1932

Abstract

Our genome is a historic record of successive invasions of mobile genetic elements. Like other eukaryotes, we have evolved mechanisms to limit their propagation and minimize the functional impact of new insertions. Although these mechanisms are vitally important, they are imperfect, and a handful of retroelement families remain active in modern humans. This review introduces the intrinsic functions of transposons, the tactics employed in their restraint, and the relevance of this conflict to human pathology. The most straightforward examples of disease-causing transposable elements are germline insertions that disrupt a gene and result in a monogenic disease allele. More enigmatic are the abnormal patterns of transposable element expression in disease states. Changes in transposon regulation and cellular responses to their expression have implicated these sequences in diseases as diverse as cancer, autoimmunity, and neurodegeneration. Distinguishing their epiphenomenal from their pathogenic effects may provide wholly new perspectives on our understanding of disease.

Loading

Article metrics loading...

/content/journals/10.1146/annurev-pathmechdis-012419-032633
2020-01-24
2024-04-25
Loading full text...

Full text loading...

/deliver/fulltext/pathol/15/1/annurev-pathmechdis-012419-032633.html?itemId=/content/journals/10.1146/annurev-pathmechdis-012419-032633&mimeType=html&fmt=ahah

Literature Cited

  1. 1. 
    Huang CR, Burns KH, Boeke JD 2012. Active transposition in genomes. Annu. Rev. Genet. 46:651–75
    [Google Scholar]
  2. 2. 
    Burns KH, Boeke JD. 2012. Human transposon tectonics. Cell 149:740–52
    [Google Scholar]
  3. 3. 
    Jurka J, Kapitonov VV, Pavlicek A, Klonowski P, Kohany O, Walichiewicz J 2005. Repbase Update, a database of eukaryotic repetitive elements. Cytogenet. Genome Res. 110:462–67
    [Google Scholar]
  4. 4. 
    Smit AFA, Hubley R, Green P 2015. RepeatMasker Open–4.0. Institute for Systems Biology http://www.repeatmasker.org
    [Google Scholar]
  5. 5. 
    Lander ES, Linton LM, Birren B, Nusbaum C, Zody MC et al. 2001. Initial sequencing and analysis of the human genome. Nature 409:860–921
    [Google Scholar]
  6. 6. 
    Batzer MA, Deininger PL. 2002. Alu repeats and human genomic diversity. Nat. Rev. Genet. 3:370–79
    [Google Scholar]
  7. 7. 
    Jurka J, Zietkiewicz E, Labuda D 1995. Ubiquitous mammalian-wide interspersed repeats (MIRs) are molecular fossils from the Mesozoic era. Nucleic Acids Res 23:170–75
    [Google Scholar]
  8. 8. 
    Smit AFA, Riggs AD. 1995. MIRs are classic, tRNA-derived SINEs that amplified before the mammalian radiation. Nucleic Acids Res 23:98–102
    [Google Scholar]
  9. 9. 
    Konkel MK, Walker JA, Hotard AB, Ranck MC, Fontenot CC et al. 2015. Sequence analysis and characterization of active human Alu subfamilies based on the 1000 Genomes Pilot Project. Genome Biol. Evol. 7:2608–22
    [Google Scholar]
  10. 10. 
    Kryatova MS, Steranka JP, Burns KH, Payer LM 2017. Insertion and deletion polymorphisms of the ancient AluS family in the human genome. Mob. DNA 8:6
    [Google Scholar]
  11. 11. 
    Kapitonov V, Jurka J. 1996. The age of Alu subfamilies. J. Mol. Evol. 42:59–65
    [Google Scholar]
  12. 12. 
    Shen MR, Batzer MA, Deininger PL 1991. Evolution of the master Alu gene(s). J. Mol. Evol. 33:311–20
    [Google Scholar]
  13. 13. 
    Konkel MK, Walker JA, Batzer MA 2010. LINEs and SINEs of primate evolution. Evol. Anthropol. 19:236–49
    [Google Scholar]
  14. 14. 
    Dewannieux M, Esnault C, Heidmann T 2003. LINE-mediated retrotransposition of marked Alu sequences. Nat. Genet. 35:41–48
    [Google Scholar]
  15. 15. 
    Boissinot S, Chevret P, Furano AV 2000. L1 (LINE-1) retrotransposon evolution and amplification in recent human history. Mol. Biol. Evol. 17:915–28
    [Google Scholar]
  16. 16. 
    Boissinot S, Entezam A, Young L, Munson PJ, Furano AV 2004. The insertional history of an active family of L1 retrotransposons in humans. Genome Res 14:1221–31
    [Google Scholar]
  17. 17. 
    Boissinot S, Furano AV. 2005. The recent evolution of human L1 retrotransposons. Cytogenet. Genome Res. 110:402–6
    [Google Scholar]
  18. 18. 
    Beck CR, Garcia-Perez JL, Badge RM, Moran JV 2011. LINE-1 elements in structural variation and disease. Annu. Rev. Genom. Hum. Genet. 12:187–215
    [Google Scholar]
  19. 19. 
    Brouha B, Schustak J, Badge RM, Lutz-Prigge S, Farley AH et al. 2003. Hot L1s account for the bulk of retrotransposition in the human population. PNAS 100:5280–85
    [Google Scholar]
  20. 20. 
    Beck CR, Collier P, Macfarlane C, Malig M, Kidd JM et al. 2010. LINE-1 retrotransposition activity in human genomes. Cell 141:1159–70
    [Google Scholar]
  21. 21. 
    Raiz J, Damert A, Chira S, Held U, Klawitter S et al. 2012. The non-autonomous retrotransposon SVA is trans-mobilized by the human LINE-1 protein machinery. Nucleic Acids Res 40:1666–83
    [Google Scholar]
  22. 22. 
    Hancks DC, Goodier JL, Mandal PK, Cheung LE, Kazazian HH Jr 2011. Retrotransposition of marked SVA elements by human L1s in cultured cells. Hum. Mol. Genet. 20:3386–400
    [Google Scholar]
  23. 23. 
    Huang CR, Schneider AM, Lu Y, Niranjan T, Shen P et al. 2010. Mobile interspersed repeats are major structural variants in the human genome. Cell 141:1171–82
    [Google Scholar]
  24. 24. 
    Ewing AD, Kazazian HH Jr 2010. High-throughput sequencing reveals extensive variation in human-specific L1 content in individual human genomes. Genome Res 20:1262–70
    [Google Scholar]
  25. 25. 
    Iskow RC, McCabe MT, Mills RE, Torene S, Pittard WS et al. 2010. Natural mutagenesis of human genomes by endogenous retrotransposons. Cell 141:1253–61
    [Google Scholar]
  26. 26. 
    Stewart C, Kural D, Stromberg MP, Walker JA, Konkel MK et al. 2011. A comprehensive map of mobile element insertion polymorphisms in humans. PLOS Genet 7:e1002236
    [Google Scholar]
  27. 27. 
    Sudmant PH, Rausch T, Gardner EJ, Handsaker RE, Abyzov A et al. 2015. An integrated map of structural variation in 2,504 human genomes. Nature 526:75–81
    [Google Scholar]
  28. 28. 
    Dewannieux M, Harper F, Richaud A, Letzelter C, Ribet D et al. 2006. Identification of an infectious progenitor for the multiple-copy HERV-K human endogenous retroelements. Genome Res 16:1548–56
    [Google Scholar]
  29. 29. 
    Belshaw R, Dawson AL, Woolven-Allen J, Redding J, Burt A, Tristem M 2005. Genomewide screening reveals high levels of insertional polymorphism in the human endogenous retrovirus family HERV-K(HML2): implications for present-day activity. J. Virol. 79:12507–14
    [Google Scholar]
  30. 30. 
    Thomas J, Perron H, Feschotte C 2018. Variation in proviral content among human genomes mediated by LTR recombination. Mob. DNA 9:36
    [Google Scholar]
  31. 31. 
    Skowronski J, Fanning TG, Singer MF 1988. Unit-length line-1 transcripts in human teratocarcinoma cells. Mol. Cell. Biol. 8:1385–97
    [Google Scholar]
  32. 32. 
    Deininger P, Morales ME, White TB, Baddoo M, Hedges DJ et al. 2016. A comprehensive approach to expression of L1 loci. Nucleic Acids Res 45:e31
    [Google Scholar]
  33. 33. 
    Hall LL, Carone DM, Gomez AV, Kolpa HJ, Byron M et al. 2014. Stable C0T-1 repeat RNA is abundant and is associated with euchromatic interphase chromosomes. Cell 156:907–19
    [Google Scholar]
  34. 34. 
    Rodic N, Sharma R, Sharma R, Zampella J, Dai L et al. 2014. Long interspersed element-1 protein expression is a hallmark of many human cancers. Am. J. Pathol. 184:1280–86
    [Google Scholar]
  35. 35. 
    Sharma R, Rodic N, Burns KH, Taylor MS 2016. Immunodetection of human LINE-1 expression in cultured cells and human tissues. Methods Mol. Biol. 1400:261–80
    [Google Scholar]
  36. 36. 
    Ardeljan D, Taylor MS, Ting DT, Burns KH 2017. The human long interspersed element-1 retrotransposon: an emerging biomarker of neoplasia. Clin. Chem. 63:816–22
    [Google Scholar]
  37. 37. 
    Martin SL, Cruceanu M, Branciforte D, Wai-Lun Li P, Kwok SC et al. 2005. LINE-1 retrotransposition requires the nucleic acid chaperone activity of the ORF1 protein. J. Mol. Biol. 348:549–61
    [Google Scholar]
  38. 38. 
    Khazina E, Truffault V, Buttner R, Schmidt S, Coles M, Weichenrieder O 2011. Trimeric structure and flexibility of the L1ORF1 protein in human L1 retrotransposition. Nat. Struct. Mol. Biol. 18:1006–14
    [Google Scholar]
  39. 39. 
    Naufer MN, Callahan KE, Cook PR, Perez-Gonzalez CE, Williams MC, Furano AV 2016. L1 retrotransposition requires rapid ORF1p oligomerization, a novel coiled coil–dependent property conserved despite extensive remodeling. Nucleic Acids Res 44:281–93
    [Google Scholar]
  40. 40. 
    Khazina E, Weichenrieder O. 2018. Human LINE-1 retrotransposition requires a metastable coiled coil and a positively charged N-terminus in L1ORF1p. eLife 7:e34960
    [Google Scholar]
  41. 41. 
    Sokolowski M, Chynces M, deHaro D, Christian CM, Belancio VP 2017. Truncated ORF1 proteins can suppress LINE-1 retrotransposition in trans. Nucleic Acids Res 45:5294–308
    [Google Scholar]
  42. 42. 
    Kolosha VO, Martin SL. 2003. High-affinity, non-sequence-specific RNA binding by the open reading frame 1 (ORF1) protein from long interspersed nuclear element 1 (LINE-1). J. Biol. Chem. 278:8112–17
    [Google Scholar]
  43. 43. 
    Feng Q, Moran JV, Kazazian HH Jr., Boeke JD 1996. Human L1 retrotransposon encodes a conserved endonuclease required for retrotransposition. Cell 87:905–16
    [Google Scholar]
  44. 44. 
    Weichenrieder O, Repanas K, Perrakis A 2004. Crystal structure of the targeting endonuclease of the human LINE-1 retrotransposon. Structure 12:975–86
    [Google Scholar]
  45. 45. 
    Mathias SL, Scott AF, Kazazian HH Jr., Boeke JD, Gabriel A. 1991. Reverse transcriptase encoded by a human transposable element. Science 254:1808–10
    [Google Scholar]
  46. 46. 
    Kopera HC, Moldovan JB, Morrish TA, Garcia-Perez JL, Moran JV 2011. Similarities between long interspersed element-1 (LINE-1) reverse transcriptase and telomerase. PNAS 108:20345–50
    [Google Scholar]
  47. 47. 
    Luan DD, Korman MH, Jakubczak JL, Eickbush TH 1993. Reverse transcription of R2Bm RNA is primed by a nick at the chromosomal target site: a mechanism for non-LTR retrotransposition. Cell 72:595–605
    [Google Scholar]
  48. 48. 
    Ostertag EM, Kazazian HH Jr 2001. Twin priming: a proposed mechanism for the creation of inversions in L1 retrotransposition. Genome Res 11:2059–65
    [Google Scholar]
  49. 49. 
    Moran JV, DeBerardinis RJ, Kazazian HH 1999. Exon shuffling by L1 retrotransposition. Science 283:1530–34
    [Google Scholar]
  50. 50. 
    Wei W, Gilbert N, Ooi SL, Lawler JF, Ostertag EM et al. 2001. Human L1 retrotransposition: cis preference versus trans complementation. Mol. Cell. Biol. 21:1429–39
    [Google Scholar]
  51. 51. 
    Kulpa DA, Moran JV. 2006. Cis-preferential LINE-1 reverse transcriptase activity in ribonucleoprotein particles. Nat. Struct. Mol. Biol. 13:655–60
    [Google Scholar]
  52. 52. 
    Weichenrieder O, Wild K, Strub K, Cusack S 2000. Structure and assembly of the Alu domain of the mammalian signal recognition particle. Nature 408:167–73
    [Google Scholar]
  53. 53. 
    Ahl V, Keller H, Schmidt S, Weichenrieder O 2015. Retrotransposition and crystal structure of an Alu RNP in the ribosome-stalling conformation. Mol. Cell 60:715–27
    [Google Scholar]
  54. 54. 
    Xing J, Zhang Y, Han K, Salem AH, Sen SK et al. 2009. Mobile elements create structural variation: analysis of a complete human genome. Genome Res 19:1516–26
    [Google Scholar]
  55. 55. 
    Esnault C, Maestre J, Heidmann T 2000. Human LINE retrotransposons generate processed pseudogenes. Nat. Genet. 24:363–67
    [Google Scholar]
  56. 56. 
    Buzdin A, Ustyugova S, Gogvadze E, Vinogradova T, Lebedev Y, Sverdlov E 2002. A new family of chimeric retrotranscripts formed by a full copy of U6 small nuclear RNA fused to the 3′ terminus of L1. Genomics 80:402–6
    [Google Scholar]
  57. 57. 
    Doucet AJ, Droc G, Siol O, Audoux J, Gilbert N 2015. U6 snRNA pseudogenes: markers of retrotransposition dynamics in mammals. Mol. Biol. Evol. 32:1815–32
    [Google Scholar]
  58. 58. 
    Boissinot S, Davis J, Entezam A, Petrov D, Furano AV 2006. Fitness cost of LINE-1 (L1) activity in humans. PNAS 103:9590–94
    [Google Scholar]
  59. 59. 
    Aravin AA, Hannon GJ, Brennecke J 2007. The Piwi–piRNA pathway provides an adaptive defense in the transposon arms race. Science 318:761–64
    [Google Scholar]
  60. 60. 
    Newkirk SJ, Lee S, Grandi FC, Gaysinskaya V, Rosser JM et al. 2017. Intact piRNA pathway prevents L1 mobilization in male meiosis. PNAS 114:E5635–44
    [Google Scholar]
  61. 61. 
    Jacobs FM, Greenberg D, Nguyen N, Haeussler M, Ewing AD et al. 2014. An evolutionary arms race between KRAB zinc-finger genes ZNF91/93 and SVA/L1 retrotransposons. Nature 516:242–45
    [Google Scholar]
  62. 62. 
    Quenneville S, Turelli P, Bojkowska K, Raclot C, Offner S et al. 2012. The KRAB-ZFP/KAP1 system contributes to the early embryonic establishment of site-specific DNA methylation patterns maintained during development. Cell Rep 2:766–73
    [Google Scholar]
  63. 63. 
    Rowe HM, Jakobsson J, Mesnard D, Rougemont J, Reynard S et al. 2010. KAP1 controls endogenous retroviruses in embryonic stem cells. Nature 463:237–40
    [Google Scholar]
  64. 64. 
    Robbez-Masson L, Tie CHC, Conde L, Tunbak H, Husovsky C et al. 2018. The HUSH complex cooperates with TRIM28 to repress young retrotransposons and new genes. Genome Res 28:836–45
    [Google Scholar]
  65. 65. 
    Liu N, Lee CH, Swigut T, Grow E, Gu B et al. 2018. Selective silencing of euchromatic L1s revealed by genome-wide screens for L1 regulators. Nature 553:228–32
    [Google Scholar]
  66. 66. 
    Ohtani H, Liu M, Zhou W, Liang G, Jones PA 2018. Switching roles for DNA and histone methylation depend on evolutionary ages of human endogenous retroviruses. Genome Res 28:1147–57
    [Google Scholar]
  67. 67. 
    Walter M, Teissandier A, Perez-Palacios R, Bourc'his D 2016. An epigenetic switch ensures transposon repression upon dynamic loss of DNA methylation in embryonic stem cells. eLife 5:e11418
    [Google Scholar]
  68. 68. 
    Scott EC, Gardner EJ, Masood A, Chuang NT, Vertino PM, Devine SE 2016. A hot L1 retrotransposon evades somatic repression and initiates human colorectal cancer. Genome Res 26:745–55
    [Google Scholar]
  69. 69. 
    Chuong EB, Elde NC, Feschotte C 2016. Regulatory evolution of innate immunity through co-option of endogenous retroviruses. Science 351:1083–87
    [Google Scholar]
  70. 70. 
    Chuong EB, Elde NC, Feschotte C 2017. Regulatory activities of transposable elements: from conflicts to benefits. Nat. Rev. Genet. 18:71–86
    [Google Scholar]
  71. 71. 
    Macfarlan TS, Gifford WD, Driscoll S, Lettieri K, Rowe HM et al. 2012. Embryonic stem cell potency fluctuates with endogenous retrovirus activity. Nature 487:57–63
    [Google Scholar]
  72. 72. 
    Turelli P, Castro-Diaz N, Marzetta F, Kapopoulou A, Raclot C et al. 2014. Interplay of TRIM28 and DNA methylation in controlling human endogenous retroelements. Genome Res 24:1260–70
    [Google Scholar]
  73. 73. 
    Ecco G, Cassano M, Kauzlaric A, Duc J, Coluccio A et al. 2016. Transposable elements and their KRAB–ZFP controllers regulate gene expression in adult tissues. Dev. Cell 36:611–23
    [Google Scholar]
  74. 74. 
    Imbeault M, Helleboid PY, Trono D 2017. KRAB zinc-finger proteins contribute to the evolution of gene regulatory networks. Nature 543:550–54
    [Google Scholar]
  75. 75. 
    Agrawal A, Eastman QM, Schatz DG 1998. Transposition mediated by RAG1 and RAG2 and its implications for the evolution of the immune system. Nature 394:744–51
    [Google Scholar]
  76. 76. 
    Hencken CG, Li X, Craig NL 2012. Functional characterization of an active Rag-like transposase. Nat. Struct. Mol. Biol. 19:834–36
    [Google Scholar]
  77. 77. 
    Hiom K, Melek M, Gellert M 1998. DNA transposition by the RAG1 and RAG2 proteins: a possible source of oncogenic translocations. Cell 94:463–70
    [Google Scholar]
  78. 78. 
    Huang S, Tao X, Yuan S, Zhang Y, Li P et al. 2016. Discovery of an active RAG transposon illuminates the origins of V(D)J recombination. Cell 166:102–14
    [Google Scholar]
  79. 79. 
    Kapitonov VV, Jurka J. 2005. RAG1 core and V(D)J recombination signal sequences were derived from Transib transposons. PLOS Biol 3:e181
    [Google Scholar]
  80. 80. 
    Sinzelle L, Izsvak Z, Ivics Z 2009. Molecular domestication of transposable elements: from detrimental parasites to useful host genes. Cell. Mol. Life Sci. 66:1073–93
    [Google Scholar]
  81. 81. 
    Smit AF, Riggs AD. 1996. Tiggers and DNA transposon fossils in the human genome. PNAS 93:1443–48
    [Google Scholar]
  82. 82. 
    Sarkar A, Sim C, Hong YS, Hogan JR, Fraser MJ et al. 2003. Molecular evolutionary analysis of the widespread piggyBac transposon family and related “domesticated” sequences. Mol. Genet. Genom. 270:173–80
    [Google Scholar]
  83. 83. 
    Deciphering Dev. Disord. Study 2015. Large-scale discovery of novel genetic causes of developmental disorders. Nature 519:223–28
    [Google Scholar]
  84. 84. 
    Stessman HAF, Willemsen MH, Fenckova M, Penn O, Hoischen A et al. 2016. Disruption of POGZ is associated with intellectual disability and autism spectrum disorders. Am. J. Hum. Genet. 98:541–52
    [Google Scholar]
  85. 85. 
    Fuentes DR, Swigut T, Wysocka J 2018. Systematic perturbation of retroviral LTRs reveals widespread long-range effects on human gene regulation. eLife 7:e35989
    [Google Scholar]
  86. 86. 
    Kobayashi K, Nakahori Y, Miyake M, Matsumura K, Kondo-Iida E et al. 1998. An ancient retrotransposal insertion causes Fukuyama-type congenital muscular dystrophy. Nature 394:388–92
    [Google Scholar]
  87. 87. 
    Narita N, Nishio H, Kitoh Y, Ishikawa Y, Ishikawa Y et al. 1993. Insertion of a 5′ truncated L1 element into the 3′ end of exon 44 of the dystrophin gene resulted in skipping of the exon during splicing in a case of Duchenne muscular dystrophy. J. Clin. Investig. 91:1862–67
    [Google Scholar]
  88. 88. 
    Taniguchi-Ikeda M, Kobayashi K, Kanagawa M, Yu CC, Mori K et al. 2011. Pathogenic exon-trapping by SVA retrotransposon and rescue in Fukuyama muscular dystrophy. Nature 478:127–31
    [Google Scholar]
  89. 89. 
    Hancks DC, Kazazian HH Jr. 2016. Roles for retrotransposon insertions in human disease. Mob. DNA 7:9
    [Google Scholar]
  90. 90. 
    Kazazian HH Jr., Wong C, Youssoufian H, Scott AF, Phillips DG, Antonarakis SE. 1988. Haemophilia A resulting from de novo insertion of L1 sequences represents a novel mechanism for mutation in man. Nature 332:164–66
    [Google Scholar]
  91. 91. 
    Hancks DC, Ewing AD, Chen JE, Tokunaga K, Kazazian HH Jr 2009. Exon-trapping mediated by the human retrotransposon SVA. Genome Res 19:1983–91
    [Google Scholar]
  92. 92. 
    de Boer M, van Leeuwen K, Geissler J, Weemaes CM, van den Berg TK et al. 2014. Primary immunodeficiency caused by an exonized retroposed gene copy inserted in the CYBB gene. Hum. Mutat. 35:486–96
    [Google Scholar]
  93. 93. 
    Kopera HC, Larson PA, Moldovan JB, Richardson SR, Liu Y, Moran JV 2016. LINE-1 cultured cell retrotransposition assay. Methods Mol. Biol. 1400:139–56
    [Google Scholar]
  94. 94. 
    Payer LM, Steranka JP, Yang WR, Kryatova M, Medabalimi S et al. 2017. Structural variants caused by Alu insertions are associated with risks for many human diseases. PNAS 114:E3984–92
    [Google Scholar]
  95. 95. 
    Int. Mult. Scler. Genet. Consort 2013. Analysis of immune-related loci identifies 48 new susceptibility variants for multiple sclerosis. Nat. Genet 45:1353–60
    [Google Scholar]
  96. 96. 
    Int. Mult. Scler. Genet. Consort, Wellcome Trust Case Control Consort 2011. Genetic risk and a primary role for cell-mediated immune mechanisms in multiple sclerosis. Nature 476:214–19
    [Google Scholar]
  97. 97. 
    De Jager PL, Baecher-Allan C, Maier LM, Arthur AT, Ottoboni L et al. 2009. The role of the CD58 locus in multiple sclerosis. PNAS 106:5264–69
    [Google Scholar]
  98. 98. 
    Payer LM, Steranka JP, Ardeljan D, Walker J, Fitzgerald KC et al. 2019. Alu insertion variants alter mRNA splicing. Nucleic Acids Res 47:421–31
    [Google Scholar]
  99. 99. 
    Makino S, Kaji R, Ando S, Tomizawa M, Yasuno K et al. 2007. Reduced neuron-specific expression of the TAF1 gene is associated with X-linked dystonia–parkinsonism. Am. J. Hum. Genet. 80:393–406
    [Google Scholar]
  100. 100. 
    Bragg DC, Mangkalaphiban K, Vaine CA, Kulkarni NJ, Shin D et al. 2017. Disease onset in X-linked dystonia–parkinsonism correlates with expansion of a hexameric repeat within an SVA retrotransposon in TAF1. . PNAS 114:E11020–28
    [Google Scholar]
  101. 101. 
    Aneichyk T, Hendriks WT, Yadav R, Shin D, Gao D et al. 2018. Dissecting the causal mechanism of X-linked dystonia–parkinsonism by integrating genome and transcriptome assembly. Cell 172:897–909
    [Google Scholar]
  102. 102. 
    Cordaux R, Batzer MA. 2009. The impact of retrotransposons on human genome evolution. Nat. Rev. Genet. 10:691–703
    [Google Scholar]
  103. 103. 
    Rickman KA, Lach FP, Abhyankar A, Donovan FX, Sanborn EM et al. 2015. Deficiency of UBE2T, the E2 ubiquitin ligase necessary for FANCD2 and FANCI ubiquitination, causes FA-T subtype of Fanconi anemia. Cell Rep 12:35–41
    [Google Scholar]
  104. 104. 
    Pisanic TR 2nd, S Asaka, Lin SF, Yen TT, Sun H et al. 2019. Long interspersed nuclear element 1 retrotransposons become deregulated during the development of ovarian cancer precursor lesions. Am. J. Pathol. 189:513–20
    [Google Scholar]
  105. 105. 
    Miki Y, Nishisho I, Horii A, Miyoshi Y, Utsunomiya J et al. 1992. Disruption of the APC gene by a retrotransposal insertion of L1 sequence in a colon cancer. Cancer Res 52:643–45
    [Google Scholar]
  106. 106. 
    Achanta P, Steranka JP, Tang Z, Rodic N, Sharma R et al. 2016. Somatic retrotransposition is infrequent in glioblastomas. Mob. DNA 7:22
    [Google Scholar]
  107. 107. 
    Carreira PE, Ewing AD, Li G, Schauer SN, Upton KR et al. 2016. Evidence for L1-associated DNA rearrangements and negligible L1 retrotransposition in glioblastoma multiforme. Mob. DNA 7:21
    [Google Scholar]
  108. 108. 
    Lee E, Iskow R, Yang L, Gokcumen O, Haseley P et al. 2012. Landscape of somatic retrotransposition in human cancers. Science 337:967–71
    [Google Scholar]
  109. 109. 
    Helman E, Lawrence MS, Stewart C, Sougnez C, Getz G, Meyerson M 2014. Somatic retrotransposition in human cancer revealed by whole-genome and exome sequencing. Genome Res 24:1053–63
    [Google Scholar]
  110. 110. 
    Tubio JM, Li Y, Ju YS, Martincorena I, Cooke SL et al. 2014. Extensive transduction of nonrepetitive DNA mediated by L1 retrotransposition in cancer genomes. Science 345:1251343
    [Google Scholar]
  111. 111. 
    Tang Z, Steranka JP, Ma S, Grivainis M, Rodic N et al. 2017. Human transposon insertion profiling: analysis, visualization and identification of somatic LINE-1 insertions in ovarian cancer. PNAS 114:E733–40
    [Google Scholar]
  112. 112. 
    Nguyen THM, Carreira PE, Sanchez-Luque FJ, Schauer SN, Fagg AC et al. 2018. L1 retrotransposon heterogeneity in ovarian tumor cell evolution. Cell Rep 23:3730–40
    [Google Scholar]
  113. 113. 
    Solyom S, Ewing AD, Rahrmann EP, Doucet T, Nelson HH et al. 2012. Extensive somatic L1 retrotransposition in colorectal tumors. Genome Res 22:2328–38
    [Google Scholar]
  114. 114. 
    Rodic N, Steranka JP, Makohon-Moore A, Moyer A, Shen P et al. 2015. Retrotransposon insertions in the clonal evolution of pancreatic ductal adenocarcinoma. Nat. Med. 21:1060–64
    [Google Scholar]
  115. 115. 
    Cooke SL, Shlien A, Marshall J, Pipinikas CP, Martincorena I et al. 2014. Processed pseudogenes acquired somatically during cancer development. Nat. Commun. 5:3644
    [Google Scholar]
  116. 116. 
    Burns KH. 2017. Transposable elements in cancer. Nat. Rev. Cancer 17:415–24
    [Google Scholar]
  117. 117. 
    Belgnaoui SM, Gosden RG, Semmes OJ, Haoudi A 2006. Human LINE-1 retrotransposon induces DNA damage and apoptosis in cancer cells. Cancer Cell Int 6:13
    [Google Scholar]
  118. 118. 
    Haoudi A, Semmes OJ, Mason JM, Cannon RE 2004. Retrotransposition-competent human LINE-1 induces apoptosis in cancer cells with intact p53. . J. Biomed. Biotechnol 2004:185–94
    [Google Scholar]
  119. 119. 
    Gasior SL, Wakeman TP, Xu B, Deininger PL 2006. The human LINE-1 retrotransposon creates DNA double-strand breaks. J. Mol. Biol. 357:1383–93
    [Google Scholar]
  120. 120. 
    Bartkova J, Rezaei N, Liontos M, Karakaidos P, Kletsas D et al. 2006. Oncogene-induced senescence is part of the tumorigenesis barrier imposed by DNA damage checkpoints. Nature 444:633–37
    [Google Scholar]
  121. 121. 
    Di Micco R, Fumagalli M, Cicalese A, Piccinin S, Gasparini P et al. 2006. Oncogene-induced senescence is a DNA damage response triggered by DNA hyper-replication. Nature 444:638–42
    [Google Scholar]
  122. 122. 
    Leonova KI, Brodsky L, Lipchick B, Pal M, Novototskaya L et al. 2013. p53 cooperates with DNA methylation and a suicidal interferon response to maintain epigenetic silencing of repeats and noncoding RNAs. PNAS 110:E89–98
    [Google Scholar]
  123. 123. 
    Chiappinelli KB, Strissel PL, Desrichard A, Li H, Henke C et al. 2015. Inhibiting DNA methylation causes an interferon response in cancer via dsRNA including endogenous retroviruses. Cell 162:974–86
    [Google Scholar]
  124. 124. 
    Roulois D, Loo Yau H, Singhania R, Wang Y, Danesh A et al. 2015. DNA-demethylating agents target colorectal cancer cells by inducing viral mimicry by endogenous transcripts. Cell 162:961–73
    [Google Scholar]
  125. 125. 
    Jones PA, Ohtani H, Chakravarthy A, De Carvalho DD 2019. Epigenetic therapy in immune-oncology. Nat. Rev. Cancer 19:151–61
    [Google Scholar]
  126. 126. 
    Sheng W, LaFleur MW, Nguyen TH, Chen S, Chakravarthy A et al. 2018. LSD1 ablation stimulates anti-tumor immunity and enables checkpoint blockade. Cell 174:549–63
    [Google Scholar]
  127. 127. 
    Ahmad S, Mu X, Yang F, Greenwald E, Park JW et al. 2018. Breaching self-tolerance to Alu duplex RNA underlies MDA5-mediated inflammation. Cell 172:797–810
    [Google Scholar]
  128. 128. 
    Lim YW, Sanz LA, Xu X, Hartono SR, Chedin F 2015. Genome-wide DNA hypomethylation and RNA:DNA hybrid accumulation in Aicardi-Goutières syndrome. eLife 4:e08007
    [Google Scholar]
  129. 129. 
    Thomas CA, Tejwani L, Trujillo CA, Negraes PD, Herai RH et al. 2017. Modeling of TREX1-dependent autoimmune disease using human stem cells highlights L1 accumulation as a source of neuroinflammation. Cell Stem Cell 21:319–31
    [Google Scholar]
  130. 130. 
    De Cecco M, Ito T, Petrashen AP, Elias AE, Skvir NJ et al. 2019. L1 drives IFN in senescent cells and promotes age-associated inflammation. Nature 566:73–78
    [Google Scholar]
  131. 131. 
    Fahrner JA, Bjornsson HT. 2014. Mendelian disorders of the epigenetic machinery: tipping the balance of chromatin states. Annu. Rev. Genom. Hum. Genet. 15:269–93
    [Google Scholar]
  132. 132. 
    Denli AM, Narvaiza I, Kerman BE, Pena M, Benner C et al. 2015. Primate-specific ORF0 contributes to retrotransposon-mediated diversity. Cell 163:583–93
    [Google Scholar]
/content/journals/10.1146/annurev-pathmechdis-012419-032633
Loading
/content/journals/10.1146/annurev-pathmechdis-012419-032633
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