisms of PIKK regulation. DNA Repair 8:1004–8 [Google Scholar]
  11. Kim ST, Lim DS, Canman CE, Kastan MB. 11.  1999. Substrate specificities and identification of putative substrates of ATM kinase family members. J. Biol. Chem. 274:37538–43 [Google Scholar]
  12. Abraham RT. 12.  2004. PI 3-kinase related kinases: ‘big’ players in stress-induced signaling pathways. DNA Repair 3:883–87 [Google Scholar]
  13. Vadas O, Burke JE, Zhang X, Berndt A, Williams RL. 13.  2011. Structural basis for activation and inhibition of class I phosphoinositide 3-kinases. Sci. Signal. 4:re2 [Google Scholar]
  14. Lempiainen H, Halazonetis TD. 14.  2009. Emerging common themes in regulation of PIKKs and PI3Ks. EMBO J. 28:3067–73 [Google Scholar]
  15. Perry J, Kleckner N. 15.  2003. The ATRs, ATMs, and TORs are giant HEAT repeat proteins. Cell 112:151–55 [Google Scholar]
  16. Andrade MA, Petosa C, O'Donoghue SI, Müller CW, Bork P. 16.  2001. Comparison of ARM and HEAT protein repeats. J. Mol. Biol. 309:1–18 [Google Scholar]
  17. Stracker TH, Petrini JH. 17.  2011. The MRE11 complex: starting from the ends. Nat. Rev. Mol. Cell Biol. 12:90–103 [Google Scholar]
  18. Williams RS, Williams JS, Tainer JA. 18.  2007. Mre11–Rad50–Nbs1 is a keystone complex connecting DNA repair machinery, double-strand break signaling, and the chromatin template. Biochem. Cell Biol. 85:509–20 [Google Scholar]
  19. Paull TT. 19.  2010. Making the best of the loose ends: Mre11/Rad50 complexes and Sae2 promote DNA double-strand break resection. DNA Repair 9:1283–91 [Google Scholar]
  20. Symington LS, Gautier J. 20.  2011. Double-strand break end resection and repair pathway choice. Annu. Rev. Genet. 45:247–71 [Google Scholar]
  21. Paull TT, Deshpande RA. 21.  2014. The Mre11/Rad50/Nbs1 complex: recent insights into catalytic activities and ATP-driven conformational changes. Exp. Cell Res. 329139–47
  22. Stewart GS, Maser RS, Stankovic T, Bressan DA, Kaplan MI. 22.  et al. 1999. The DNA double-strand break repair gene hMre11 is mutated in individuals with an ataxia-telangiectasia-like disorder. Cell 99:577–87 [Google Scholar]
  23. 23.  Deleted in proof
  24. Luo G, Yao MS, Bender CF, Mills M, Bladl AR. 24.  et al. 1999. Disruption of mRad50 causes embryonic stem cell lethality, abnormal embryonic development, and sensitivity to ionizing radiation. PNAS 96:7376–81 [Google Scholar]
  25. Xiao Y, Weaver DT. 25.  1997. Conditional gene targeted deletion by Cre recombinase demonstrates the requirement for the double-strand break repair Mre11 protein in murine embryonic stem cells. Nucleic Acids Res. 25:2985–91 [Google Scholar]
  26. Zhu J, Petersen S, Tessarollo L, Nussenzweig A. 26.  2001. Targeted disruption of the Nijmegen breakage syndrome gene NBS1 leads to early embryonic lethality in mice. Curr. Biol. 11:105–9 [Google Scholar]
  27. Taylor AM, Groom A, Byrd PJ. 27.  2004. Ataxia-telangiectasia-like disorder (ATLD)—its clinical presentation and molecular basis. DNA Repair 3:1219–25 [Google Scholar]
  28. Uziel T, Lerenthal Y, Moyal L, Andegeko Y, Mittelman L, Shiloh Y. 28.  2003. Requirement of the MRN complex for ATM activation by DNA damage. EMBO J. 22:5612–21 [Google Scholar]
  29. Yamaguchi-Iwai Y, Sonoda E, Sasaki MS, Morrison C, Haraguchi T. 29.  et al. 1999. Mre11 is essential for the maintenance of chromosomal DNA in vertebrate cells. EMBO J. 18:6619–29 [Google Scholar]
  30. Adelman CA, De S, Petrini JH. 30.  2009. Rad50 is dispensable for the maintenance and viability of postmitotic tissues. Mol. Cell. Biol. 29:483–92 [Google Scholar]
  31. Shiloh Y. 31.  1997. Ataxia-telangiectasia and the Nijmegen breakage syndrome: related disorders but genes apart. Annu. Rev. Genet. 31:635–62 [Google Scholar]
  32. Varon R, Vissinga C, Platzer M, Cerosaletti KM, Chrzanowska KH. 32.  et al. 1998. Nibrin, a novel DNA double-strand break repair protein, is mutated in Nijmegen breakage syndrome. Cell 93:467–76 [Google Scholar]
  33. Frappart PO, McKinnon PJ. 33.  2006. Ataxia-telangiectasia and related diseases. Neuromol. Med. 8:495–511 [Google Scholar]
  34. Matsumoto Y, Miyamoto T, Sakamoto H, Izumi H, Nakazawa Y. 34.  et al. 2011. Two unrelated patients with MRE11A mutations and Nijmegen breakage syndrome–like severe microcephaly. DNA Repair 10:314–21 [Google Scholar]
  35. Waltes R, Kalb R, Gatei M, Kijas AW, Stumm M. 35.  et al. 2009. Human RAD50 deficiency in a Nijmegen breakage syndrome–like disorder. Am. J. Hum. Genet. 84:605–16 [Google Scholar]
  36. Paull TT, Gellert M. 36.  1999. Nbs1 potentiates ATP-driven DNA unwinding and endonuclease cleavage by the Mre11/Rad50 complex. Genes Dev. 13:1276–88 [Google Scholar]
  37. Lee J-H, Ghirlando R, Bhaskara V, Hoffmeyer MR, Gu J, Paull TT. 37.  2003. Regulation of Mre11/Rad50 by Nbs1: effects on nucleotide-dependent DNA binding and association with ATLD mutant complexes. J. Biol. Chem. 278:45171–81 [Google Scholar]
  38. Cerosaletti KM, Desai-Mehta A, Yeo TC, Kraakman–Van Der Zwet M, Zdzienicka MZ, Concannon P. 38.  2000. Retroviral expression of the NBS1 gene in cultured Nijmegen breakage syndrome cells restores normal radiation sensitivity and nuclear focus formation. Mutagenesis 15:281–86 [Google Scholar]
  39. Desai-Mehta A, Cerosaletti KM, Concannon P. 39.  2001. Distinct functional domains of nibrin mediate Mre11 binding, focus formation, and nuclear localization. Mol. Cell. Biol. 21:2184–91 [Google Scholar]
  40. Regal JA, Festerling TA, Buis JM, Ferguson DO. 40.  2013. Disease-associated MRE11 mutants impact ATM/ATR DNA damage signaling by distinct mechanisms. Hum. Mol. Genet. 22:5146–59 [Google Scholar]
  41. Shull ER, Lee Y, Nakane H, Stracker TH, Zhao J. 41.  et al. 2009. Differential DNA damage signaling accounts for distinct neural apoptotic responses in ATLD and NBS. Genes Dev. 23:171–80 [Google Scholar]
  42. Morales M, Theunissen JW, Kim CF, Kitagawa R, Kastan MB, Petrini JH. 42.  2005. The Rad50S allele promotes ATM-dependent DNA damage responses and suppresses ATM deficiency: implications for the Mre11 complex as a DNA damage sensor. Genes Dev. 19:3043–54 [Google Scholar]
  43. Roset R, Inagaki A, Hohl M, Brenet F, Lafrance-Vanasse J. 43.  et al. 2014. The Rad50 hook domain regulates DNA damage signaling and tumorigenesis. Genes Dev. 28:451–62 [Google Scholar]
  44. Alani E, Padmore R, Kleckner N. 44.  1990. Analysis of wild-type and rad50 mutants of yeast suggests an intimate relationship between meiotic chromosome synapsis and recombination. Cell 61:419–36 [Google Scholar]
  45. Bakkenist CJ, Kastan MB. 45.  2003. DNA damage activates ATM through intermolecular autophosphorylation and dimer dissociation. Nature 421:499–506 [Google Scholar]
  46. Matsuoka S, Ballif BA, Smogorzewska A, McDonald ER 3rd, Hurov KE. 46.  et al. 2007. ATM and ATR substrate analysis reveals extensive protein networks responsive to DNA damage. Science 316:1160–66 [Google Scholar]
  47. Bensimon A, Schmidt A, Ziv Y, Elkon R, Wang SY. 47.  et al. 2010. ATM-dependent and -independent dynamics of the nuclear phosphoproteome after DNA damage. Sci. Signal. 3:rs3 [Google Scholar]
  48. Shiloh Y, Ziv Y. 48.  2013. The ATM protein kinase: regulating the cellular response to genotoxic stress, and more. Nat. Rev. Mol. Cell Biol. 14:197–210 [Google Scholar]
  49. Berkovich E, Monnat RJ Jr, Kastan MB. 49.  2007. Roles of ATM and NBS1 in chromatin structure modulation and DNA double-strand break repair. Nat. Cell Biol. 9:683–90 [Google Scholar]
  50. Kozlov SV, Graham ME, Peng C, Chen P, Robinson PJ, Lavin MF. 50.  2006. Involvement of novel autophosphorylation sites in ATM activation. EMBO J. 25:3504–14 [Google Scholar]
  51. Pellegrini M, Celeste A, Difilippantonio S, Guo R, Wang W. 51.  et al. 2006. Autophosphorylation at serine 1987 is dispensable for murine Atm activation in vivo. Nature 443:222–25 [Google Scholar]
  52. Daniel JA, Pellegrini M, Lee JH, Paull TT, Feigenbaum L, Nussenzweig A. 52.  2008. Multiple autophosphorylation sites are dispensable for murine ATM activation in vivo. J. Cell Biol. 183:777–83 [Google Scholar]
  53. Lee JH, Paull TT. 53.  2005. ATM activation by DNA double-strand breaks through the Mre11–Rad50–Nbs1 complex. Science 308:551–54 [Google Scholar]
  54. Dupre A, Boyer-Chatenet L, Gautier J. 54.  2006. Two-step activation of ATM by DNA and the Mre11–Rad50–Nbs1 complex. Nat. Struct. Mol. Biol. 13:451–57 [Google Scholar]
  55. You Z, Chahwan C, Bailis J, Hunter T, Russell P. 55.  2005. ATM activation and its recruitment to damaged DNA require binding to the C terminus of Nbs1. Mol. Cell. Biol. 25:5363–79 [Google Scholar]
  56. So S, Davis AJ, Chen DJ. 56.  2009. Autophosphorylation at serine 1981 stabilizes ATM at DNA damage sites. J. Cell Biol. 187:977–90 [Google Scholar]
  57. Sun Y, Jiang X, Chen S, Fernandes N, Price BD. 57.  2005. A role for the Tip60 histone acetyltransferase in the acetylation and activation of ATM. PNAS 102:13182–87 [Google Scholar]
  58. Sun Y, Xu Y, Roy K, Price BD. 58.  2007. DNA damage-induced acetylation of lysine 3016 of ATM activates ATM kinase activity. Mol. Cell. Biol. 27:8502–9 [Google Scholar]
  59. Kaidi A, Jackson SP. 59.  2013. KAT5 tyrosine phosphorylation couples chromatin sensing to ATM signalling. Nature 498:70–74 [Google Scholar]
  60. Liu S, Shiotani B, Lahiri M, Marechal A, Tse A. 60.  et al. 2011. ATR autophosphorylation as a molecular switch for checkpoint activation. Mol. Cell 43:192–202 [Google Scholar]
  61. Douglas P, Sapkota GP, Morrice N, Yu Y, Goodarzi AA. 61.  et al. 2002. Identification of in vitro and in vivo phosphorylation sites in the catalytic subunit of the DNA-dependent protein kinase. Biochem. J. 368:243–51 [Google Scholar]
  62. Lee JH, Paull TT. 62.  2004. Direct activation of the ATM protein kinase by the Mre11/Rad50/Nbs1 complex. Science 304:93–96 [Google Scholar]
  63. Shiotani B, Zou L. 63.  2009. Single-stranded DNA orchestrates an ATM-to-ATR switch at DNA breaks. Mol. Cell 33:547–58 [Google Scholar]
  64. Smith GC, Cary RB, Lakin ND, Hann BC, Teo SH. 64.  et al. 1999. Purification and DNA binding properties of the ataxia-telangiectasia gene product ATM. PNAS 96:11134–39 [Google Scholar]
  65. Deshpande RA, Williams GJ, Limbo O, Williams RS, Kuhnlein J. 65.  et al. 2014. ATP-driven Rad50 conformations regulate DNA tethering, end resection, and ATM checkpoint signaling. EMBO J. 33:482–500 [Google Scholar]
  66. Cannon B, Kuhnlein J, Yang SH, Cheng A, Schindler D. 66.  et al. 2013. Visualization of local DNA unwinding by Mre11/Rad50/Nbs1 using single-molecule FRET. PNAS 110:18868–73 [Google Scholar]
  67. Costanzo V, Paull TT, Gottesman M, Gautier J. 67.  2004. Mre11 assembles linear DNA fragments into DNA damage signaling complexes. PLOS Biol. 2:e110 [Google Scholar]
  68. Lee JH, Mand MR, Deshpande RA, Kinoshita E, Yang SH. 68.  et al. 2013. Ataxia telangiectasia-mutated (ATM) kinase activity is regulated by ATP-driven conformational changes in the Mre11/Rad50/Nbs1 (MRN) complex. J. Biol. Chem. 288:12840–51 [Google Scholar]
  69. Arlett CF, Lehmann AR. 69.  1978. Human disorders showing increased sensitivity to the induction of genetic damage. Annu. Rev. Genet. 12:95–115 [Google Scholar]
  70. Andegeko Y, Moyal L, Mittelman L, Tsarfaty I, Shiloh Y, Rotman G. 70.  2001. Nuclear retention of ATM at sites of DNA double strand breaks. J. Biol. Chem. 276:38224–30 [Google Scholar]
  71. You Z, Bailis JM, Johnson SA, Dilworth SM, Hunter T. 71.  2007. Rapid activation of ATM on DNA flanking double-strand breaks. Nat. Cell Biol. 9:1311–18 [Google Scholar]
  72. Falck J, Coates J, Jackson SP. 72.  2005. Conserved modes of recruitment of ATM, ATR and DNA-PKcs to sites of DNA damage. Nature 434:605–11 [Google Scholar]
  73. Cerosaletti K, Wright J, Concannon P. 73.  2006. Active role for nibrin in the kinetics of atm activation. Mol. Cell. Biol. 26:1691–99 [Google Scholar]
  74. Paull TT, Gellert M. 74.  1998. The 3′ to 5′ exonuclease activity of Mre 11 facilitates repair of DNA double-strand breaks. Mol. Cell 1:969–79 [Google Scholar]
  75. Hopkins B, Paull TT. 75.  2008. The P. furiosus Mre11/Rad50 complex promotes 5′ strand resection at a DNA double-strand break. Cell 135:250–60 [Google Scholar]
  76. Trujillo KM, Yuan SS, Lee EY, Sung P. 76.  1998. Nuclease activities in a complex of human recombination and DNA repair factors Rad50, Mre11, and p95. J. Biol. Chem. 273:21447–50 [Google Scholar]
  77. Trujillo KM, Sung P. 77.  2001. DNA structure-specific nuclease activities in the Saccharomyces cerevisiae Rad50/Mre11 complex. J. Biol. Chem. 13:13 [Google Scholar]
  78. Buis J, Wu Y, Deng Y, Leddon J, Westfield G. 77a.  et al. 2008. Mre11 nuclease activity has essential roles in DNA repair and genomic stability distinct from ATM activation. Cell 135:85–96 [Google Scholar]
  79. Shibata A, Moiani D, Arvai AS, Perry J, Harding SM. 78.  et al. 2014. DNA double-strand break repair pathway choice is directed by distinct MRE11 nuclease activities. Mol. Cell 53:7–18 [Google Scholar]
  80. Jazayeri A, Balestrini A, Garner E, Haber JE, Costanzo V. 79.  2008. Mre11–Rad50–Nbs1 dependent processing of DNA breaks generates oligonucleotides that stimulate ATM activity. EMBO J. 27:1953–62 [Google Scholar]
  81. Usui T, Ogawa H, Petrini JH. 80.  2001. A DNA damage response pathway controlled by Tel1 and the Mre11 complex. Mol. Cell 7:1255–66 [Google Scholar]
  82. Limbo O, Porter-Goff ME, Rhind N, Russell P. 81.  2011. Mre11 nuclease activity and Ctp1 regulate Chk1 activation by Rad3ATR and Tel1ATM checkpoint kinases at double-strand breaks. Mol. Cell. Biol. 31:573–83 [Google Scholar]
  83. Morrow DM, Tagle DA, Shiloh Y, Collins FS, Hieter P. 82.  1995. TEL1, an S. cerevisiae homolog of the human gene mutated in ataxia telangiectasia, is functionally related to the yeast checkpoint gene MEC1. Cell 82:831–40 [Google Scholar]
  84. Hohl M, Kwon Y, Galvan SM, Xue X, Tous C. 83.  et al. 2011. The Rad50 coiled-coil domain is indispensable for Mre11 complex functions. Nat. Struct. Mol. Biol. 18:1124–31 [Google Scholar]
  85. Usui T, Petrini JH, Morales M. 84.  2006. Rad50S alleles of the Mre11 complex: questions answered and questions raised. Exp. Cell Res. 312:2694–99 [Google Scholar]
  86. Di Virgilio M, Ying CY, Gautier J. 85.  2009. PIKK-dependent phosphorylation of Mre11 induces MRN complex inactivation by disassembly from chromatin. DNA Repair 8:1311–20 [Google Scholar]
  87. Gatei M, Jakob B, Chen P, Kijas AW, Becherel OJ. 86.  et al. 2011. ATM protein–dependent phosphorylation of Rad50 protein regulates DNA repair and cell cycle control. J. Biol. Chem. 286:31542–56 [Google Scholar]
  88. Gatei M, Kijas AW, Biard D, Dork T, Lavin MF. 87.  2014. RAD50 phosphorylation promotes ATR downstream signaling and DNA restart following replication stress. Hum. Mol. Genet. 23:4232–48 [Google Scholar]
  89. Lee JH, Xu B, Lee CH, Ahn JY, Song MS. 88.  et al. 2003. Distinct functions of Nijmegen breakage syndrome in ataxia telangiectasia mutated–dependent responses to DNA damage. Mol. Cancer Res. 1:674–81 [Google Scholar]
  90. Gatei M, Young D, Cerosaletti KM, Desai-Mehta A, Spring K. 89.  et al. 2000. ATM-dependent phosphorylation of nibrin in response to radiation exposure. Nat. Genet. 25:115–19 [Google Scholar]
  91. Zhao S, Weng YC, Yuan SS, Lin YT, Hsu HC. 90.  et al. 2000. Functional link between ataxia-telangiectasia and Nijmegen breakage syndrome gene products. Nature 405:473–77 [Google Scholar]
  92. Lim DS, Kim ST, Xu B, Maser RS, Lin J. 91.  et al. 2000. ATM phosphorylates p95/Nbs1 in an S-phase checkpoint pathway. Nature 404:613–17 [Google Scholar]
  93. Li T, Wang ZQ. 92.  2011. Point mutation at the Nbs1 threonine 278 site does not affect mouse development, but compromises the Chk2 and Smc1 phosphorylation after DNA damage. Mech. Ageing Dev. 132:382–88 [Google Scholar]
  94. Buscemi G, Savio C, Zannini L, Micciche F, Masnada D. 93.  et al. 2001. Chk2 activation dependence on Nbs1 after DNA damage. Mol. Cell. Biol. 21:5214–22 [Google Scholar]
  95. Yazdi PT, Wang Y, Zhao S, Patel N, Lee EY, Qin J. 94.  2002. SMC1 is a downstream effector in the ATM/NBS1 branch of the human S-phase checkpoint. Genes Dev. 16:571–82 [Google Scholar]
  96. Bai Y, Murnane JP. 95.  2003. Telomere instability in a human tumor cell line expressing NBS1 with mutations at sites phosphorylated by ATM. Mol. Cancer Res. 1:1058–69 [Google Scholar]
  97. Lee AY, Liu E, Wu X. 96.  2007. The Mre11/Rad50/Nbs1 complex plays an important role in the prevention of DNA rereplication in mammalian cells. J. Biol. Chem. 282:32243–55 [Google Scholar]
  98. Wu J, Zhang X, Zhang L, Wu CY, Rezaeian AH. 97.  et al. 2012. Skp2 E3 ligase integrates ATM activation and homologous recombination repair by ubiquitinating NBS1. Mol. Cell 46:351–61 [Google Scholar]
  99. Li R, Yang YG, Gao Y, Wang ZQ, Tong WM. 98.  2012. A distinct response to endogenous DNA damage in the development of Nbs1-deficient cortical neurons. Cell Res. 22:859–72 [Google Scholar]
  100. Frappart PO, Tong WM, Demuth I, Radovanovic I, Herceg Z. 99.  et al. 2005. An essential function for NBS1 in the prevention of ataxia and cerebellar defects. Nat. Med. 11:538–44 [Google Scholar]
  101. Barlow C, Hirotsune S, Paylor R, Liyanage M, Eckhaus M. 100.  et al. 1996. Atm-deficient mice: a paradigm of ataxia telangiectasia. Cell 86:159–71 [Google Scholar]
  102. Xu Y, Ashley T, Brainerd EE, Bronson RT, Meyn MS, Baltimore D. 101.  1996. Targeted disruption of ATM leads to growth retardation, chromosomal fragmentation during meiosis, immune defects, and thymic lymphoma. Genes Dev. 10:2411–22 [Google Scholar]
  103. Barlow C, Ribaut-Barassin C, Zwingman TA, Pope AJ, Brown KD. 102.  et al. 2000. ATM is a cytoplasmic protein in mouse brain required to prevent lysosomal accumulation. PNAS 97:871–76 [Google Scholar]
  104. Elson A, Wang Y, Daugherty CJ, Morton CC, Zhou F. 103.  et al. 1996. Pleiotropic defects in ataxia-telangiectasia protein–deficient mice. PNAS 93:13084–89 [Google Scholar]
  105. Kuljis RO, Xu Y, Aguila MC, Baltimore D. 104.  1997. Degeneration of neurons, synapses, and neuropil and glial activation in a murine Atm knockout model of ataxia-telangiectasia. PNAS 94:12688–93 [Google Scholar]
  106. Barzilai A, Rotman G, Shiloh Y. 105.  2002. ATM deficiency and oxidative stress: a new dimension of defective response to DNA damage. DNA Repair 1:3–25 [Google Scholar]
  107. Rotman G, Shiloh Y. 106.  1997. Ataxia-telangiectasia: Is ATM a sensor of oxidative damage and stress?. Bioessays 19:911–17 [Google Scholar]
  108. Kamsler A, Daily D, Hochman A, Stern N, Shiloh Y. 107.  et al. 2001. Increased oxidative stress in ataxia telangiectasia evidenced by alterations in redox state of brains from Atm-deficient mice. Cancer Res. 61:1849–54 [Google Scholar]
  109. Quick KL, Dugan LL. 108.  2001. Superoxide stress identifies neurons at risk in a model of ataxia-telangiectasia. Ann. Neurol. 49:627–35 [Google Scholar]
  110. Yi M, Rosin MP, Anderson CK. 109.  1990. Response of fibroblast cultures from ataxia-telangiectasia patients to oxidative stress. Cancer Lett. 54:43–50 [Google Scholar]
  111. Ward AJ, Olive PL, Burr AH, Rosin MP. 110.  1994. Response of fibroblast cultures from ataxia-telangiectasia patients to reactive oxygen species generated during inflammatory reactions. Environ. Mol. Mutagen. 24:103–11 [Google Scholar]
  112. Shackelford RE, Innes CL, Sieber SO, Heinloth AN, Leadon SA, Paules RS. 111.  2001. The Ataxia telangiectasia gene product is required for oxidative stress–induced G1 and G2 checkpoint function in human fibroblasts. J. Biol. Chem. 276:21951–59 [Google Scholar]
  113. Clarke DD, Solokoff L. 112.  2006. Circulation and energy metabolism of the brain. Basic Neurochemistry: Molecular, Cellular, and Medical Aspects S Brady, G Siegel, RW Albers, D Price 637–70 Philadelphia: Lippincott-Raven [Google Scholar]
  114. Kurz EU, Douglas P, Lees-Miller SP. 113.  2004. Doxorubicin activates ATM-dependent phosphorylation of multiple downstream targets in part through the generation of reactive oxygen species. J. Biol. Chem. 279:53272–81 [Google Scholar]
  115. Ito K, Takubo K, Arai F, Satoh H, Matsuoka S. 114.  et al. 2007. Regulation of reactive oxygen species by Atm is essential for proper response to DNA double-strand breaks in lymphocytes. J. Immunol. 178:103–10 [Google Scholar]
  116. Ito K, Hirao A, Arai F, Matsuoka S, Takubo K. 115.  et al. 2004. Regulation of oxidative stress by ATM is required for self-renewal of haematopoietic stem cells. Nature 431:997–1002 [Google Scholar]
  117. Kim J, Wong PK. 116.  2009. Oxidative stress is linked to ERK1/2-p16 signaling–mediated growth defect in ATM-deficient astrocytes. J. Biol. Chem. 284:14396–404 [Google Scholar]
  118. Reliene R, Schiestl RH. 117.  2007. Antioxidants suppress lymphoma and increase longevity in Atm-deficient mice. J. Nutr. 137:S229–32 [Google Scholar]
  119. Guo Z, Kozlov S, Lavin MF, Person MD, Paull TT. 118.  2010. ATM activation by oxidative stress. Science 330:517–21 [Google Scholar]
  120. Bencokova Z, Kaufmann MR, Pires IM, Lecane PS, Giaccia AJ, Hammond EM. 119.  2009. ATM activation and signaling under hypoxic conditions. Mol. Cell. Biol. 29:526–37 [Google Scholar]
  121. Chandel NS, McClintock DS, Feliciano CE, Wood TM, Melendez JA. 120.  et al. 2000. Reactive oxygen species generated at mitochondrial complex III stabilize hypoxia-inducible factor 1α during hypoxia: a mechanism of O2 sensing. J. Biol. Chem. 275:25130–38 [Google Scholar]
  122. Hunt CR, Pandita RK, Laszlo A, Higashikubo R, Agarwal M. 121.  et al. 2007. Hyperthermia activates a subset of ataxia-telangiectasia mutated effectors independent of DNA strand breaks and heat shock protein 70 status. Cancer Res. 67:3010–17 [Google Scholar]
  123. Chessa L, Petrinelli P, Antonelli A, Fiorilli M, Elli R. 122.  et al. 1992. Heterogeneity in ataxia-telangiectasia: classical phenotype associated with intermediate cellular radiosensitivity. Am. J. Med. Genet. 42:741–46 [Google Scholar]
  124. Gilad S, Chessa L, Khosravi R, Russell P, Galanty Y. 123.  et al. 1998. Genotype–phenotype relationships in ataxia-telangiectasia and variants. Am. J. Hum. Genet. 62:551–61 [Google Scholar]
  125. Toyoshima M, Hara T, Zhang H, Yamamoto T, Akaboshi S. 124.  et al. 1998. Ataxia-telangiectasia without immunodeficiency: novel point mutations within and adjacent to the phosphatidylinositol 3-kinase–like domain. Am. J. Med. Genet. 75:141–14 [Google Scholar]
  126. Guo Z, Deshpande R, Paull TT. 125.  2010. ATM activation in the presence of oxidative stress. Cell Cycle 9:4805–11 [Google Scholar]
  127. Yang DQ, Halaby MJ, Li Y, Hibma JC, Burn P. 126.  2011. Cytoplasmic ATM protein kinase: an emerging therapeutic target for diabetes, cancer and neuronal degeneration. Drug Discov. Today 16:332–38 [Google Scholar]
  128. Valentin-Vega YA, Maclean KH, Tait-Mulder J, Milasta S, Steeves M. 127.  et al. 2011. Mitochondrial dysfunction in ataxia telangiectasia. Blood 119:1490–500 [Google Scholar]
  129. Watters D, Kedar P, Spring K, Bjorkman J, Chen P. 128.  et al. 1999. Localization of a portion of extranuclear ATM to peroxisomes. J. Biol. Chem. 274:34277–82 [Google Scholar]
  130. Watters D, Khanna KK, Beamish H, Birrell G, Spring K. 129.  et al. 1997. Cellular localisation of the ataxia-telangiectasia (ATM) gene product and discrimination between mutated and normal forms. Oncogene 14:1911–21 [Google Scholar]
  131. Sibanda BL, Chirgadze DY, Blundell TL. 130.  2010. Crystal structure of DNA-PKcs reveals a large open-ring cradle comprised of HEAT repeats. Nature 463:118–21 [Google Scholar]
  132. Williams GJ, Hammel M, Radhakrishnan SK, Ramsden D, Lees-Miller SP, Tainer JA. 131.  2014. Structural insights into NHEJ: building up an integrated picture of the dynamic DSB repair super complex, one component and interaction at a time. DNA Repair 17:110–20 [Google Scholar]
  133. Yang H, Rudge DG, Koos JD, Vaidialingam B, Yang HJ, Pavletich NP. 132.  2013. mTOR kinase structure, mechanism and regulation. Nature 497:217–23 [Google Scholar]
  134. Mordes DA, Cortez D. 133.  2008. Activation of ATR and related PIKKs. Cell Cycle 7:2809–12 [Google Scholar]
  135. Lawrence JC, Lin TA, McMahon LP, Choi KM. 134.  2004. Modulation of the protein kinase activity of mTOR. Curr. Top. Microbiol. Immunol. 279:199–213 [Google Scholar]
  136. Schroeder EA, Raimundo N, Shadel GS. 135.  2013. Epigenetic silencing mediates mitochondria stress-induced longevity. Cell Metab. 17:954–64 [Google Scholar]
  137. Laplante M, Sabatini DM. 136.  2012. mTOR signaling in growth control and disease. Cell 149:274–93 [Google Scholar]
  138. Weber JD, Gutmann DH. 137.  2012. Deconvoluting mTOR biology. Cell Cycle 11:236–48 [Google Scholar]
  139. Lakin ND, Weber P, Stankovic T, Rottinghaus ST, Taylor AM, Jackson SP. 138.  1996. Analysis of the ATM protein in wild-type and ataxia telangiectasia cells. Oncogene 13:2707–16 [Google Scholar]
  140. Ohne Y, Takahara T, Hatakeyama R, Matsuzaki T, Noda M. 139.  et al. 2008. Isolation of hyperactive mutants of mammalian target of rapamycin. J. Biol. Chem. 283:31861–70 [Google Scholar]
  141. Urano J, Sato T, Matsuo T, Otsubo Y, Yamamoto M, Tamanoi F. 140.  2007. Point mutations in TOR confer Rheb-independent growth in fission yeast and nutrient-independent mammalian TOR signaling in mammalian cells. PNAS 104:3514–19 [Google Scholar]
  142. Baldo V, Testoni V, Lucchini G, Longhese MP. 141.  2008. Dominant TEL1-hy mutations compensate for Mec1 lack of functions in the DNA damage response. Mol. Cell. Biol. 28:358–75 [Google Scholar]
  143. Goodarzi AA, Jonnalagadda JC, Douglas P, Young D, Ye R. 142.  et al. 2004. Autophosphorylation of ataxia-telangiectasia mutated is regulated by protein phosphatase 2A. EMBO J. 23:4451–61 [Google Scholar]
  144. Peng A, Lewellyn AL, Schiemann WP, Maller JL. 143.  2010. Repo-Man controls a protein phosphatase 1–dependent threshold for DNA damage checkpoint activation. Curr. Biol. 20:387–96 [Google Scholar]
  145. Fiscella M, Zhang H, Fan S, Sakaguchi K, Shen S. 144.  et al. 1997. Wip1, a novel human protein phosphatase that is induced in response to ionizing radiation in a p53-dependent manner. PNAS 94:6048–53 [Google Scholar]
  146. Shreeram S, Demidov ON, Hee WK, Yamaguchi H, Onishi N. 145.  et al. 2006. Wip1 phosphatase modulates ATM-dependent signaling pathways. Mol. Cell 23:757–64 [Google Scholar]
  147. Bulavin DV, Phillips C, Nannenga B, Timofeev O, Donehower LA. 146.  et al. 2004. Inactivation of the Wip1 phosphatase inhibits mammary tumorigenesis through p38 MAPK–mediated activation of the p16Ink4a–p1Arf pathway. Nat. Genet. 36:343–50 [Google Scholar]
  148. Batchelor E, Mock CS, Bhan I, Loewer A, Lahav G. 147.  2008. Recurrent initiation: a mechanism for triggering p53 pulses in response to DNA damage. Mol. Cell 30:277–89 [Google Scholar]
  149. Ali A, Zhang J, Bao S, Liu I, Otterness D. 148.  et al. 2004. Requirement of protein phosphatase 5 in DNA-damage-induced ATM activation. Genes Dev. 18:249–54 [Google Scholar]
  150. Yong W, Bao S, Chen H, Li D, Sanchez ER, Shou W. 149.  2007. Mice lacking protein phosphatase 5 are defective in ataxia telangiectasia mutated (ATM)-mediated cell cycle arrest. J. Biol. Chem. 282:14690–94 [Google Scholar]
  151. Lou Z, Minter-Dykhouse K, Franco S, Gostissa M, Rivera MA. 150.  et al. 2006. MDC1 maintains genomic stability by participating in the amplification of ATM-dependent DNA damage signals. Mol. Cell 21:187–200 [Google Scholar]
  152. Stucki M, Clapperton JA, Mohammad D, Yaffe MB, Smerdon SJ, Jackson SP. 151.  2005. MDC1 directly binds phosphorylated histone H2AX to regulate cellular responses to DNA double-strand breaks. Cell 123:1213–26 [Google Scholar]
  153. Chapman JR, Jackson SP. 152.  2008. Phospho-dependent interactions between NBS1 and MDC1 mediate chromatin retention of the MRN complex at sites of DNA damage. EMBO Rep. 9:795–801 [Google Scholar]
  154. Melander F, Bekker-Jensen S, Falck J, Bartek J, Mailand N, Lukas J. 153.  2008. Phosphorylation of SDT repeats in the MDC1 N terminus triggers retention of NBS1 at the DNA damage–modified chromatin. J. Cell Biol. 181:213–26 [Google Scholar]
  155. Spycher C, Miller ES, Townsend K, Pavic L, Morrice NA. 154.  et al. 2008. Constitutive phosphorylation of MDC1 physically links the MRE11–RAD50–NBS1 complex to damaged chromatin. J. Cell Biol. 181:227–40 [Google Scholar]
  156. Wu L, Luo K, Lou Z, Chen J. 155.  2008. MDC1 regulates intra-S-phase checkpoint by targeting NBS1 to DNA double-strand breaks. PNAS 105:11200–5 [Google Scholar]
  157. Peng A, Lewellyn AL, Maller JL. 156.  2007. Undamaged DNA transmits and enhances DNA damage checkpoint signals in early embryos. Mol. Cell. Biol. 27:6852–62 [Google Scholar]
  158. Wang Q, Alexander P, Goldstein M, Wakeman TP, Sun T. 157.  et al. 2014. Rad17 recruits the MRE11–RAD50–NBS1 complex to regulate the cellular response to DNA double-strand breaks. EMBO J. 33:862–77 [Google Scholar]
  159. Ayrapetov MK, Gursoy-Yuzugullu O, Xu C, Xu Y, Price BD. 158.  2014. DNA double-strand breaks promote methylation of histone H3 on lysine 9 and transient formation of repressive chromatin. PNAS 111:9169–74 [Google Scholar]
  160. Carvalho S, Vitor AC, Sridhara SC, Martins FB, Raposo AC. 159.  et al. 2014. SETD2 is required for DNA double-strand break repair and activation of the p53-mediated checkpoint. eLife 3:e02482 [Google Scholar]
  161. Zhang L, Chen H, Gong M, Gong F. 160.  2013. The chromatin remodeling protein BRG1 modulates BRCA1 response to UV irradiation by regulating ATR/ATM activation. Front. Oncol. 3:7 [Google Scholar]
  162. Ray A, Mir SN, Wani G, Zhao Q, Battu A. 161.  et al. 2009. Human SNF5/INI1, a component of the human SWI/SNF chromatin remodeling complex, promotes nucleotide excision repair by influencing ATM recruitment and downstream H2AX phosphorylation. Mol. Cell. Biol. 29:6206–19 [Google Scholar]
  163. Liang B, Qiu J, Ratnakumar K, Laurent BC. 162.  2007. RSC functions as an early double-strand-break sensor in the cell's response to DNA damage. Curr. Biol. 17:1432–37 [Google Scholar]
  164. Park JH, Park EJ, Lee HS, Kim SJ, Hur SK. 163.  et al. 2006. Mammalian SWI/SNF complexes facilitate DNA double-strand break repair by promoting γ-H2AX induction. EMBO J. 25:3986–97 [Google Scholar]
  165. Gupta A, Sharma GG, Young CS, Agarwal M, Smith ER. 164.  et al. 2005. Involvement of human MOF in ATM function. Mol. Cell. Biol. 25:5292–305 [Google Scholar]
  166. Kim YC, Gerlitz G, Furusawa T, Catez F, Nussenzweig A. 165.  et al. 2009. Activation of ATM depends on chromatin interactions occurring before induction of DNA damage. Nat. Cell Biol. 11:92–96 [Google Scholar]
  167. Soutoglou E, Misteli T. 166.  2008. Activation of the cellular DNA damage response in the absence of DNA lesions. Science 320:1507–10 [Google Scholar]
  168. White JS, Choi S, Bakkenist CJ. 167.  2010. Transient ATM kinase inhibition disrupts DNA damage–induced sister chromatid exchange. Sci. Signal. 3:ra44 [Google Scholar]
  169. Daniel JA, Pellegrini M, Lee BS, Guo Z, Filsuf D. 168.  et al. 2012. Loss of ATM kinase activity leads to embryonic lethality in mice. J. Cell Biol. 198:295–304 [Google Scholar]
  170. Yamamoto K, Wang Y, Jiang W, Liu X, Dubois RL. 169.  et al. 2012. Kinase-dead ATM protein causes genomic instability and early embryonic lethality in mice. J. Cell Biol. 198:305–13 [Google Scholar]
  171. Kanu N, Behrens A. 170.  2007. ATMIN defines an NBS1-independent pathway of ATM signalling. EMBO J. 26:2933–41 [Google Scholar]
  172. Zhang T, Penicud K, Bruhn C, Loizou JI, Kanu N. 171.  et al. 2012. Competition between NBS1 and ATMIN controls ATM signaling pathway choice. Cell Rep. 2:1498–504 [Google Scholar]
  173. Zhang T, Cronshaw J, Kanu N, Snijders AP, Behrens A. 172.  2014. UBR5-mediated ubiquitination of ATMIN is required for ionizing radiation–induced ATM signaling and function. PNAS 111:12091–96 [Google Scholar]
  174. Lee JH, Goodarzi AA, Jeggo PA, Paull TT. 173.  2010. 53BP1 promotes ATM activity through direct interactions with the MRN complex. EMBO J. 29:574–85 [Google Scholar]
  175. Bowen C, Ju JH, Lee JH, Paull TT, Gelmann EP. 174.  2013. Functional activation of ATM by the prostate cancer suppressor NKX3.1. Cell Rep. 4:516–29 [Google Scholar]
  176. Chiba N, Comaills V, Shiotani B, Takahashi F, Shimada T. 175.  et al. 2012. Homeobox B9 induces epithelial-to-mesenchymal transition–associated radioresistance by accelerating DNA damage responses. PNAS 109:2760–65 [Google Scholar]
  177. Wang M, Saha J, Hada M, Anderson JA, Pluth JM. 176.  et al. 2013. Novel Smad proteins localize to IR-induced double-strand breaks: interplay between TGFβ and ATM pathways. Nucleic Acids Res. 41:933–42 [Google Scholar]
  178. Park S, Kang JM, Kim SJ, Kim H, Hong S. 177.  et al. 2014. Smad7 enhances ATM activity by facilitating the interaction between ATM and Mre11–Rad50–Nbs1 complex in DNA double-strand break repair. Cell. Mol. Life Sci. 72583–96
  179. Guo JY, Yamada A, Kajino T, Wu JQ, Tang W. 178.  et al. 2008. Aven-dependent activation of ATM following DNA damage. Curr. Biol. 18:933–42 [Google Scholar]
  180. Hong ST, Choi KW. 179.  2013. TCTP directly regulates ATM activity to control genome stability and organ development in Drosophila melanogaster. Nat. Commun. 4:2986 [Google Scholar]
  181. Zhang J, de Toledo SM, Pandey BN, Guo G, Pain D. 180.  et al. 2012. Role of the translationally controlled tumor protein in DNA damage sensing and repair. PNAS 109:e926–33 [Google Scholar]
  182. Lim HC, Xie L, Zhang W, Li R, Chen ZC. 181.  et al. 2013. Ribosomal S6 kinase 2 (RSK2) maintains genomic stability by activating the Atm/p53-dependent DNA damage pathway. PLOS ONE 8:e74334 [Google Scholar]
  183. Chen C, Zhang L, Huang NJ, Huang B, Kornbluth S. 182.  2013. Suppression of DNA-damage checkpoint signaling by Rsk-mediated phosphorylation of Mre11. PNAS 110:20605–10 [Google Scholar]
  184. Heiss EH, Schilder YD, Dirsch VM. 183.  2007. Chronic treatment with resveratrol induces redox stress- and ataxia telangiectasia–mutated (ATM)-dependent senescence in p53-positive cancer cells. J. Biol. Chem. 282:26759–66 [Google Scholar]
  185. Tyagi A, Singh RP, Agarwal C, Siriwardana S, Sclafani RA, Agarwal R. 184.  2005. Resveratrol causes Cdc2–Tyr15 phosphorylation via ATM/ATR–Chk1/2–Cdc25C pathway as a central mechanism for S phase arrest in human ovarian carcinoma Ovcar-3 cells. Carcinogenesis 26:1978–87 [Google Scholar]
  186. Gatz SA, Keimling M, Baumann C, Dork T, Debatin K-M. 185.  et al. 2008. Resveratrol modulates DNA double-strand break repair pathways in an ATM/ATR-p53- and -Nbs1-dependent manner. Carcinogenesis 29:519–27 [Google Scholar]
  187. Lee JH, Guo Z, Myler LR, Zheng S, Paull TT. 186.  2014. Direct activation of ATM by resveratrol under oxidizing conditions. PLOS ONE 9:e97969 [Google Scholar]
  188. Pirola L, Fröjdö S. 187.  2008. Resveratrol: one molecule, many targets. IUBMB Life 60:323–32 [Google Scholar]
  189. Bosotti R, Isacchi A, Sonnhammer EL. 188.  2000. FAT: a novel domain in PIK-related kinases. Trends Biochem. Sci. 25:225–27 [Google Scholar]
  190. Keith CT, Schreiber SL. 189.  1995. PIK-related kinases: DNA repair, recombination, and cell cycle checkpoints. Science 270:50–51 [Google Scholar]
  191. Efeyan A, Zoncu R, Sabatini DM. 190.  2012. Amino acids and mTORC1: from lysosomes to disease. Trends Mol. Med. 18:524–33 [Google Scholar]
  192. Hammel M, Yu Y, Mahaney BL, Cai B, Ye R. 191.  et al. 2010. Ku and DNA-dependent protein kinase dynamic conformations and assembly regulate DNA binding and the initial non-homologous end joining complex. J. Biol. Chem. 285:1414–23 [Google Scholar]
/content/journals/10.1146/annurev-biochem-060614-034335
Loading
/content/journals/10.1146/annurev-biochem-060614-034335
Loading

Data & Media loading...

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