1932

Abstract

Viruses must interact with their hosts in order to replicate; these interactions often provoke the evolutionarily conserved response to DNA damage, known as the DNA damage response (DDR). The DDR can be activated by incoming viral DNA, during the integration of retroviruses, or in response to the aberrant DNA structures generated upon replication of DNA viruses. Furthermore, DNA and RNA viral proteins can induce the DDR by promoting inappropriate S phase entry, by modifying cellular DDR factors directly, or by unintentionally targeting host DNA. The DDR may be antiviral, although viruses often require proximal DDR activation of repair and recombination factors to facilitate replication as well as downstream DDR signaling suppression to ensure cell survival. An unintended consequence of DDR attenuation during infection is the long-term survival and proliferation of precancerous cells. Therefore, the molecular basis for DDR activation and attenuation by viruses remains an important area of study that will likely provide key insights into how viruses have evolved with their hosts.

Loading

Article metrics loading...

/content/journals/10.1146/annurev-virology-031413-085548
2014-09-29
2024-03-28
Loading full text...

Full text loading...

/deliver/fulltext/virology/1/1/annurev-virology-031413-085548.html?itemId=/content/journals/10.1146/annurev-virology-031413-085548&mimeType=html&fmt=ahah

Literature Cited

  1. Ciccia A, Elledge SJ. 1.  2010. The DNA damage response: making it safe to play with knives. Mol. Cell 40:179–204 [Google Scholar]
  2. Smith J, Tho LM, Xu N, Gillespie DA. 2.  2010. The ATM-Chk2 and ATR-Chk1 pathways in DNA damage signaling and cancer. Adv. Cancer Res. 108:73–112 [Google Scholar]
  3. Lee JH, Paull TT. 3.  2005. ATM activation by DNA double-strand breaks through the Mre11-Rad50-Nbs1 complex. Science 308:551–54 [Google Scholar]
  4. Bakkenist CJ, Kastan MB. 4.  2003. DNA damage activates ATM through intermolecular autophosphorylation and dimer dissociation. Nature 421:499–506 [Google Scholar]
  5. Sun Y, Jiang X, Chen S, Fernandes N, Price BD. 5.  2005. A role for the Tip60 histone acetyltransferase in the acetylation and activation of ATM. Proc. Natl. Acad. Sci. USA 102:13182–87 [Google Scholar]
  6. Burma S, Chen BP, Murphy M, Kurimasa A, Chen DJ. 6.  2001. ATM phosphorylates histone H2AX in response to DNA double-strand breaks. J. Biol. Chem. 276:42462–67 [Google Scholar]
  7. Stewart GS, Wang B, Bignell CR, Taylor AM, Elledge SJ. 7.  2003. MDC1 is a mediator of the mammalian DNA damage checkpoint. Nature 421:961–66 [Google Scholar]
  8. Doil C, Mailand N, Bekker-Jensen S, Menard P, Larsen DH. 8.  et al. 2009. RNF168 binds and amplifies ubiquitin conjugates on damaged chromosomes to allow accumulation of repair proteins. Cell 136:435–46 [Google Scholar]
  9. Mailand N, Bekker-Jensen S, Faustrup H, Melander F, Bartek J. 9.  et al. 2007. RNF8 ubiquitylates histones at DNA double-strand breaks and promotes assembly of repair proteins. Cell 131:887–900 [Google Scholar]
  10. Huen MS, Grant R, Manke I, Minn K, Yu X. 10.  et al. 2007. RNF8 transduces the DNA-damage signal via histone ubiquitylation and checkpoint protein assembly. Cell 131:901–14 [Google Scholar]
  11. Kolas NK, Chapman JR, Nakada S, Ylanko J, Chahwan R. 11.  et al. 2007. Orchestration of the DNA-damage response by the RNF8 ubiquitin ligase. Science 318:1637–40 [Google Scholar]
  12. Matsuoka S, Ballif BA, Smogorzewska A, McDonald ER III, Hurov KE. 12.  et al. 2007. ATM and ATR substrate analysis reveals extensive protein networks responsive to DNA damage. Science 316:1160–66 [Google Scholar]
  13. Nam EA, Cortez D. 13.  2011. ATR signalling: more than meeting at the fork. Biochem. J. 436:527–36 [Google Scholar]
  14. Zou L, Elledge SJ. 14.  2003. Sensing DNA damage through ATRIP recognition of RPA-ssDNA complexes. Science 300:1542–48 [Google Scholar]
  15. Kumagai A, Lee J, Yoo HY, Dunphy WG. 15.  2006. TopBP1 activates the ATR-ATRIP complex. Cell 124:943–55 [Google Scholar]
  16. Jeong SY, Kumagai A, Lee J, Dunphy WG. 16.  2003. Phosphorylated Claspin interacts with a phosphate-binding site in the kinase domain of Chk1 during ATR-mediated activation. J. Biol. Chem. 278:46782–88 [Google Scholar]
  17. Kumagai A, Dunphy WG. 17.  2003. Repeated phosphopeptide motifs in Claspin mediate the regulated binding of Chk1. Nat. Cell Biol. 5:161–65 [Google Scholar]
  18. Mahaney BL, Meek K, Lees-Miller SP. 18.  2009. Repair of ionizing radiation–induced DNA double-strand breaks by non-homologous end-joining. Biochem. J. 417:639–50 [Google Scholar]
  19. Burma S, Chen DJ. 19.  2004. Role of DNA-PK in the cellular response to DNA double-strand breaks. DNA Repair 3:909–18 [Google Scholar]
  20. Meek K, Dang V, Lees-Miller SP. 20.  2008. DNA-PK: the means to justify the ends?. Adv. Immunol. 99:33–58 [Google Scholar]
  21. Bartek J, Lukas J. 21.  2003. Chk1 and Chk2 kinases in checkpoint control and cancer. Cancer Cell 3:421–29 [Google Scholar]
  22. Stracker TH, Usui T, Petrini JH. 22.  2009. Taking the time to make important decisions: the checkpoint effector kinases Chk1 and Chk2 and the DNA damage response. DNA Repair 8:1047–54 [Google Scholar]
  23. Weitzman MD, Lilley CE, Chaurushiya MS. 23.  2010. Genomes in conflict: maintaining genome integrity during virus infection. Annu. Rev. Microbiol. 64:61–81 [Google Scholar]
  24. Everett RD. 24.  2006. Interactions between DNA viruses, ND10 and the DNA damage response. Cell Microbiol. 8:365–74 [Google Scholar]
  25. Lilley CE, Chaurushiya MS, Boutell C, Everett RD, Weitzman MD. 25.  2011. The intrinsic antiviral defense to incoming HSV-1 genomes includes specific DNA repair proteins and is counteracted by the viral protein ICP0. PLoS Pathog. 7:e1002084Defined the viral ICP0 protein as a major ubiquitin ligase important for targeting the DDR to permit virus (HSV) replication. [Google Scholar]
  26. Kerur N, Veettil MV, Sharma-Walia N, Bottero V, Sadagopan S. 26.  et al. 2011. IFI16 acts as a nuclear pathogen sensor to induce the inflammasome in response to Kaposi sarcoma–associated herpesvirus infection. Cell Host Microbe 9:363–75 [Google Scholar]
  27. Barber GN. 27.  2014. STING-dependent cytosolic DNA sensing pathways. Trends Immunol. 35:88–93 [Google Scholar]
  28. Everett RD. 28.  2013. The spatial organization of DNA virus genomes in the nucleus. PLoS Pathog. 9:e1003386 [Google Scholar]
  29. Daniel R, Katz RA, Skalka AM. 29.  1999. A role for DNA-PK in retroviral DNA integration. Science 284:644–47 [Google Scholar]
  30. Schmid M, Speiseder T, Dobner T, Gonzalez RA. 30.  2014. DNA virus replication compartments. J. Virol. 881404–20
  31. Stracker TH, Carson CT, Weitzman MD. 31.  2002. Adenovirus oncoproteins inactivate the Mre11-Rad50-NBS1 DNA repair complex. Nature 418:348–52First identified DDR as an antiviral pathway in response to adenovirus infection; the E1B55k/E4orf6 ubiquitin ligase targets the MRN complex to prevent genome concatemerization. [Google Scholar]
  32. Carson CT, Schwartz RA, Stracker TH, Lilley CE, Lee DV, Weitzman MD. 32.  2003. The Mre11 complex is required for ATM activation and the G2/M checkpoint. EMBO J. 22:6610–20 [Google Scholar]
  33. Blackford AN, Bruton RK, Dirlik O, Stewart GS, Taylor AM. 33.  et al. 2008. A role for E1B-AP5 in ATR signaling pathways during adenovirus infection. J. Virol. 82:7640–52 [Google Scholar]
  34. Cervelli T, Palacios JA, Zentilin L, Mano M, Schwartz RA. 34.  et al. 2008. Processing of recombinant AAV genomes occurs in specific nuclear structures that overlap with foci of DNA-damage-response proteins. J. Cell Sci. 121:349–57 [Google Scholar]
  35. Collaco RF, Bevington JM, Bhrigu V, Kalman-Maltese V, Trempe JP. 35.  2009. Adeno-associated virus and adenovirus coinfection induces a cellular DNA damage and repair response via redundant phosphatidylinositol 3-like kinase pathways. Virology 392:24–33 [Google Scholar]
  36. Schwartz RA, Carson CT, Schuberth C, Weitzman MD. 36.  2009. Adeno-associated virus replication induces a DNA damage response coordinated by DNA-dependent protein kinase. J. Virol. 83:6269–78 [Google Scholar]
  37. Adeyemi RO, Landry S, Davis ME, Weitzman MD, Pintel DJ. 37.  2010. Parvovirus minute virus of mice induces a DNA damage response that facilitates viral replication. PLoS Pathog. 6:e1001141 [Google Scholar]
  38. Taylor TJ, Knipe DM. 38.  2004. Proteomics of herpes simplex virus replication compartments: association of cellular DNA replication, repair, recombination, and chromatin remodeling proteins with ICP8. J. Virol. 78:5856–66 [Google Scholar]
  39. Wilkinson DE, Weller SK. 39.  2004. Recruitment of cellular recombination and repair proteins to sites of herpes simplex virus type 1 DNA replication is dependent on the composition of viral proteins within prereplicative sites and correlates with the induction of the DNA damage response. J. Virol. 78:4783–96 [Google Scholar]
  40. Lilley CE, Carson CT, Muotri AR, Gage FH, Weitzman MD. 40.  2005. DNA repair proteins affect the lifecycle of herpes simplex virus 1. Proc. Natl. Acad. Sci. USA 102:5844–49 [Google Scholar]
  41. Shirata N, Kudoh A, Daikoku T, Tatsumi Y, Fujita M. 41.  et al. 2005. Activation of ataxia telangiectasia–mutated DNA damage checkpoint signal transduction elicited by herpes simplex virus infection. J. Biol. Chem. 280:30336–41 [Google Scholar]
  42. Xiaofei E, Pickering MT, Debatis M, Castillo J, Lagadinos A. 42.  et al. 2011. An E2F1-mediated DNA damage response contributes to the replication of human cytomegalovirus. PLoS Pathog. 7:e1001342 [Google Scholar]
  43. Kudoh A, Fujita M, Zhang L, Shirata N, Daikoku T. 43.  et al. 2005. Epstein–Barr virus lytic replication elicits ATM checkpoint signal transduction while providing an S-phase-like cellular environment. J. Biol. Chem. 280:8156–63 [Google Scholar]
  44. Daikoku T, Kudoh A, Sugaya Y, Iwahori S, Shirata N. 44.  et al. 2006. Postreplicative mismatch repair factors are recruited to Epstein–Barr virus replication compartments. J. Biol. Chem. 281:11422–30 [Google Scholar]
  45. Moody CA, Laimins LA. 45.  2009. Human papillomaviruses activate the ATM DNA damage pathway for viral genome amplification upon differentiation. PLoS Pathog. 5:e1000605 [Google Scholar]
  46. Hong S, Laimins LA. 46.  2013. The JAK-STAT transcriptional regulator, STAT-5, activates the ATM DNA damage pathway to induce HPV 31 genome amplification upon epithelial differentiation. PLoS Pathog. 9:e1003295 [Google Scholar]
  47. Shi Y, Dodson GE, Shaikh S, Rundell K, Tibbetts RS. 47.  2005. Ataxia-telangiectasia-mutated (ATM) is a T-antigen kinase that controls SV40 viral replication in vivo. J. Biol. Chem. 280:40195–200 [Google Scholar]
  48. Zhao X, Madden-Fuentes RJ, Lou BX, Pipas JM, Gerhardt J. 48.  et al. 2008. Ataxia telangiectasia–mutated damage-signaling kinase– and proteasome-dependent destruction of Mre11-Rad50-Nbs1 subunits in simian virus 40–infected primate cells. J. Virol. 82:5316–28 [Google Scholar]
  49. Dahl J, You J, Benjamin TL. 49.  2005. Induction and utilization of an ATM signaling pathway by polyomavirus. J. Virol. 79:13007–17 [Google Scholar]
  50. Lau A, Swinbank KM, Ahmed PS, Taylor DL, Jackson SP. 50.  et al. 2005. Suppression of HIV-1 infection by a small molecule inhibitor of the ATM kinase. Nat. Cell Biol. 7:493–500 [Google Scholar]
  51. Munger K, Werness BA, Dyson N, Phelps WC, Harlow E, Howley PM. 51.  1989. Complex formation of human papillomavirus E7 proteins with the retinoblastoma tumor suppressor gene product. EMBO J. 8:4099–105 [Google Scholar]
  52. Dyson N, Howley PM, Munger K, Harlow E. 52.  1989. The human papilloma virus-16 E7 oncoprotein is able to bind to the retinoblastoma gene product. Science 243:934–37 [Google Scholar]
  53. DeCaprio JA, Ludlow JW, Figge J, Shew JY, Huang CM. 53.  et al. 1988. SV40 large tumor antigen forms a specific complex with the product of the retinoblastoma susceptibility gene. Cell 54:275–83 [Google Scholar]
  54. Bester AC, Roniger M, Oren YS, Im MM, Sarni D. 54.  et al. 2011. Nucleotide deficiency promotes genomic instability in early stages of cancer development. Cell 145:435–46 [Google Scholar]
  55. Frame FM, Rogoff HA, Pickering MT, Cress WD, Kowalik TF. 55.  2006. E2F1 induces MRN foci formation and a cell cycle checkpoint response in human fibroblasts. Oncogene 25:3258–66 [Google Scholar]
  56. Pickering MT, Kowalik TF. 56.  2006. Rb inactivation leads to E2F1-mediated DNA double-strand break accumulation. Oncogene 25:746–55 [Google Scholar]
  57. Castillo JP, Frame FM, Rogoff HA, Pickering MT, Yurochko AD, Kowalik TF. 57.  2005. Human cytomegalovirus IE1-72 activates ataxia telangiectasia mutated kinase and a p53/p21-mediated growth arrest response. J. Virol. 79:11467–75 [Google Scholar]
  58. Koopal S, Furuhjelm JH, Jarviluoma A, Jaamaa S, Pyakurel P. 58.  et al. 2007. Viral oncogene-induced DNA damage response is activated in Kaposi sarcoma tumorigenesis. PLoS Pathog. 3:1348–6058 and 59: Defined the DDR as an innate tumor-suppressor response to oncogenic herpesvirus infection. [Google Scholar]
  59. Nikitin PA, Yan CM, Forte E, Bocedi A, Tourigny JP. 59.  et al. 2010. An ATM/Chk2-mediated DNA damage–responsive signaling pathway suppresses Epstein–Barr virus transformation of primary human B cells. Cell Host Microbe 8:510–2258 and 59: Defined the DDR as an innate tumor-suppressor response to oncogenic herpesvirus infection. [Google Scholar]
  60. Boichuk S, Hu L, Hein J, Gjoerup OV. 60.  2010. Multiple DNA damage signaling and repair pathways deregulated by simian virus 40 large T antigen. J. Virol. 84:8007–20 [Google Scholar]
  61. Hein J, Boichuk S, Wu J, Cheng Y, Freire R. 61.  et al. 2009. Simian virus 40 large T antigen disrupts genome integrity and activates a DNA damage response via Bub1 binding. J. Virol. 83:117–27 [Google Scholar]
  62. Gruhne B, Sompallae R, Masucci MG. 62.  2009. Three Epstein–Barr virus latency proteins independently promote genomic instability by inducing DNA damage, inhibiting DNA repair and inactivating cell cycle checkpoints. Oncogene 28:3997–4008 [Google Scholar]
  63. Spardy N, Duensing A, Hoskins EE, Wells SI, Duensing S. 63.  2008. HPV-16 E7 reveals a link between DNA replication stress, Fanconi anemia D2 protein, and alternative lengthening of telomere-associated promyelocytic leukemia bodies. Cancer Res. 68:9954–63 [Google Scholar]
  64. Duensing S, Duensing A, Crum CP, Munger K. 64.  2001. Human papillomavirus type 16 E7 oncoprotein-induced abnormal centrosome synthesis is an early event in the evolving malignant phenotype. Cancer Res. 61:2356–60 [Google Scholar]
  65. Liang X, Pickering MT, Cho NH, Chang H, Volkert MR. 65.  et al. 2006. Deregulation of DNA damage signal transduction by herpesvirus latency-associated M2. J. Virol. 80:5862–74 [Google Scholar]
  66. Belgnaoui SM, Fryrear KA, Nyalwidhe JO, Guo X, Semmes OJ. 66.  2010. The viral oncoprotein Tax sequesters DNA damage response factors by tethering MDC1 to chromatin. J. Biol. Chem. 285:32897–905 [Google Scholar]
  67. Durkin SS, Guo X, Fryrear KA, Mihaylova VT, Gupta SK. 67.  et al. 2008. HTLV-1 Tax oncoprotein subverts the cellular DNA damage response via binding to DNA-dependent protein kinase. J. Biol. Chem. 283:36311–20 [Google Scholar]
  68. Wu X, Avni D, Chiba T, Yan F, Zhao Q. 68.  et al. 2004. SV40 T antigen interacts with Nbs1 to disrupt DNA replication control. Genes Dev. 18:1305–16 [Google Scholar]
  69. Schrofelbauer B, Hakata Y, Landau NR. 69.  2007. HIV-1 Vpr function is mediated by interaction with the damage-specific DNA-binding protein DDB1. Proc. Natl. Acad. Sci. USA 104:4130–35 [Google Scholar]
  70. Gillespie KA, Mehta KP, Laimins LA, Moody CA. 70.  2012. Human papillomaviruses recruit cellular DNA repair and homologous recombination factors to viral replication centers. J. Virol. 86:9520–26 [Google Scholar]
  71. Sakakibara N, Mitra R, McBride AA. 71.  2011. The papillomavirus E1 helicase activates a cellular DNA damage response in viral replication foci. J. Virol. 85:8981–95 [Google Scholar]
  72. Fradet-Turcotte A, Bergeron-Labrecque F, Moody CA, Lehoux M, Laimins LA, Archambault J. 72.  2011. Nuclear accumulation of the papillomavirus E1 helicase blocks S-phase progression and triggers an ATM-dependent DNA damage response. J. Virol. 85:8996–9012 [Google Scholar]
  73. Gastaldello S, Hildebrand S, Faridani O, Callegari S, Palmkvist M. 73.  et al. 2010. A deneddylase encoded by Epstein–Barr virus promotes viral DNA replication by regulating the activity of cullin-RING ligases. Nat. Cell Biol. 12:351–61 [Google Scholar]
  74. Whitehurst CB, Vaziri C, Shackelford J, Pagano JS. 74.  2012. Epstein–Barr virus BPLF1 deubiquitinates PCNA and attenuates polymerase η recruitment to DNA damage sites. J. Virol. 86:8097–106 [Google Scholar]
  75. Hume AJ, Finkel JS, Kamil JP, Coen DM, Culbertson MR, Kalejta RF. 75.  2008. Phosphorylation of retinoblastoma protein by viral protein with cyclin-dependent kinase function. Science 320:797–99 [Google Scholar]
  76. Kuny CV, Chinchilla K, Culbertson MR, Kalejta RF. 76.  2010. Cyclin-dependent kinase–like function is shared by the beta- and gamma- subset of the conserved herpesvirus protein kinases. PLoS Pathog. 6:e1001092 [Google Scholar]
  77. Tarakanova VL, Leung-Pineda V, Hwang S, Yang CW, Matatall K. 77.  et al. 2007. γ-Herpesvirus kinase actively initiates a DNA damage response by inducing phosphorylation of H2AX to foster viral replication. Cell Host Microbe 1:275–86First study to define the host DDR as a target of a viral kinase, promoting replication of MHV68. [Google Scholar]
  78. Li R, Zhu J, Xie Z, Liao G, Liu J. 78.  et al. 2011. Conserved herpesvirus kinases target the DNA damage response pathway and TIP60 histone acetyltransferase to promote virus replication. Cell Host Microbe 10:390–400Proteomic identification of the DDR as a global target for the CHPKs. [Google Scholar]
  79. Gruhne B, Sompallae R, Marescotti D, Kamranvar SA, Gastaldello S, Masucci MG. 79.  2009. The Epstein–Barr virus nuclear antigen-1 promotes genomic instability via induction of reactive oxygen species. Proc. Natl. Acad. Sci. USA 106:2313–18 [Google Scholar]
  80. Kinjo T, Ham-Terhune J, Peloponese JM Jr, Jeang KT. 80.  2010. Induction of reactive oxygen species by human T-cell leukemia virus type 1 Tax correlates with DNA damage and expression of cellular senescence marker. J. Virol. 84:5431–37 [Google Scholar]
  81. Weiden MD, Ginsberg HS. 81.  1994. Deletion of the E4 region of the genome produces adenovirus DNA concatemers. Proc. Natl. Acad. Sci. USA 91:153–57 [Google Scholar]
  82. Li R, Hayward SD. 82.  2011. The Ying-Yang of the virus-host interaction: control of the DNA damage response. Future Microbiol. 6:379–83 [Google Scholar]
  83. Hagemeier SR, Barlow EA, Meng Q, Kenney SC. 83.  2012. The cellular ataxia telangiectasia–mutated kinase promotes Epstein–Barr virus lytic reactivation in response to multiple different types of lytic reactivation–inducing stimuli. J. Virol. 86:13360–70 [Google Scholar]
  84. Tarakanova VL, Stanitsa E, Leonardo SM, Bigley TM, Gauld SB. 84.  2010. Conserved gammaherpesvirus kinase and histone variant H2AX facilitate gammaherpesvirus latency in vivo. Virology 405:50–61 [Google Scholar]
  85. Choi YK, Nash K, Byrne BJ, Muzyczka N, Song S. 85.  2010. The effect of DNA-dependent protein kinase on adeno-associated virus replication. PLoS ONE 5:e15073 [Google Scholar]
  86. Luo Y, Chen AY, Qiu J. 86.  2011. Bocavirus infection induces a DNA damage response that facilitates viral DNA replication and mediates cell death. J. Virol. 85:133–45 [Google Scholar]
  87. Daniel R, Marusich E, Argyris E, Zhao RY, Skalka AM, Pomerantz RJ. 87.  2005. Caffeine inhibits human immunodeficiency virus type 1 transduction of nondividing cells. J. Virol. 79:2058–65 [Google Scholar]
  88. Ariumi Y, Turelli P, Masutani M, Trono D. 88.  2005. DNA damage sensors ATM, ATR, DNA-PKcs, and PARP-1 are dispensable for human immunodeficiency virus type 1 integration. J. Virol. 79:2973–78 [Google Scholar]
  89. Cooper A, Garcia M, Petrovas C, Yamamoto T, Koup RA, Nabel GJ. 89.  2013. HIV-1 causes CD4 cell death through DNA-dependent protein kinase during viral integration. Nature 498:376–79 [Google Scholar]
  90. Sowd GA, Li NY, Fanning E. 90.  2013. ATM and ATR activities maintain replication fork integrity during SV40 chromatin replication. PLoS Pathog. 9:e1003283 [Google Scholar]
  91. Mohni KN, Dee AR, Smith S, Schumacher AJ, Weller SK. 91.  2013. Efficient herpes simplex virus 1 replication requires cellular ATR pathway proteins. J. Virol. 87:531–42 [Google Scholar]
  92. Dheekollu J, Deng Z, Wiedmer A, Weitzman MD, Lieberman PM. 92.  2007. A role for MRE11, NBS1, and recombination junctions in replication and stable maintenance of EBV episomes. PLoS ONE 2:e1257 [Google Scholar]
  93. Baker A, Rohleder KJ, Hanakahi LA, Ketner G. 93.  2007. Adenovirus E4 34k and E1b 55k oncoproteins target host DNA ligase IV for proteasomal degradation. J. Virol. 81:7034–40 [Google Scholar]
  94. Orazio NI, Naeger CM, Karlseder J, Weitzman MD. 94.  2011. The adenovirus E1b55K/E4orf6 complex induces degradation of the Bloom helicase during infection. J. Virol. 85:1887–92 [Google Scholar]
  95. Blackford AN, Patel RN, Forrester NA, Theil K, Groitl P. 95.  et al. 2010. Adenovirus 12 E4orf6 inhibits ATR activation by promoting TOPBP1 degradation. Proc. Natl. Acad. Sci. USA 107:12251–56 [Google Scholar]
  96. Parkinson J, Lees-Miller SP, Everett RD. 96.  1999. Herpes simplex virus type 1 immediate-early protein vmw110 induces the proteasome-dependent degradation of the catalytic subunit of DNA-dependent protein kinase. J. Virol. 73:650–57 [Google Scholar]
  97. Chaurushiya MS, Lilley CE, Aslanian A, Meisenhelder J, Scott DC. 97.  et al. 2012. Viral E3 ubiquitin ligase–mediated degradation of a cellular E3: Viral mimicry of a cellular phosphorylation mark targets the RNF8 FHA domain. Mol. Cell 46:79–9097 and 98: Demonstrated viral evasion of the DDR by HSV ICP0 phosphopeptide mimicry of RNF8 binding to ATM-phosphorylated MDC1. [Google Scholar]
  98. Lilley CE, Chaurushiya MS, Boutell C, Landry S, Suh J. 98.  et al. 2010. A viral E3 ligase targets RNF8 and RNF168 to control histone ubiquitination and DNA damage responses. EMBO J. 29:943–5597 and 98: Demonstrated viral evasion of the DDR by HSV ICP0 phosphopeptide mimicry of RNF8 binding to ATM-phosphorylated MDC1. [Google Scholar]
  99. Mohni KN, Livingston CM, Cortez D, Weller SK. 99.  2010. ATR and ATRIP are recruited to herpes simplex virus type 1 replication compartments even though ATR signaling is disabled. J. Virol. 84:12152–64 [Google Scholar]
  100. Wilkinson DE, Weller SK. 100.  2006. Herpes simplex virus type I disrupts the ATR-dependent DNA-damage response during lytic infection. J. Cell Sci. 119:2695–703 [Google Scholar]
  101. Mohni KN, Smith S, Dee AR, Schumacher AJ, Weller SK. 101.  2013. Herpes simplex virus type 1 single strand DNA binding protein and helicase/primase complex disable cellular ATR signaling. PLoS Pathog. 9:e1003652Showed that the HSV helicase/primase complex mimics RPA/ATRIP recognition of ssDNA-dsDNA junctions and suppresses ATR signaling. [Google Scholar]
  102. Luo MH, Rosenke K, Czornak K, Fortunato EA. 102.  2007. Human cytomegalovirus disrupts both ataxia telangiectasia mutated protein (ATM)- and ATM-Rad3-related kinase–mediated DNA damage responses during lytic infection. J. Virol. 81:1934–50 [Google Scholar]
  103. O'Dowd JM, Zavala AG, Brown CJ, Mori T, Fortunato EA. 103.  2012. HCMV-infected cells maintain efficient nucleotide excision repair of the viral genome while abrogating repair of the host genome. PLoS Pathog. 8:e1003038 [Google Scholar]
  104. Carvalho T, Seeler JS, Ohman K, Jordan P, Pettersson U. 104.  et al. 1995. Targeting of adenovirus E1A and E4-ORF3 proteins to nuclear matrix–associated PML bodies. J. Cell Biol. 131:45–56 [Google Scholar]
  105. Doucas V, Ishov AM, Romo A, Juguilon H, Weitzman MD. 105.  et al. 1996. Adenovirus replication is coupled with the dynamic properties of the PML nuclear structure. Genes Dev. 10:196–207 [Google Scholar]
  106. Ou HD, Kwiatkowski W, Deerinck TJ, Noske A, Blain KY. 106.  et al. 2012. A structural basis for the assembly and functions of a viral polymer that inactivates multiple tumor suppressors. Cell 151:304–19Used the three-dimensional structure of adenovirus E4orf3 and innovative microscopy to define the mechanism of broad suppression of MRN and other intrinsic viral defenses. [Google Scholar]
  107. Liu Y, Shevchenko A, Shevchenko A, Berk AJ. 107.  2005. Adenovirus exploits the cellular aggresome response to accelerate inactivation of the MRN complex. J. Virol. 79:14004–16 [Google Scholar]
  108. Araujo FD, Stracker TH, Carson CT, Lee DV, Weitzman MD. 108.  2005. Adenovirus type 5 E4orf3 protein targets the Mre11 complex to cytoplasmic aggresomes. J. Virol. 79:11382–91 [Google Scholar]
  109. Carson CT, Orazio NI, Lee DV, Suh J, Bekker-Jensen S. 109.  et al. 2009. Mislocalization of the MRN complex prevents ATR signaling during adenovirus infection. EMBO J. 28:652–62 [Google Scholar]
  110. Shin YC, Nakamura H, Liang X, Feng P, Chang H. 110.  et al. 2006. Inhibition of the ATM/p53 signal transduction pathway by Kaposi's sarcoma–associated herpesvirus interferon regulatory factor 1. J. Virol. 80:2257–66 [Google Scholar]
  111. Lai CK, Jeng KS, Machida K, Cheng YS, Lai MM. 111.  2008. Hepatitis C virus NS3/4A protein interacts with ATM, impairs DNA repair and enhances sensitivity to ionizing radiation. Virology 370:295–309 [Google Scholar]
  112. Machida K, McNamara G, Cheng KT, Huang J, Wang CH. 112.  et al. 2010. Hepatitis C virus inhibits DNA damage repair through reactive oxygen and nitrogen species and by interfering with the ATM-NBS1/Mre11/Rad50 DNA repair pathway in monocytes and hepatocytes. J. Immunol. 185:6985–98 [Google Scholar]
  113. Querido E, Blanchette P, Yan Q, Kamura T, Morrison M. 113.  et al. 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]
  114. Scheffner M, Werness BA, Huibregtse JM, Levine AJ, Howley PM. 114.  1990. The E6 oncoprotein encoded by human papillomavirus types 16 and 18 promotes the degradation of p53. Cell 63:1129–36 [Google Scholar]
  115. Sato Y, Kamura T, Shirata N, Murata T, Kudoh A. 115.  et al. 2009. Degradation of phosphorylated p53 by viral protein–ECS E3 ligase complex. PLoS Pathog. 5:e1000530 [Google Scholar]
  116. Muller S, Dobner T. 116.  2008. The adenovirus E1B-55K oncoprotein induces SUMO modification of p53. Cell Cycle 7:754–58 [Google Scholar]
  117. Pennella MA, Liu Y, Woo JL, Kim CA, Berk AJ. 117.  2010. Adenovirus E1B 55-kilodalton protein is a p53-SUMO1 E3 ligase that represses p53 and stimulates its nuclear export through interactions with promyelocytic leukemia nuclear bodies. J. Virol. 84:12210–25 [Google Scholar]
  118. Soria C, Estermann FE, Espantman KC, O'Shea CC. 118.  2010. Heterochromatin silencing of p53 target genes by a small viral protein. Nature 466:1076–81 [Google Scholar]
  119. Friborg J Jr, Kong W, Hottiger MO, Nabel GJ. 119.  1999. p53 inhibition by the LANA protein of KSHV protects against cell death. Nature 402:889–94 [Google Scholar]
  120. Chen W, Hilton IB, Staudt MR, Burd CE, Dittmer DP. 120.  2010. Distinct p53, p53:LANA, and LANA complexes in Kaposi's sarcoma–associated herpesvirus lymphomas. J. Virol. 84:3898–908 [Google Scholar]
  121. Yi F, Saha A, Murakami M, Kumar P, Knight JS. 121.  et al. 2009. Epstein–Barr virus nuclear antigen 3C targets p53 and modulates its transcriptional and apoptotic activities. Virology 388:236–47 [Google Scholar]
  122. Wang XW, Forrester K, Yeh H, Feitelson MA, Gu JR, Harris CC. 122.  1994. Hepatitis B virus X protein inhibits p53 sequence-specific DNA binding, transcriptional activity, and association with transcription factor ERCC3. Proc. Natl. Acad. Sci. USA 91:2230–34 [Google Scholar]
  123. Ariumi Y, Kaida A, Lin JY, Hirota M, Masui O. 123.  et al. 2000. HTLV-1 Tax oncoprotein represses the p53-mediated trans-activation function through coactivator CBP sequestration. Oncogene 19:1491–99 [Google Scholar]
  124. Kaida A, Ariumi Y, Ueda Y, Lin JY, Hijikata M. 124.  et al. 2000. Functional impairment of p73 and p51, the p53-related proteins, by the human T-cell leukemia virus type 1 Tax oncoprotein. Oncogene 19:827–30 [Google Scholar]
  125. Varon R, Vissinga C, Platzer M, Cerosaletti KM, Chrzanowska KH. 125.  et al. 1998. Nibrin, a novel DNA double-strand break repair protein, is mutated in Nijmegen breakage syndrome. Cell 93:467–76 [Google Scholar]
  126. Vorechovsky I, Luo L, Lindblom A, Negrini M, Webster AD. 126.  et al. 1996. ATM mutations in cancer families. Cancer Res. 56:4130–33 [Google Scholar]
  127. Meijers-Heijboer H, van den Ouweland A, Klijn J, Wasielewski M, de Snoo A. 127.  et al. 2002. Low-penetrance susceptibility to breast cancer due to CHEK2*1100delC in noncarriers of BRCA1 or BRCA2 mutations. Nat. Genet. 31:55–59 [Google Scholar]
  128. Hangaishi A, Ogawa S, Qiao Y, Wang L, Hosoya N. 128.  et al. 2002. Mutations of Chk2 in primary hematopoietic neoplasms. Blood 99:3075–77 [Google Scholar]
  129. Leidal AM, Cyr DP, Hill RJ, Lee PW, McCormick C. 129.  2012. Subversion of autophagy by Kaposi's sarcoma–associated herpesvirus impairs oncogene-induced senescence. Cell Host Microbe 11:167–80Described suppression of autophagy by a viral protein as the mechanism for overcoming DDR-mediated senescence induced by KSHV. [Google Scholar]
  130. Skalska L, White RE, Parker GA, Sinclair AJ, Paschos K, Allday MJ. 130.  2013. Induction of p16INK4a is the major barrier to proliferation when Epstein–Barr virus (EBV) transforms primary B cells into lymphoblastoid cell lines. PLoS Pathog. 9:e1003187 [Google Scholar]
  131. Ishov AM, Maul GG. 131.  1996. The periphery of nuclear domain 10 (ND10) as site of DNA virus deposition. J. Cell Biol. 134:815–26 [Google Scholar]
  132. Everett RD, Murray J. 132.  2005. ND10 components relocate to sites associated with herpes simplex virus type 1 nucleoprotein complexes during virus infection. J. Virol. 79:5078–89 [Google Scholar]
  133. Maul GG, Ishov AM, Everett RD. 133.  1996. Nuclear domain 10 as preexisting potential replication start sites of herpes simplex virus type-1. Virology 217:67–75 [Google Scholar]
  134. Carbone R, Pearson M, Minucci S, Pelicci PG. 134.  2002. PML NBs associate with the hMre11 complex and p53 at sites of irradiation induced DNA damage. Oncogene 21:1633–40 [Google Scholar]
  135. Rodier F, Munoz DP, Teachenor R, Chu V, Le O. 135.  et al. 2011. DNA-SCARS: distinct nuclear structures that sustain damage-induced senescence growth arrest and inflammatory cytokine secretion. J. Cell Sci. 124:68–81 [Google Scholar]
  136. Cuchet-Lourenco D, Boutell C, Lukashchuk V, Grant K, Sykes A. 136.  et al. 2011. SUMO pathway dependent recruitment of cellular repressors to herpes simplex virus type 1 genomes. PLoS Pathog. 7:e1002123 [Google Scholar]
  137. Jackson SP, Durocher D. 137.  2013. Regulation of DNA damage responses by ubiquitin and SUMO. Mol. Cell 49:795–807 [Google Scholar]
  138. Hwang J, Kalejta RF. 138.  2007. Proteasome-dependent, ubiquitin-independent degradation of Daxx by the viral pp71 protein in human cytomegalovirus–infected cells. Virology 367:334–38 [Google Scholar]
  139. Tsai K, Thikmyanova N, Wojcechowskyj JA, Delecluse HJ, Lieberman PM. 139.  2011. EBV tegument protein BNRF1 disrupts DAXX-ATRX to activate viral early gene transcription. PLoS Pathog. 7:e1002376 [Google Scholar]
  140. Everett RD, Freemont P, Saitoh H, Dasso M, Orr A. 140.  et al. 1998. The disruption of ND10 during herpes simplex virus infection correlates with the Vmw110- and proteasome-dependent loss of several PML isoforms. J. Virol. 72:6581–91 [Google Scholar]
  141. Weitzman MD, Lilley CE, Chaurushiya MS. 141.  2011. Changing the ubiquitin landscape during viral manipulation of the DNA damage response. FEBS Lett. 585:2897–906 [Google Scholar]
  142. Refsland EW, Harris RS. 142.  2013. The APOBEC3 family of retroelement restriction factors. Curr. Top. Microbiol. Immunol. 371:1–27 [Google Scholar]
  143. Sheehy AM, Gaddis NC, Choi JD, Malim MH. 143.  2002. Isolation of a human gene that inhibits HIV-1 infection and is suppressed by the viral Vif protein. Nature 418:646–50 [Google Scholar]
  144. Vartanian JP, Henry M, Marchio A, Suspene R, Aynaud MM. 144.  et al. 2010. Massive APOBEC3 editing of hepatitis B viral DNA in cirrhosis. PLoS Pathog. 6:e1000928 [Google Scholar]
  145. Suspene R, Aynaud MM, Koch S, Pasdeloup D, Labetoulle M. 145.  et al. 2011. Genetic editing of herpes simplex virus 1 and Epstein–Barr herpesvirus genomes by human APOBEC3 cytidine deaminases in culture and in vivo. J. Virol. 85:7594–602 [Google Scholar]
  146. Lawrence MS, Stojanov P, Polak P, Kryukov GV, Cibulskis K. 146.  et al. 2013. Mutational heterogeneity in cancer and the search for new cancer-associated genes. Nature 499:214–18 [Google Scholar]
  147. Burns MB, Lackey L, Carpenter MA, Rathore A, Land AM. 147.  et al. 2013. APOBEC3B is an enzymatic source of mutation in breast cancer. Nature 494:366–70 [Google Scholar]
/content/journals/10.1146/annurev-virology-031413-085548
Loading
/content/journals/10.1146/annurev-virology-031413-085548
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