1932

Abstract

Genomic instability in disease and its fidelity in health depend on the DNA damage response (DDR), regulated in part from the complex of meiotic recombination 11 homolog 1 (MRE11), ATP-binding cassette–ATPase (RAD50), and phosphopeptide-binding Nijmegen breakage syndrome protein 1 (NBS1). The MRE11–RAD50–NBS1 (MRN) complex forms a multifunctional DDR machine. Within its network assemblies, MRN is the core conductor for the initial and sustained responses to DNA double-strand breaks, stalled replication forks, dysfunctional telomeres, and viral DNA infection. MRN can interfere with cancer therapy and is an attractive target for precision medicine. Its conformations change the paradigm whereby kinases initiate damage sensing. Delineated results reveal kinase activation, posttranslational targeting, functional scaffolding, conformations storing binding energy and enabling access, interactions with hub proteins such as replication protein A (RPA), and distinct networks at DNA breaks and forks. MRN biochemistry provides prototypic insights into how it initiates, implements, and regulates multifunctional responses to genomic stress.

Loading

Article metrics loading...

/content/journals/10.1146/annurev-biochem-062917-012415
2018-06-20
2024-04-19
Loading full text...

Full text loading...

/deliver/fulltext/biochem/87/1/annurev-biochem-062917-012415.html?itemId=/content/journals/10.1146/annurev-biochem-062917-012415&mimeType=html&fmt=ahah

Literature Cited

  1. 1.  Ciccia A, Elledge SJ 2010. The DNA damage response: making it safe to play with knives. Mol. Cell 40:179–204
    [Google Scholar]
  2. 2.  D'Amours D, Jackson SP 2002. The MRE11 complex: at the crossroads of DNA repair and checkpoint signalling. Nat. Rev. Mol. Cell Biol. 3:317–27
    [Google Scholar]
  3. 3.  Connelly JC, Leach DRF 2002. Tethering on the brink: the evolutionarily conserved Mre11–Rad50 complex. Trends Biochem. Sci. 27:410–18
    [Google Scholar]
  4. 4.  Stracker TH, Petrini JHJ 2011. The MRE11 complex: starting from the ends. Nat. Rev. Mol. Cell Biol. 12:90–103
    [Google Scholar]
  5. 5.  Uchisaka N, Takahashi N, Sato M, Kikuchi A, Mochizuki S et al. 2009. Two brothers with ataxia-telangiectasia-like disorder with lung adenocarcinoma. J. Pediatr. 155:435–38
    [Google Scholar]
  6. 6.  Matsumoto Y, Miyamoto T, Sakamoto H, Izumi H, Nakazawa Y et al. 2011. Two unrelated patients with MRE11A mutations and Nijmegen breakage syndrome-like severe microcephaly. DNA Repair 10:314–21
    [Google Scholar]
  7. 7.  Miyamoto R, Morino H, Yoshizawa A, Miyazaki Y, Maruyama H et al. 2014. Exome sequencing reveals a novel MRE11 mutation in a patient with progressive myoclonic ataxia. J. Neurol. Sci. 337:219–23
    [Google Scholar]
  8. 8.  Stewart GS, Maser RS, Stankovic T, Bressan DA, Kaplan MI 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]
  9. 9.  Delia D. 2004. MRE11 mutations and impaired ATM-dependent responses in an Italian family with ataxia-telangiectasia-like disorder. Hum. Mol. Genet. 13:2155–63
    [Google Scholar]
  10. 10.  Fernet M, Gribaa M, Salih MAM, Seidahmed MZ, Hall J, Koenig M 2005. Identification and functional consequences of a novel MRE11 mutation affecting 10 Saudi Arabian patients with the ataxia telangiectasia-like disorder. Hum. Mol. Genet. 14:307–18
    [Google Scholar]
  11. 11.  Brandt S, Samartzis EP, Zimmermann A-K, Fink D, Moch H et al. 2017. Lack of MRE11-RAD50-NBS1 (MRN) complex detection occurs frequently in low-grade epithelial ovarian cancer. BMC Cancer 17:44
    [Google Scholar]
  12. 12.  Gravells P, Grant E, Smith KM, James DI, Bryant HE 2017. Specific killing of DNA damage-response deficient cells with inhibitors of poly(ADP-ribose) glycohydrolase. DNA Repair 52:81–91
    [Google Scholar]
  13. 13.  Kim I-K, Kiefer JR, Ho CMW, Stegeman RA, Classen S et al. 2012. Structure of mammalian poly(ADP-ribose) glycohydrolase reveals a flexible tyrosine clasp as a substrate-binding element. Nat. Struct. Mol. Biol. 19:653–56
    [Google Scholar]
  14. 14.  Wang Y, Gudikote J, Giri U, Yan J, Deng W et al. 2018. RAD50 expression is associated with poor clinical outcomes after radiotherapy for resected non–small cell lung cancer. Clin. Cancer Res. 24:341–50
    [Google Scholar]
  15. 15.  Lamarche BJ, Orazio NI, Weitzman MD 2010. The MRN complex in double-strand break repair and telomere maintenance. FEBS Lett 584:3682–95
    [Google Scholar]
  16. 16.  Williams GJ, Lees-Miller SP, Tainer JA 2010. Mre11–Rad50–Nbs1 conformations and the control of sensing, signaling, and effector responses at DNA double-strand breaks. DNA Repair 9:1299–306
    [Google Scholar]
  17. 17.  Lavin M, Kozlov S, Gatei M, Kijas A 2015. ATM-dependent phosphorylation of all three members of the MRN complex: from sensor to adaptor. Biomolecules 5:2877–902
    [Google Scholar]
  18. 18.  Matsuoka S, Ballif BA, Smogorzewska A, McDonald ER, Hurov KE et al. 2007. ATM and ATR substrate analysis reveals extensive protein networks responsive to DNA damage. Science 316:1160–66
    [Google Scholar]
  19. 19.  Williams RS, Dodson GE, Limbo O, Yamada Y, Williams JS et al. 2009. Nbs1 flexibly tethers Ctp1 and Mre11-Rad50 to coordinate DNA double-strand break processing and repair. Cell 139:87–99
    [Google Scholar]
  20. 20.  Lloyd J, Chapman JR, Clapperton JA, Haire LF, Hartsuiker E et al. 2009. A supramodular FHA/BRCT-repeat architecture mediates Nbs1 adaptor function in response to DNA damage. Cell 139:100–11
    [Google Scholar]
  21. 21.  Liu Y, Sung S, Kim Y, Li F, Gwon G et al. 2016. ATP‐dependent DNA binding, unwinding, and resection by the Mre11/Rad50 complex. EMBO J 35:743–58
    [Google Scholar]
  22. 22.  Sung S, Li F, Park YB, Kim JS, Kim AK et al. 2014. DNA end recognition by the Mre11 nuclease dimer: insights into resection and repair of damaged DNA. EMBO J 33:2422–35
    [Google Scholar]
  23. 23.  Schiller CB, Lammens K, Guerini I, Coordes B, Feldmann H et al. 2012. Structure of Mre11–Nbs1 complex yields insights into ataxia-telangiectasia–like disease mutations and DNA damage signaling. Nat. Struct. Mol. Biol. 19:693–700
    [Google Scholar]
  24. 24.  Hammel M, Yu Y, Radhakrishnan SK, Chokshi C, Tsai M-S et al. 2016. An intrinsically disordered APLF links Ku, DNA-PKcs, and XRCC4-DNA ligase IV in an extended flexible non-homologous end joining complex. J. Biol. Chem. 291:26987–7006
    [Google Scholar]
  25. 25.  Lavin MF. 2007. ATM and the Mre11 complex combine to recognize and signal DNA double-strand breaks. Oncogene 26:7749–58
    [Google Scholar]
  26. 26.  Marechal A, Zou L 2013. DNA damage sensing by the ATM and ATR kinases. Cold Spring Harb. Perspect. Biol. 5:a012716
    [Google Scholar]
  27. 27.  Shibata A, Moiani D, Arvai AS, Perry J, Harding SM et al. 2014. DNA double-strand break repair pathway choice is directed by distinct MRE11 nuclease activities. Mol. Cell 53:7–18
    [Google Scholar]
  28. 28.  Trenz K, Smith E, Smith S, Costanzo V 2006. ATM and ATR promote Mre11 dependent restart of collapsed replication forks and prevent accumulation of DNA breaks. EMBO J 25:1764–74
    [Google Scholar]
  29. 29.  Lafrance-Vanasse J, Williams GJ, Tainer JA 2015. Envisioning the dynamics and flexibility of Mre11-Rad50-Nbs1 complex to decipher its roles in DNA replication and repair. Prog. Biophys. Mol. Biol. 117:182–93
    [Google Scholar]
  30. 30.  Hopfner K-P, Karcher A, Shin DS, Craig L, Arthur LM et al. 2000. Structural biology of Rad50 ATPase. Cell 101:789–800
    [Google Scholar]
  31. 31.  Cahill D, Carney JP 2007. Dimerization of the Rad50 protein is independent of the conserved hook domain. Mutagenesis 22:269–74
    [Google Scholar]
  32. 32.  Hopfner K-P, Craig L, Moncalian G, Zinkel RA, Usui T et al. 2002. The Rad50 zinc-hook is a structure joining Mre11 complexes in DNA recombination and repair. Nature 418:562–66
    [Google Scholar]
  33. 33.  Park YB, Hohl M, Padjasek M, Jeong E, Jin KS et al. 2017. Eukaryotic Rad50 functions as a rod-shaped dimer. Nat. Struct. Mol. Biol. 24:248–57
    [Google Scholar]
  34. 34.  Myler LR, Gallardo IF, Soniat MM, Deshpande RA, Gonzalez XB et al. 2017. Single-molecule imaging reveals how Mre11-Rad50-Nbs1 initiates DNA break repair.. Mol. Cell 67:891–98.e4
    [Google Scholar]
  35. 35.  Williams GJ, Williams RS, Williams JS, Moncalian G, Arvai AS et al. 2011. ABC ATPase signature helices in Rad50 link nucleotide state to Mre11 interface for DNA repair. Nat. Struct. Mol. Biol. 18:423–41
    [Google Scholar]
  36. 36.  Williams RS, Moncalian G, Williams JS, Yamada Y, Limbo O et al. 2008. Mre11 dimers coordinate DNA end bridging and nuclease processing in double-strand-break repair. Cell 135:97–109
    [Google Scholar]
  37. 37.  Das D, Moiani D, Axelrod HL, Miller MD, McMullan D et al. 2010. Crystal structure of the first eubacterial Mre11 nuclease reveals novel features that may discriminate substrates during DNA repair. J. Mol. Biol. 397:647–63
    [Google Scholar]
  38. 38.  Seifert FU, Lammens K, Stoehr G, Kessler B, Hopfner KP 2016. Structural mechanism of ATP‐dependent DNA binding and DNA end bridging by eukaryotic Rad50. EMBO J 35:759–72
    [Google Scholar]
  39. 39.  Seifert FU, Lammens K, Hopfner K-P 2015. Structure of the catalytic domain of Mre11 from Chaetomium thermophilum. Acta Crystallogr. Sect. F 71:752–57
    [Google Scholar]
  40. 40.  Hopfner KP, Karcher A, Craig L, Woo TT, Carney JP, Tainer JA 2001. Structural biochemistry and interaction architecture of the DNA double-strand break repair Mre11 nuclease and Rad50-ATPase. Cell 105:473–85
    [Google Scholar]
  41. 41.  Lee JH. 2004. Direct activation of the ATM protein kinase by the Mre11/Rad50/Nbs1 complex. Science 304:93–96
    [Google Scholar]
  42. 42.  Falck J, Coates J, Jackson SP 2005. Conserved modes of recruitment of ATM, ATR and DNA-PKcs to sites of DNA damage. Nature 434:605–11
    [Google Scholar]
  43. 43.  You Z, Chahwan C, Bailis J, Hunter T, Russell P 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]
  44. 44.  Sibanda BL, Chirgadze DY, Ascher DB, Blundell TL 2017. DNA-PKcs structure suggests an allosteric mechanism modulating DNA double-strand break repair. Science 355:520–24
    [Google Scholar]
  45. 45.  Stracker TH, Morales M, Couto SS, Hussein H, Petrini JHJ 2007. The carboxy terminus of NBS1 is required for induction of apoptosis by the MRE11 complex. Nature 447:218–21
    [Google Scholar]
  46. 46.  Difilippantonio S, Celeste A, Kruhlak MJ, Lee Y, Difilippantonio MJ et al. 2007. Distinct domains in Nbs1 regulate irradiation-induced checkpoints and apoptosis. J. Exp. Med. 204:1003–11
    [Google Scholar]
  47. 47.  Moiani D, Ronato DA, Brosey CA, Arvai AS, Syed A et al. 2017. Targeting allostery with avatars to design inhibitors assessed by cell activity: dissecting MRE11 endo- and exonuclease activities. Methods Enzymol 601:205–41
    [Google Scholar]
  48. 48.  Davies OR, Forment JV, Sun M, Belotserkovskaya R, Coates J et al. 2015. CtIP tetramer assembly is required for DNA-end resection and repair. Nat. Struct. Mol. Biol. 22:150–57
    [Google Scholar]
  49. 49.  Andres SN, Appel CD, Westmoreland JW, Williams JS, Nguyen Y et al. 2015. Tetrameric Ctp1 coordinates DNA binding and DNA bridging in DNA double-strand-break repair. Nat. Struct. Mol. Biol. 22:158–66
    [Google Scholar]
  50. 50.  Andres SN, Williams RS 2017. CtIP/Ctp1/Sae2, molecular form fit for function. DNA Repair 56:109–17
    [Google Scholar]
  51. 51.  Lammens K, Bemeleit DJ, Möckel C, Clausing E, Schele A et al. 2011. The Mre11:Rad50 structure shows an ATP-dependent molecular clamp in DNA double-strand break repair. Cell 145:54–66
    [Google Scholar]
  52. 52.  Deshpande RA, Williams GJ, Limbo O, Williams RS, Kuhnlein J et al. 2014. ATP-driven Rad50 conformations regulate DNA tethering, end resection, and ATM checkpoint signaling. EMBO J 33:482–500
    [Google Scholar]
  53. 53.  Rojowska A, Lammens K, Seifert FU, Direnberger C, Feldmann H, Hopfner KP 2014. Structure of the Rad50 DNA double‐strand break repair protein in complex with DNA. EMBO J 33:2847–59
    [Google Scholar]
  54. 54.  He J, Shi LZ, Truong LN, Lu C-S, Razavian N et al. 2012. Rad50 zinc hook is important for the Mre11 complex to bind chromosomal DNA double-stranded breaks and initiate various DNA damage responses. J. Biol. Chem. 287:31747–56
    [Google Scholar]
  55. 55.  Barfoot T, Herdendorf TJ, Behning BR, Stohr BA, Gao Y et al. 2015. Functional analysis of the bacteriophage T4 Rad50 homolog (gp46) coiled-coil domain. J. Biol. Chem. 290:23905–15
    [Google Scholar]
  56. 56.  Roset R, Inagaki A, Hohl M, Brenet F, Lafrance-Vanasse J et al. 2014. The Rad50 hook domain regulates DNA damage signaling and tumorigenesis. Genes Dev 28:451–62
    [Google Scholar]
  57. 57.  Hohl M, Kwon Y, Galván SM, Xue X, Tous C et al. 2011. The Rad50 coiled-coil domain is indispensable for Mre11 complex functions. Nat. Struct. Mol. Biol. 18:1124–31
    [Google Scholar]
  58. 58.  Reindl S, Ghosh A, Williams GJ, Lassak K, Neiner T et al. 2013. Insights into FlaI functions in archaeal motor assembly and motility from structures, conformations, and genetics. Mol. Cell 49:1069–82
    [Google Scholar]
  59. 59.  Acharya SN, Many AM, Schroeder AP, Kennedy FM, Savytskyy OP et al. 2008. Coprinus cinereus Rad50 mutants reveal an essential structural role for Rad50 in axial element and synaptonemal complex formation, homolog pairing and meiotic recombination. Genetics 180:1889–907
    [Google Scholar]
  60. 60.  Al-Ahmadie H, Iyer G, Hohl M, Asthana S, Inagaki A et al. 2014. Synthetic lethality in ATM-deficient RAD50-mutant tumors underlies outlier response to cancer therapy. Cancer Discov 4:1014–21
    [Google Scholar]
  61. 61.  Gao Y, Nelson SW 2014. Autoinhibition of bacteriophage T4 Mre11 by its C-terminal domain. J. Biol. Chem. 289:26505–13
    [Google Scholar]
  62. 62.  Tsutakawa SE, Lafrance-Vanasse J, Tainer JA 2014. The cutting edges in DNA repair, licensing, and fidelity: DNA and RNA repair nucleases sculpt DNA to measure twice, cut once. DNA Repair 19:95–107
    [Google Scholar]
  63. 63.  Kim JH, Grosbart M, Anand R, Wyman C, Cejka P, Petrini JHJ 2017. The Mre11-Nbs1 interface is essential for viability and tumor suppression. Cell Rep 18:496–507
    [Google Scholar]
  64. 64.  Oh J, Al-Zain A, Cannavo E, Cejka P, Symington LS 2016. Xrs2 dependent and independent functions of the Mre11-Rad50 complex. Mol. Cell 64:405–15
    [Google Scholar]
  65. 65.  Crown KN, Savytskyy OP, Malik SB, Logsdon J, Williams RS et al. 2013. A mutation in the FHA domain of Coprinus cinereus Nbs1 leads to Spo11-independent meiotic recombination and chromosome segregation. G3 3:1927–43
    [Google Scholar]
  66. 66.  Makharashvili N, Paull TT 2015. CtIP: a DNA damage response protein at the intersection of DNA metabolism. DNA Repair 32:75–81
    [Google Scholar]
  67. 67.  Sartori AA, Lukas C, Coates J, Mistrik M, Fu S et al. 2007. Human CtIP promotes DNA end resection. Nature 450:509–14
    [Google Scholar]
  68. 68.  Anand R, Ranjha L, Cannavo E, Cejka P 2016. Phosphorylated CtIP functions as a co-factor of the MRE11-RAD50-NBS1 endonuclease in DNA end resection. Mol. Cell 64:940–50
    [Google Scholar]
  69. 69.  Liao S, Tammaro M, Yan H 2016. The structure of ends determines the pathway choice and Mre11 nuclease dependency of DNA double-strand break repair. Nucleic Acids Res 44:5689–701
    [Google Scholar]
  70. 70.  Jensen KL, Russell P 2016. Ctp1-dependent clipping and resection of DNA double-strand breaks by Mre11 endonuclease complex are not genetically separable. Nucleic Acids Res 44:8241–49
    [Google Scholar]
  71. 71.  Wang Q, Goldstein M, Alexander P, Wakeman TP, Sun T 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]
  72. 72.  Deshpande RA, Lee J-H, Arora S, Paull TT 2016. Nbs1 converts the human Mre11/Rad50 nuclease complex into an endo/exonuclease machine specific for protein-DNA adducts. Mol. Cell 64:593–606
    [Google Scholar]
  73. 73.  Cannavo E, Cejka P 2014. Sae2 promotes dsDNA endonuclease activity within Mre11–Rad50–Xrs2 to resect DNA breaks. Nature 514:122–25
    [Google Scholar]
  74. 74.  Eisenstein M. 2015. The field that came in from the cold. Nat. Methods 13:19–22
    [Google Scholar]
  75. 75.  Gradia SD, Ishida JP, Tsai M-S, Jeans C, Tainer JA, Fuss JO 2017. MacroBac: new technologies for robust and efficient large-scale production of recombinant multiprotein complexes. Methods Enzymol 592:1–26
    [Google Scholar]
  76. 76.  Williams GJ, Hammel M, Radhakrishnan SK, Ramsden D, Lees-Miller SP, Tainer JA 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]
  77. 77.  Jackson SP, Bartek J 2009. The DNA-damage response in human biology and disease. Nature 461:1071–78
    [Google Scholar]
  78. 78.  Lomax ME, Folkes LK, O'Neill P 2013. Biological consequences of radiation-induced DNA damage: relevance to radiotherapy. Clin. Oncol. 25:578–85
    [Google Scholar]
  79. 79.  Paull TT. 2015. Mechanisms of ATM activation. Annu. Rev. Biochem. 84:711–38
    [Google Scholar]
  80. 80.  Wu L, Luo K, Lou Z, Chen J 2008. MDC1 regulates intra-S-phase checkpoint by targeting NBS1 to DNA double-strand breaks. PNAS 105:11200–5
    [Google Scholar]
  81. 81.  Pennisi R, Antoccia A, Leone S, Ascenzi P, di Masi A 2017. Hsp90α regulates ATM and NBN functions in sensing and repair of DNA double-strand breaks. FEBS J 284:2378–95
    [Google Scholar]
  82. 82.  von Morgen P, Burdova K, Flower TG, O'Reilly NJ, Boulton SJ et al. 2017. MRE11 stability is regulated by CK2-dependent interaction with R2TP complex. Oncogene 36:4943–50
    [Google Scholar]
  83. 83.  Lu C-S, Truong LN, Aslanian A, Shi LZ, Li Y et al. 2012. The RING finger protein RNF8 ubiquitinates Nbs1 to promote DNA double-strand break repair by homologous recombination. J. Biol. Chem. 287:43984–94
    [Google Scholar]
  84. 84.  Li Z, Li J, Kong Y, Yan S, Ahmad N, Liu X 2017. Plk1 phosphorylation of Mre11 antagonizes the DNA damage response. Cancer Res 77:3169–80
    [Google Scholar]
  85. 85.  Zhang T, Penicud K, Bruhn C, Loizou JI, Kanu N et al. 2012. Competition between NBS1 and ATMIN controls ATM signaling pathway choice. Cell Rep 2:1498–504
    [Google Scholar]
  86. 86.  Zhang T, Cronshaw J, Kanu N, Snijders AP, Behrens A 2014. UBR5-mediated ubiquitination of ATMIN is required for ionizing radiation-induced ATM signaling and function. PNAS 111:12091–96
    [Google Scholar]
  87. 87.  Staples CJ, Barone G, Myers KN, Ganesh A, Gibbs-Seymour I et al. 2016. MRNIP/C5orf45 interacts with the MRN complex and contributes to the DNA damage response. Cell Rep 16:2565–75
    [Google Scholar]
  88. 88.  Abraham RT. 2001. Cell cycle checkpoint signaling through the ATM and ATR kinases. Genes Dev 15:2177–96
    [Google Scholar]
  89. 89.  Shin DS, Chahwan C, Huffman JL, Tainer JA 2004. Structure and function of the double-strand break repair machinery. DNA Repair 3:863–73
    [Google Scholar]
  90. 90.  Mahaney BL, Meek K, Lees-Miller SP 2009. Repair of ionizing radiation-induced DNA double-strand breaks by non-homologous end-joining. Biochem. J. 417:639–50
    [Google Scholar]
  91. 91.  Sfeir A, Symington LS 2015. Microhomology-mediated end joining: A back-up survival mechanism or dedicated pathway?. Trends Biochem. Sci. 40:701–14
    [Google Scholar]
  92. 92.  Biehs R, Steinlage M, Barton O, Juhász S, Künzel J et al. 2017. DNA double-strand break resection occurs during non-homologous end joining in G1 but is distinct from resection during homologous recombination. Mol. Cell 65:671–84.e5
    [Google Scholar]
  93. 93.  Dutta A, Eckelmann B, Adhikari S, Ahmed KM, Sengupta S et al. 2016. Microhomology-mediated end joining is activated in irradiated human cells due to phosphorylation-dependent formation of the XRCC1 repair complex. Nucleic Acids Res 45:2585–99
    [Google Scholar]
  94. 94.  Sharma S, Javadekar SM, Pandey M, Srivastava M, Kumari R, Raghavan SC 2015. Homology and enzymatic requirements of microhomology-dependent alternative end joining. Cell Death Dis 6:e1697
    [Google Scholar]
  95. 95.  Shamanna RA, Lu H, de Freitas JK, Tian J, Croteau DL, Bohr VA 2016. WRN regulates pathway choice between classical and alternative non-homologous end joining. Nat. Commun. 7:13785
    [Google Scholar]
  96. 96.  Grabarz A, Guirouilh-Barbat J, Barascu A, Pennarun G, Genet D et al. 2013. A role for BLM in double-strand break repair pathway choice: prevention of CtIP/Mre11-mediated alternative nonhomologous end-joining. Cell Rep 5:21–28
    [Google Scholar]
  97. 97.  Tomimatsu N, Mukherjee B, Deland K, Kurimasa A, Bolderson E et al. 2012. Exo1 plays a major role in DNA end resection in humans and influences double-strand break repair and damage signaling decisions. DNA Repair 11:441–48
    [Google Scholar]
  98. 98.  Tomimatsu N, Mukherjee B, Hardebeck MC, Ilcheva M, Camacho CV et al. 2014. Phosphorylation of EXO1 by CDKs 1 and 2 regulates DNA end resection and repair pathway choice. Nat. Commun. 5:3561
    [Google Scholar]
  99. 99.  Myler LR, Gallardo IF, Zhou Y, Gong F, Yang S-H et al. 2016. Single-molecule imaging reveals the mechanism of Exo1 regulation by single-stranded DNA binding proteins. PNAS 113:E1170–79
    [Google Scholar]
  100. 100.  Genschel J, Modrich P 2003. Mechanism of 5′-directed excision in human mismatch repair. Mol. Cell 12:1077–86
    [Google Scholar]
  101. 101.  Lu H, Shamanna RA, Keijzers G, Anand R, Rasmussen LJ et al. 2016. RECQL4 promotes DNA end resection in repair of DNA double-strand breaks. Cell Rep 16:161–73
    [Google Scholar]
  102. 102.  Broderick R, Nieminuszczy J, Baddock HT, Deshpande RA, Gileadi O et al. 2016. EXD2 promotes homologous recombination by facilitating DNA end resection. Nat. Cell Biol. 18:271–80
    [Google Scholar]
  103. 103.  Silva J, Aivio S, Knobel PA, Bailey LJ, Casali A et al. 2018. EXD2 governs germ stem cell homeostasis and lifespan by promoting mitoribosome integrity and translation. Nat. Cell Biol. 20:162–74
    [Google Scholar]
  104. 104.  Puddu F, Oelschlaegel T, Guerini I, Geisler NJ, Niu H et al. 2015. Synthetic viability genomic screening defines Sae2 function in DNA repair. EMBO J 34:1509–22
    [Google Scholar]
  105. 105.  Chen H, Donnianni RA, Handa N, Deng SK, Oh J et al. 2015. Sae2 promotes DNA damage resistance by removing the Mre11–Rad50–Xrs2 complex from DNA and attenuating Rad53 signaling. PNAS 112:E1880–87
    [Google Scholar]
  106. 106.  Nitiss JL, Ferrari M, Dibitetto D, De Gregorio G, Eapen VV et al. 2015. Functional interplay between the 53BP1-ortholog Rad9 and the Mre11 complex regulates resection, end-tethering and repair of a double-strand break. PLOS Genet 11:e1004928
    [Google Scholar]
  107. 107.  Bunting SF, Callén E, Wong N, Chen H-T, Polato F et al. 2010. 53BP1 inhibits homologous recombination in Brca1-deficient cells by blocking resection of DNA breaks. Cell 141:243–54
    [Google Scholar]
  108. 108.  Bothmer A, Robbiani DF, Feldhahn N, Gazumyan A, Nussenzweig A, Nussenzweig MC 2010. 53BP1 regulates DNA resection and the choice between classical and alternative end joining during class switch recombination. J. Exp. Med. 207:855–65
    [Google Scholar]
  109. 109.  Bouwman P, Aly A, Escandell JM, Pieterse M, Bartkova J et al. 2010. 53BP1 loss rescues BRCA1 deficiency and is associated with triple-negative and BRCA-mutated breast cancers. Nat. Struct. Mol. Biol. 17:688–95
    [Google Scholar]
  110. 110.  Daley JM, Sung P 2014. 53BP1, BRCA1, and the choice between recombination and end joining at DNA double-strand breaks. Mol. Cell. Biol. 34:1380–88
    [Google Scholar]
  111. 111.  Thompson LH. 2012. Recognition, signaling, and repair of DNA double-strand breaks produced by ionizing radiation in mammalian cells: the molecular choreography. Mutat. Res. 751:158–246
    [Google Scholar]
  112. 112.  Prakash R, Zhang Y, Feng W, Jasin M 2015. Homologous recombination and human health: the roles of BRCA1, BRCA2, and associated proteins. Cold Spring Harb. Perspect. Biol. 7:a016600
    [Google Scholar]
  113. 113.  Symington LS. 2016. Mechanism and regulation of DNA end resection in eukaryotes. Crit. Rev. Biochem. Mol. Biol. 51:195–212
    [Google Scholar]
  114. 114.  Zhao W, Steinfeld JB, Liang F, Chen X, Maranon DG et al. 2017. BRCA1–BARD1 promotes RAD51-mediated homologous DNA pairing. Nature 550:360–65
    [Google Scholar]
  115. 115.  Takeda S, Hoa NN, Sasanuma H 2016. The role of the Mre11–Rad50–Nbs1 complex in double-strand break repair—facts and myths. J. Radiat. Res. 57:i25–32
    [Google Scholar]
  116. 116.  Branzei D, Foiani M 2008. Regulation of DNA repair throughout the cell cycle. Nat. Rev. Mol. Cell Biol. 9:297–308
    [Google Scholar]
  117. 117.  Aceytuno RD, Piett CG, Havali-Shahriari Z, Edwards RA, Rey M et al. 2017. Structural and functional characterization of the PNKP–XRCC4–LigIV DNA repair complex. Nucleic Acids Res 45:6238–51
    [Google Scholar]
  118. 118.  Aparicio T, Baer R, Gottesman M, Gautier J 2016. MRN, CtIP, and BRCA1 mediate repair of topoisomerase II–DNA adducts. J. Cell Biol. 212:399–408
    [Google Scholar]
  119. 119.  Lee KC, Padget K, Curtis H, Cowell IG, Moiani D et al. 2012. MRE11 facilitates the removal of human topoisomerase II complexes from genomic DNA. Biol. Open 1:863–73
    [Google Scholar]
  120. 120.  Wang H, Li Y, Truong LN, Shi LZ, Hwang PY-H et al. 2014. CtIP maintains stability at common fragile sites and inverted repeats by end resection-independent endonuclease activity. Mol. Cell 54:1012–21
    [Google Scholar]
  121. 121.  Zeman MK, Cimprich KA 2014. Causes and consequences of replication stress. Nat. Cell Biol. 16:2–9
    [Google Scholar]
  122. 122.  Neelsen KJ, Lopes M 2015. Replication fork reversal in eukaryotes: from dead end to dynamic response. Nat. Rev. Mol. Cell Biol. 16:207–20
    [Google Scholar]
  123. 123.  Aze A, Zhou JC, Costa A, Costanzo V 2013. DNA replication and homologous recombination factors: acting together to maintain genome stability. Chromosoma 122:401–13
    [Google Scholar]
  124. 124.  Carr AM, Lambert S 2013. Replication stress-induced genome instability: the dark side of replication maintenance by homologous recombination. J. Mol. Biol. 425:4733–44
    [Google Scholar]
  125. 125.  Kawabata T, Luebben SW, Yamaguchi S, Ilves I, Matise I et al. 2011. Stalled fork rescue via dormant replication origins in unchallenged S phase promotes proper chromosome segregation and tumor suppression. Mol. Cell 41:543–53
    [Google Scholar]
  126. 126.  Woodward AM, Göhler T, Luciani MG, Oehlmann M, Ge X et al. 2006. Excess Mcm2–7 license dormant origins of replication that can be used under conditions of replicative stress. J. Cell Biol. 173:673–83
    [Google Scholar]
  127. 127.  Ge XQ, Jackson DA, Blow JJ 2007. Dormant origins licensed by excess Mcm2–7 are required for human cells to survive replicative stress. Genes Dev 21:3331–41
    [Google Scholar]
  128. 128.  Blow JJ, Ge XQ, Jackson DA 2011. How dormant origins promote complete genome replication. Trends Biochem. Sci. 36:405–14
    [Google Scholar]
  129. 129.  Xu X, Vaithiyalingam S, Glick GG, Mordes DA, Chazin WJ, Cortez D 2008. The basic cleft of RPA70N binds multiple checkpoint proteins, including RAD9, to regulate ATR signaling. Mol. Cell. Biol. 28:7345–53
    [Google Scholar]
  130. 130.  Fanning E. 2006. A dynamic model for replication protein A (RPA) function in DNA processing pathways. Nucleic Acids Res 34:4126–37
    [Google Scholar]
  131. 131.  Bolderson E, Petermann E, Croft L, Suraweera A, Pandita RK et al. 2014. Human single-stranded DNA binding protein 1 (hSSB1/NABP2) is required for the stability and repair of stalled replication forks. Nucleic Acids Res 42:6326–36
    [Google Scholar]
  132. 132.  Deng SK, Chen H, Symington LS 2015. Replication protein A prevents promiscuous annealing between short sequence homologies: implications for genome integrity. BioEssays 37:305–13
    [Google Scholar]
  133. 133.  Oakley GG, Tillison K, Opiyo SA, Glanzer JG, Horn JM, Patrick SM 2009. Physical interaction between replication protein A (RPA) and MRN: involvement of RPA2 phosphorylation and the N-terminus of RPA1. Biochemistry 48:7473–81
    [Google Scholar]
  134. 134.  Cimprich KA, Cortez D 2008. ATR: an essential regulator of genome integrity. Nat. Rev. Mol. Cell Biol. 9:616–27
    [Google Scholar]
  135. 135.  Duursma AM, Driscoll R, Elias JE, Cimprich KA 2013. A role for the MRN complex in ATR activation via TOPBP1 recruitment. Mol. Cell 50:116–22
    [Google Scholar]
  136. 136.  Lee J, Dunphy WG 2013. The Mre11-Rad50-Nbs1 (MRN) complex has a specific role in the activation of Chk1 in response to stalled replication forks. Mol. Biol. Cell 24:1343–53
    [Google Scholar]
  137. 137.  Myers JS, Cortez D 2006. Rapid activation of ATR by ionizing radiation requires ATM and Mre11. J. Biol. Chem. 281:9346–50
    [Google Scholar]
  138. 138.  Jazayeri A, Falck J, Lukas C, Bartek J, Smith GCM et al. 2005. ATM- and cell cycle-dependent regulation of ATR in response to DNA double-strand breaks. Nat. Cell Biol. 8:37–45
    [Google Scholar]
  139. 139.  Kobayashi M, Hayashi N, Takata M Yamamoto K-i 2013. NBS1 directly activates ATR independently of MRE11 and TOPBP1. Genes Cells 18:238–46
    [Google Scholar]
  140. 140.  Shiotani B, Nguyen Hai D, Håkansson P, Maréchal A, Tse A et al. 2013. Two distinct modes of ATR activation orchestrated by Rad17 and Nbs1. Cell Rep 3:1651–62
    [Google Scholar]
  141. 141.  Seeber A, Hegnauer AM, Hustedt N, Deshpande I, Poli J et al. 2016. RPA mediates recruitment of MRX to forks and double-strand breaks to hold sister chromatids together. Mol. Cell 64:951–66
    [Google Scholar]
  142. 142.  Gatei M, Kijas AW, Biard D, Dörk T, Lavin MF 2014. RAD50 phosphorylation promotes ATR downstream signaling and DNA restart following replication stress. Hum. Mol. Genet. 23:4232–48
    [Google Scholar]
  143. 143.  Tittel-Elmer M, Lengronne A, Davidson MB, Bacal J, François P et al. 2012. Cohesin association to replication sites depends on Rad50 and promotes fork restart. Mol. Cell 48:98–108
    [Google Scholar]
  144. 144.  Palermo V, Rinalducci S, Sanchez M, Grillini F, Sommers JA et al. 2016. CDK1 phosphorylates WRN at collapsed replication forks. Nat. Commun. 7:12880
    [Google Scholar]
  145. 145.  Hashimoto Y, Chaudhuri AR, Lopes M, Costanzo V 2010. Rad51 protects nascent DNA from Mre11-dependent degradation and promotes continuous DNA synthesis. Nat. Struct. Mol. Biol. 17:1305–11
    [Google Scholar]
  146. 146.  Vallerga MB, Mansilla SF, Federico MB, Bertolin AP, Gottifredi V 2015. Rad51 recombinase prevents Mre11 nuclease-dependent degradation and excessive PrimPol-mediated elongation of nascent DNA after UV irradiation. PNAS 112:E6624–33
    [Google Scholar]
  147. 147.  Schlacher K, Christ N, Siaud N, Egashira A, Wu H, Jasin M 2011. Double-strand break repair-independent role for BRCA2 in blocking stalled replication fork degradation by MRE11. Cell 145:529–42
    [Google Scholar]
  148. 148.  Schlacher K, Wu H, Jasin M 2012. A distinct replication fork protection pathway connects Fanconi anemia tumor suppressors to RAD51-BRCA1/2. Cancer Cell 22:106–16
    [Google Scholar]
  149. 149.  Kim TM, Son MY, Dodds S, Hu L, Luo G, Hasty P 2014. RECQL5 and BLM exhibit divergent functions in cells defective for the Fanconi anemia pathway. Nucleic Acids Res 43:893–903
    [Google Scholar]
  150. 150.  Huh MS, Ivanochko D, Hashem LE, Curtin M, Delorme M et al. 2016. Stalled replication forks within heterochromatin require ATRX for protection. Cell Death Dis 7:e2220
    [Google Scholar]
  151. 151.  Kolinjivadi AM, Sannino V, De Antoni A, Zadorozhny K, Kilkenny M et al. 2017. Smarcal1-mediated fork reversal triggers Mre11-dependent degradation of nascent DNA in the absence of Brca2 and stable Rad51 nucleofilaments. Mol. Cell 67:867–81
    [Google Scholar]
  152. 152.  Pefani DE, O'Neill E 2015. Safeguarding genome stability: RASSF1A tumor suppressor regulates BRCA2 at stalled forks. Cell Cycle 14:1624–30
    [Google Scholar]
  153. 153.  Leuzzi G, Marabitti V, Pichierri P, Franchitto A 2016. WRNIP1 protects stalled forks from degradation and promotes fork restart after replication stress. EMBO J 35:1437–51
    [Google Scholar]
  154. 154.  Trego KS, Chernikova SB, Davalos AR, Perry JJP, Finger LD et al. 2014. The DNA repair endonuclease XPG interacts directly and functionally with the WRN helicase defective in Werner syndrome. Cell Cycle 10:1998–2007
    [Google Scholar]
  155. 155.  Perry JJP, Asaithamby A, Barnebey A, Kiamanesch F, Chen DJ et al. 2010. Identification of a coiled coil in Werner syndrome protein that facilitates multimerization and promotes exonuclease processivity. J. Biol. Chem. 285:25699–707
    [Google Scholar]
  156. 156.  Perry JJP, Yannone SM, Holden LG, Hitomi C, Asaithamby A et al. 2006. WRN exonuclease structure and molecular mechanism imply an editing role in DNA end processing. Nat. Struct. Mol. Biol. 13:414–22
    [Google Scholar]
  157. 157.  Iannascoli C, Palermo V, Murfuni I, Franchitto A, Pichierri P 2015. The WRN exonuclease domain protects nascent strands from pathological MRE11/EXO1-dependent degradation. Nucleic Acids Res 43:9788–803
    [Google Scholar]
  158. 158.  Chaudhuri AR, Callen E, Ding X, Gogola E, Duarte AA et al. 2016. Replication fork stability confers chemoresistance in BRCA-deficient cells. Nature 535:382–87
    [Google Scholar]
  159. 159.  Roy S, Tomaszowski K-H, Luzwick JW, Park S, Li J et al. 2018. p53 orchestrates DNA replication restart homeostasis by suppressing mutagenic RAD52 and POLθ pathways. eLife 7:e31723
    [Google Scholar]
  160. 160.  Zhu X-D, Küster B, Mann M, Petrini JHJ, de Lange T 2000. Cell-cycle-regulated association of RAD50/MRE11/NBS1 with TRF2 and human telomeres. Nat. Genet. 25:347–52
    [Google Scholar]
  161. 161.  O'Sullivan RJ, Karlseder J 2010. Telomeres: protecting chromosomes against genome instability. Nat. Rev. Mol. Cell Biol. 11:171–81
    [Google Scholar]
  162. 162.  Levy MZ, Allsopp RC, Futcher AB, Greider CW, Harley CB 1992. Telomere end-replication problem and cell aging. J. Mol. Biol. 225:951–60
    [Google Scholar]
  163. 163.  Wu P, Takai H, de Lange T 2012. Telomeric 3′ overhangs derive from resection by Exo1 and Apollo and fill-in by POT1b-associated CST. Cell 150:39–52
    [Google Scholar]
  164. 164.  Shippen-Lentz D, Blackburn EH 1990. Functional evidence for an RNA template in telomerase. Science 247:546–52
    [Google Scholar]
  165. 165.  Hastie ND, Dempster M, Dunlop MG, Thompson AM, Green DK, Allshire RC 1990. Telomere reduction in human colorectal carcinoma and with ageing. Nature 346:866–68
    [Google Scholar]
  166. 166.  Harley CB, Futcher AB, Greider CW 1990. Telomeres shorten during ageing of human fibroblasts. Nature 345:458–60
    [Google Scholar]
  167. 167.  Allsopp RC, Vaziri H, Patterson C, Goldstein S, Younglai EV et al. 1992. Telomere length predicts replicative capacity of human fibroblasts. PNAS 89:10114–18
    [Google Scholar]
  168. 168.  Takata H, Tanaka Y, Matsuura A 2005. Late S phase-specific recruitment of Mre11 complex triggers hierarchical assembly of telomere replication proteins in Saccharomyces cerevisiae. Mol. Cell 17:573–83
    [Google Scholar]
  169. 169.  Rivera T, Haggblom C, Cosconati S, Karlseder J 2016. A balance between elongation and trimming regulates telomere stability in stem cells. Nat. Struct. Mol. Biol. 24:30–39
    [Google Scholar]
  170. 170.  Wu Z, Liu J, Zhang Q-D, Lv D-K, Wu N-F, Zhou J-Q 2017. Rad6–Bre1-mediated H2B ubiquitination regulates telomere replication by promoting telomere-end resection. Nucleic Acids Res 45:3308–22
    [Google Scholar]
  171. 171.  Marcand S. 2014. How do telomeres and NHEJ coexist?. Mol. Cell. Oncol. 1:e963438
    [Google Scholar]
  172. 172.  Marcomini I, Gasser SM 2015. Nuclear organization in DNA end processing: telomeres versus double-strand breaks. DNA Repair 32:134–40
    [Google Scholar]
  173. 173.  Williamson JR, Raghuraman MK, Cech TR 1989. Monovalent cation-induced structure of telomeric DNA: the G-quartet model. Cell 59:871–80
    [Google Scholar]
  174. 174.  Higa M, Fujita M, Yoshida K 2017. DNA replication origins and fork progression at mammalian telomeres. Genes 8:112
    [Google Scholar]
  175. 175.  Denchi EL. 2009. Give me a break: How telomeres suppress the DNA damage response. DNA Repair 8:1118–26
    [Google Scholar]
  176. 176.  Palm W, de Lange T 2008. How shelterin protects mammalian telomeres. Annu. Rev. Genet. 42:301–34
    [Google Scholar]
  177. 177.  Denchi EL, de Lange T 2007. Protection of telomeres through independent control of ATM and ATR by TRF2 and POT1. Nature 448:1068–71
    [Google Scholar]
  178. 178.  Déjardin J, Kingston RE 2009. Purification of proteins associated with specific genomic loci. Cell 136:175–86
    [Google Scholar]
  179. 179.  Rai R, Hu C, Broton C, Chen Y, Lei M, Chang S 2017. NBS1 phosphorylation status dictates repair choice of dysfunctional telomeres. Mol. Cell 65:801–17.e4
    [Google Scholar]
  180. 180.  Cesare AJ, Reddel RR 2010. Alternative lengthening of telomeres: models, mechanisms and implications. Nat. Rev. Genet. 11:319–30
    [Google Scholar]
  181. 181.  Shay JW, Reddel RR, Wright WE 2012. Cancer and telomeres–an ALTernative to telomerase. Science 336:1388–90
    [Google Scholar]
  182. 182.  Xue Y, Marvin ME, Ivanova IG, Lydall D, Louis EJ, Maringele L 2016. Rif1 and Exo1 regulate the genomic instability following telomere losses. Aging Cell 15:553–62
    [Google Scholar]
  183. 183.  Ballew BJ, Lundblad V 2013. Multiple genetic pathways regulate replicative senescence in telomerase-deficient yeast. Aging Cell 12:719–27
    [Google Scholar]
  184. 184.  Durocher D, Cassani C, Gobbini E, Wang W, Niu H et al. 2016. Tel1 and Rif2 regulate MRX functions in end-tethering and repair of DNA double-strand breaks. PLOS Biol 14:e1002387
    [Google Scholar]
  185. 185.  Saretzki G, Panero J, Stella F, Schutz N, Fantl DB, Slavutsky I 2015. Differential expression of non-shelterin genes associated with high telomerase levels and telomere shortening in plasma cell disorders. PLOS ONE 10:e0137972
    [Google Scholar]
  186. 186.  Hoxha M, Fabris S, Agnelli L, Bollati V, Cutrona G et al. 2014. Relevance of telomere/telomerase system impairment in early stage chronic lymphocytic leukemia. Genes Chromosomes Cancer 53:612–21
    [Google Scholar]
  187. 187.  Heng J, Zhang F, Guo X, Tang L, Peng L et al. 2017. Integrated analysis of promoter methylation and expression of telomere related genes in breast cancer. Oncotarget 8:25442–54
    [Google Scholar]
  188. 188.  Attwooll CL, Akpinar M, Petrini JH 2009. The Mre11 complex and the response to dysfunctional telomeres. Mol. Cell. Biol. 29:5540–51
    [Google Scholar]
  189. 189.  Li Y, Shen Y, Hohensinner P, Ju J, Wen Z et al. 2016. Deficient activity of the nuclease MRE11A induces T cell aging and promotes arthritogenic effector functions in patients with rheumatoid arthritis. Immunity 45:903–16
    [Google Scholar]
  190. 190.  Stracker TH, Carson CT, Weitzman MD 2002. Adenovirus oncoproteins inactivate the Mre11–Rad50–NBS1 DNA repair complex. Nature 418:348–52
    [Google Scholar]
  191. 191.  Lilley CE, Schwartz RA, Weitzman MD 2007. Using or abusing: viruses and the cellular DNA damage response. Trends Microbiol 15:119–26
    [Google Scholar]
  192. 192.  Polo SE, Jackson SP 2011. Dynamics of DNA damage response proteins at DNA breaks: a focus on protein modifications. Genes Dev 25:409–33
    [Google Scholar]
  193. 193.  Shah GA, O'Shea CC 2015. Viral and cellular genomes activate distinct DNA damage responses. Cell 162:987–1002
    [Google Scholar]
  194. 194.  Weiden MD, Ginsberg HS 1994. Deletion of the E4 region of the genome produces adenovirus DNA concatemers. PNAS 91:153–57
    [Google Scholar]
  195. 195.  Boyer J, Rohleder K, Ketner G 1999. Adenovirus E4 34k and E4 11k inhibit double strand break repair and are physically associated with the cellular DNA-dependent protein kinase. Virology 263:307–12
    [Google Scholar]
  196. 196.  Rawlins DR, Rosenfeld PJ, Wides RJ, Challberg MD, Kelly TJ 1984. Structure and function of the adenovirus origin of replication. Cell 37:309–19
    [Google Scholar]
  197. 197.  Bernstein JA, Porter JM, Challberg MD 1986. Template requirements for in vivo replication of adenovirus DNA. Mol. Cell. Biol. 6:2115–24
    [Google Scholar]
  198. 198.  Kondo T, Kobayashi J, Saitoh T, Maruyama K, Ishii KJ et al. 2013. DNA damage sensor MRE11 recognizes cytosolic double-stranded DNA and induces type I interferon by regulating STING trafficking. PNAS 110:2969–74
    [Google Scholar]
  199. 199.  Bhattacharya S, Srinivasan K, Abdisalaam S, Su F, Raj P et al. 2017. RAD51 interconnects between DNA replication, DNA repair and immunity. Nucleic Acids Res 45:4590–605
    [Google Scholar]
  200. 200.  Liu X, Hammel M, He Y, Tainer JA, Jeng US et al. 2013. Structural insights into the interaction of IL-33 with its receptors. PNAS 110:14918–23
    [Google Scholar]
  201. 201.  Stav-Noraas TE, Edelmann RJ, Poulsen LLC, Sundnes O, Phung D et al. 2017. Endothelial IL-33 expression is augmented by adenoviral activation of the DNA damage machinery. J. Immunol. 198:3318–25
    [Google Scholar]
  202. 202.  Roth S, Rottach A, Lotz-Havla AS, Laux V, Muschaweckh A et al. 2014. Rad50-CARD9 interactions link cytosolic DNA sensing to IL-1β production. Nat. Immunol. 15:538–45
    [Google Scholar]
  203. 203.  Sohn S-Y, Hearing P 2012. Adenovirus regulates Sumoylation of Mre11-Rad50-Nbs1 components through a paralog-specific mechanism. J. Virol. 86:9656–65
    [Google Scholar]
  204. 204.  Bridge E, Ketner G 1989. Redundant control of adenovirus late gene expression by early region 4. J. Virol. 63:631–38
    [Google Scholar]
  205. 205.  Halbert DN, Cutt JR, Shenk T 1985. Adenovirus early region 4 encodes functions required for efficient DNA replication, late gene expression, and host cell shutoff. J. Virol. 56:250–57
    [Google Scholar]
  206. 206.  Huang MM, Hearing P 1989. Adenovirus early region 4 encodes two gene products with redundant effects in lytic infection. J. Virol. 63:2605–15
    [Google Scholar]
  207. 207.  Weinberg DH, Ketner G 1986. Adenoviral early region 4 is required for efficient viral DNA replication and for late gene expression. J. Virol. 57:833–38
    [Google Scholar]
  208. 208.  Evans JD, Hearing P 2005. Relocalization of the Mre11-Rad50-Nbs1 complex by the adenovirus E4 ORF3 protein is required for viral replication. J. Virol. 79:6207–15
    [Google Scholar]
  209. 209.  Ou HD, Kwiatkowski W, Deerinck TJ, Noske A, Blain KY et al. 2012. A structural basis for the assembly and functions of a viral polymer that inactivates multiple tumor suppressors. Cell 151:304–19
    [Google Scholar]
  210. 210.  Sohn SY, Hearing P 2012. Adenovirus regulates sumoylation of Mre11-Rad50-Nbs1 components through a paralog-specific mechanism. J. Virol. 86:9656–65
    [Google Scholar]
  211. 211.  Bridges RG, Sohn SY, Wright J, Leppard KN, Hearing P 2016. The adenovirus E4-ORF3 protein stimulates SUMOylation of general transcription factor TFII-I to direct proteasomal degradation. mBio 7:e02184–15
    [Google Scholar]
  212. 212.  Harada JN, Shevchenko A, Shevchenko A, Pallas DC, Berk AJ 2002. Analysis of the adenovirus E1B-55K-anchored proteome reveals its link to ubiquitination machinery. J. Virol. 76:9194–206
    [Google Scholar]
  213. 213.  Querido E. 2001. Degradation of p53 by adenovirus E4orf6 and E1B55K proteins occurs via a novel mechanism involving a Cullin-containing complex. Genes Dev 15:3104–17
    [Google Scholar]
  214. 214.  zur Hausen H. 2009. Papillomaviruses in the causation of human cancers—a brief historical account. Virology 384:260–65
    [Google Scholar]
  215. 215.  Johnson BA, Aloor HL, Moody CA 2017. The Rb binding domain of HPV31 E7 is required to maintain high levels of DNA repair factors in infected cells. Virology 500:22–34
    [Google Scholar]
  216. 216.  Campos-León K, Wijendra K, Siddiqa A, Pentland I, Feeney KM et al. 2017. Association of human papillomavirus 16 E2 with Rad50-interacting protein 1 enhances viral DNA replication. J. Virol. 91:e02305–16
    [Google Scholar]
  217. 217.  Robertson ES, Mariggiò G, Koch S, Zhang G, Weidner-Glunde M et al. 2017. Kaposi Sarcoma Herpesvirus (KSHV) latency-associated nuclear antigen (LANA) recruits components of the MRN (Mre11-Rad50-NBS1) repair complex to modulate an innate immune signaling pathway and viral latency. PLOS Pathog 13:e1006335
    [Google Scholar]
  218. 218.  Chang Y, Cesarman E, Pessin MS, Lee F, Culpepper J et al. 1994. Identification of herpesvirus-like DNA sequences in AIDS-associated Kaposi's sarcoma. Science 266:1865–69
    [Google Scholar]
  219. 219.  Bouvard V, Baan R, Straif K, Grosse Y, Secretan B et al. 2009. A review of human carcinogens–Part B: biological agents. Lancet Oncol 10:321–22
    [Google Scholar]
  220. 220.  Pantelidou C, Cherubini G, Lemoine NR, Halldén G 2016. The E1B19K-deleted oncolytic adenovirus mutant AdΔ19K sensitizes pancreatic cancer cells to drug-induced DNA-damage by down-regulating Claspin and Mre11. Oncotarget 7:15703–24
    [Google Scholar]
  221. 221.  Tookman LA, Browne AK, Connell CM, Bridge G, Ingemarsdotter CK et al. 2015. RAD51 and BRCA2 enhance oncolytic adenovirus type 5 activity in ovarian cancer. Mol. Cancer Res. 14:44–55
    [Google Scholar]
  222. 222.  Crick F. 1974. The double helix: a personal view. Nature 248:766–69
    [Google Scholar]
  223. 223.  Harrison SC. 2007. Comments on the NIGMS PSI. Structure 15:1344–46
    [Google Scholar]
  224. 224.  Hura GL, Menon AL, Hammel M, Rambo RP, Poole FL II et al. 2009. Robust, high-throughput solution structural analyses by small angle X-ray scattering (SAXS). Nat. Methods 6:606–12
    [Google Scholar]
  225. 225.  Rambo RP, Tainer JA 2011. Characterizing flexible and intrinsically unstructured biological macromolecules by SAS using the Porod-Debye law. Biopolymers 95:559–71
    [Google Scholar]
  226. 226.  Hura GL, Budworth H, Dyer KN, Rambo RP, Hammel M et al. 2013. Comprehensive macromolecular conformations mapped by quantitative SAXS analyses. Nat. Methods 10:453–54
    [Google Scholar]
  227. 227.  Rambo RP, Tainer JA 2013. Accurate assessment of mass, models and resolution by small-angle scattering. Nature 496:477–81
    [Google Scholar]
  228. 228.  Rambo RP, Tainer JA 2013. Super-resolution in solution X-ray scattering and its applications to structural systems biology. Annu. Rev. Biophys. 42:415–41
    [Google Scholar]
  229. 229.  Brosey CA, Ho C, Long WZ, Singh S, Burnett K et al. 2016. Defining NADH-driven allostery regulating apoptosis-inducing factor. Structure 24:2067–79
    [Google Scholar]
  230. 230.  Bacolla A, Tainer JA, Vasquez KM, Cooper DN 2016. Translocation and deletion breakpoints in cancer genomes are associated with potential non-B DNA-forming sequences. Nucleic Acids Res 44:5673–88
    [Google Scholar]
  231. 231.  Tsutakawa SE, Thompson MJ, Arvai AS, Neil AJ, Shaw SJ et al. 2017. Phosphate steering by Flap Endonuclease 1 promotes 5′-flap specificity and incision to prevent genome instability. Nat. Commun. 8:15855
    [Google Scholar]
  232. 232.  Rashid F, Harris PD, Zaher MS, Sobhy MA, Joudeh LI et al. 2017. Single-molecule FRET unveils induced-fit mechanism for substrate selectivity in flap endonuclease 1. eLife 6:e21884
    [Google Scholar]
  233. 233.  Sundheim O, Vågbø CB, Bjørås M, Sousa MML, Talstad V et al. 2006. Human ABH3 structure and key residues for oxidative demethylation to reverse DNA/RNA damage. EMBO J 25:3389–97
    [Google Scholar]
  234. 234.  Dango S, Mosammaparast N, Sowa ME, Xiong L-J, Wu F et al. 2011. DNA unwinding by ASCC3 helicase is coupled to ALKBH3-dependent DNA alkylation repair and cancer cell proliferation. Mol. Cell 44:373–84
    [Google Scholar]
  235. 235.  Brosey CA, Ahmed Z, Lees-Miller SP, Tainer JA 2017. What combined measurements from structures and imaging tell us about DNA damage responses. Methods Enzymol 592:417–55
    [Google Scholar]
  236. 236.  Loeb LA, Loeb KR, Anderson JP 2003. Multiple mutations and cancer. PNAS 100:776–81
    [Google Scholar]
/content/journals/10.1146/annurev-biochem-062917-012415
Loading
/content/journals/10.1146/annurev-biochem-062917-012415
Loading

Data & Media loading...

Supplemental Material

Supplementary Data

  • 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