1932

Abstract

Covalent DNA–protein crosslinks (DPCs) are pervasive DNA lesions that interfere with essential chromatin processes such as transcription or replication. This review strives to provide an overview of the sources and principles of cellular DPC formation. DPCs are caused by endogenous reactive metabolites and various chemotherapeutic agents. However, in certain conditions DPCs also arise physiologically in cells. We discuss the cellular mechanisms resolving these threats to genomic integrity. Detection and repair of DPCs require not only the action of canonical DNA repair pathways but also the activity of specialized proteolytic enzymes—including proteases of the SPRTN/Wss1 family—to degrade the crosslinked protein. Loss of DPC repair capacity has dramatic consequences, ranging from genome instability in yeast and worms to cancer predisposition and premature aging in mice and humans.

Loading

Article metrics loading...

/content/journals/10.1146/annurev-biochem-032620-105820
2022-06-21
2024-04-26
Loading full text...

Full text loading...

/deliver/fulltext/biochem/91/1/annurev-biochem-032620-105820.html?itemId=/content/journals/10.1146/annurev-biochem-032620-105820&mimeType=html&fmt=ahah

Literature Cited

  1. 1.
    Lindahl T. 1993. Instability and decay of the primary structure of DNA. Nature 362:709–15
    [Google Scholar]
  2. 2.
    Chatterjee N, Walker GC. 2017. Mechanisms of DNA damage, repair, and mutagenesis. Environ. Mol. Mutagen. 58:235–63
    [Google Scholar]
  3. 3.
    Schumacher B, Pothof J, Vijg J, Hoeijmakers JHJ. 2021. The central role of DNA damage in the ageing process. Nature 592:695–703
    [Google Scholar]
  4. 4.
    Roos WP, Thomas AD, Kaina B. 2016. DNA damage and the balance between survival and death in cancer biology. Nat. Rev. Cancer 16:20–33
    [Google Scholar]
  5. 5.
    Jackson SP, Bartek J. 2009. The DNA-damage response in human biology and disease. Nature 461:1071–78
    [Google Scholar]
  6. 6.
    Jeggo PA, Pearl LH, Carr AM. 2016. DNA repair, genome stability and cancer: a historical perspective. Nat. Rev. Cancer 16:35–42
    [Google Scholar]
  7. 7.
    Curtin NJ. 2012. DNA repair dysregulation from cancer driver to therapeutic target. Nat. Rev. Cancer 12:801–17
    [Google Scholar]
  8. 8.
    O'Connor MJ. 2015. Targeting the DNA damage response in cancer. Mol. Cell 60:547–60
    [Google Scholar]
  9. 9.
    O'Neil NJ, Bailey ML, Hieter P. 2017. Synthetic lethality and cancer. Nat. Rev. Genet. 18:613–23
    [Google Scholar]
  10. 10.
    Lord CJ, Ashworth A. 2017. PARP inhibitors: synthetic lethality in the clinic. Science 355:1152–58
    [Google Scholar]
  11. 11.
    Bryant HE, Schultz N, Thomas HD, Parker KM, Flower D et al. 2005. Specific killing of BRCA2-deficient tumours with inhibitors of poly(ADP-ribose) polymerase. Nature 434:913–17
    [Google Scholar]
  12. 12.
    Farmer H, McCabe N, Lord CJ, Tutt AN, Johnson DA et al. 2005. Targeting the DNA repair defect in BRCA mutant cells as a therapeutic strategy. Nature 434:917–21
    [Google Scholar]
  13. 13.
    Smith KC, O'Leary ME. 1967. Photoinduced DNA-protein cross-links and bacterial killing: a correlation at low temperatures. Science 155:1024–26
    [Google Scholar]
  14. 14.
    Moulé Y, Moreau S, Aujard C. 1980. Induction of cross-links between DNA and protein by PR toxin, a mycotoxin from Penicillium roqueforti. Mutat. Res. 77:79–89
    [Google Scholar]
  15. 15.
    Wilkins RJ, Macleod HD. 1976. Formaldehyde induced DNA–protein crosslinks in Escherichia Coli. Mutat. Res. 36:11–16
    [Google Scholar]
  16. 16.
    Solomon MJ, Varshavsky A. 1985. Formaldehyde-mediated DNA–protein crosslinking: a probe for in vivo chromatin structures. PNAS 82:6470–74
    [Google Scholar]
  17. 17.
    Barker S, Weinfeld M, Murray D. 2005. DNA–protein crosslinks: their induction, repair, and biological consequences. Mutat. Res. 589:111–35
    [Google Scholar]
  18. 18.
    Stingele J, Schwarz MS, Bloemeke N, Wolf PG, Jentsch S. 2014. A DNA-dependent protease involved in DNA-protein crosslink repair. Cell 158:327–38Identifies Wss1 as the first protease specifically dedicated to DPC repair.
    [Google Scholar]
  19. 19.
    Duxin JP, Dewar JM, Yardimci H, Walter JC. 2014. Repair of a DNA-protein crosslink by replication-coupled proteolysis. Cell 159:346–57
    [Google Scholar]
  20. 20.
    Stingele J, Bellelli R, Alte F, Hewitt G, Sarek G et al. 2016. Mechanism and regulation of DNA-protein crosslink repair by the DNA-dependent metalloprotease SPRTN. Mol. Cell 64:688–703Identifies SPRTN as the main DPC protease in higher eukaryotes.
    [Google Scholar]
  21. 21.
    Vaz B, Popovic M, Newman JA, Fielden J, Aitkenhead H et al. 2016. Metalloprotease SPRTN/DVC1 orchestrates replication-coupled DNA-protein crosslink repair. Mol. Cell 64:704–19Identifies SPRTN as the main DPC protease in higher eukaryotes.
    [Google Scholar]
  22. 22.
    Lopez-Mosqueda J, Maddi K, Prgomet S, Kalayil S, Marinovic-Terzic I et al. 2016. SPRTN is a mammalian DNA-binding metalloprotease that resolves DNA-protein crosslinks. eLife 5:e21491
    [Google Scholar]
  23. 23.
    Maskey RS, Flatten KS, Sieben CJ, Peterson KL, Baker DJ et al. 2017. Spartan deficiency causes accumulation of Topoisomerase 1 cleavage complexes and tumorigenesis. Nucleic Acids Res 45:4564–76
    [Google Scholar]
  24. 24.
    Maskey RS, Kim MS, Baker DJ, Childs B, Malureanu LA et al. 2014. Spartan deficiency causes genomic instability and progeroid phenotypes. Nat. Commun. 5:5744
    [Google Scholar]
  25. 25.
    Lessel D, Vaz B, Halder S, Lockhart PJ, Marinovic-Terzic I et al. 2014. Mutations in SPRTN cause early onset hepatocellular carcinoma, genomic instability and progeroid features. Nat. Genet. 46:1239–44
    [Google Scholar]
  26. 26.
    Stingele J, Bellelli R, Boulton SJ. 2017. Mechanisms of DNA-protein crosslink repair. Nat. Rev. Mol. Cell Biol. 18:563–73
    [Google Scholar]
  27. 27.
    Nakano T, Xu X, Salem AMH, Shoulkamy MI, Ide H. 2017. Radiation-induced DNA–protein cross-links: mechanisms and biological significance. Free Radic. . Biol. Med. 107:136–45
    [Google Scholar]
  28. 28.
    Tretyakova NY, Groehler AT, Ji S 2015. DNA–protein cross-links: formation, structural identities, and biological outcomes. Acc. Chem. Res. 48:1631–44
    [Google Scholar]
  29. 29.
    Stingele J, Habermann B, Jentsch S. 2015. DNA–protein crosslink repair: proteases as DNA repair enzymes. Trends Biochem. Sci. 40:67–71
    [Google Scholar]
  30. 30.
    Nakano T, Ouchi R, Kawazoe J, Pack SP, Makino K, Ide H. 2012. T7 RNA polymerases backed up by covalently trapped proteins catalyze highly error prone transcription. J. Biol. Chem. 287:6562–72
    [Google Scholar]
  31. 31.
    Nakano T, Miyamoto-Matsubara M, Shoulkamy MI, Salem AM, Pack SP et al. 2013. Translocation and stability of replicative DNA helicases upon encountering DNA-protein cross-links. J. Biol. Chem. 288:4649–58
    [Google Scholar]
  32. 32.
    Fu YV, Yardimci H, Long DT, Ho TV, Guainazzi A et al. 2011. Selective bypass of a lagging strand roadblock by the eukaryotic replicative DNA helicase. Cell 146:931–41
    [Google Scholar]
  33. 33.
    Yeo JE, Wickramaratne S, Khatwani S, Wang YC, Vervacke J et al. 2014. Synthesis of site-specific DNA-protein conjugates and their effects on DNA replication. ACS Chem. Biol. 9:1860–68
    [Google Scholar]
  34. 34.
    Kuo HK, Griffith JD, Kreuzer KN. 2007. 5-Azacytidine–induced methyltransferase-DNA adducts block DNA replication in vivo. Cancer Res 67:8248–54
    [Google Scholar]
  35. 35.
    Mohni KN, Wessel SR, Zhao R, Wojciechowski AC, Luzwick JW et al. 2019. HMCES maintains genome integrity by shielding abasic sites in single-strand DNA. Cell 176:144–53.e13Provides evidence that HMCES protects abasic sites during replication by forming a covalent DPC.
    [Google Scholar]
  36. 36.
    Keeney S, Giroux CN, Kleckner N. 1997. Meiosis-specific DNA double-strand breaks are catalyzed by Spo11, a member of a widely conserved protein family. Cell 88:375–84
    [Google Scholar]
  37. 37.
    O'Brien PJ, Siraki AG, Shangari N. 2005. Aldehyde sources, metabolism, molecular toxicity mechanisms, and possible effects on human health. Crit. Rev. Toxicol. 35:609–62
    [Google Scholar]
  38. 38.
    Behrens UJ, Hoerner M, Lasker JM, Lieber CS. 1988. Formation of acetaldehyde adducts with ethanol-inducible P450IIE1 in vivo. Biochem. Biophys. Res. Commun. 154:584–90
    [Google Scholar]
  39. 39.
    Shi Y, Lan F, Matson C, Mulligan P, Whetstine JR et al. 2004. Histone demethylation mediated by the nuclear amine oxidase homolog LSD1. Cell 119:941–53
    [Google Scholar]
  40. 40.
    Tsukada Y, Fang J, Erdjument-Bromage H, Warren ME, Borchers CH et al. 2006. Histone demethylation by a family of JmjC domain-containing proteins. Nature 439:811–16
    [Google Scholar]
  41. 41.
    Dingler FA, Wang M, Mu A, Millington CL, Oberbeck N et al. 2020. Two aldehyde clearance systems are essential to prevent lethal formaldehyde accumulation in mice and humans. Mol. Cell 80:996–1012.e9
    [Google Scholar]
  42. 42.
    Burgos-Barragan G, Wit N, Meiser J, Dingler FA, Pietzke M et al. 2017. Mammals divert endogenous genotoxic formaldehyde into one-carbon metabolism. Nature 548:549–54
    [Google Scholar]
  43. 43.
    Groehler AIV, Kren S, Li Q, Robledo-Villafane M, Schmidt J et al. 2018. Oxidative cross-linking of proteins to DNA following ischemia-reperfusion injury. Free Radic. . Biol. Med. 120:89–101
    [Google Scholar]
  44. 44.
    Ji S, Shao H, Han Q, Seiler CL, Tretyakova NY. 2017. Reversible DNA–protein cross-linking at epigenetic DNA marks. Angew. Chem. Int. Ed. 56:14130–34
    [Google Scholar]
  45. 45.
    Sczepanski JT, Wong RS, McKnight JN, Bowman GD, Greenberg MM. 2010. Rapid DNA–protein cross-linking and strand scission by an abasic site in a nucleosome core particle. PNAS 107:22475–80
    [Google Scholar]
  46. 46.
    Michaelson-Richie ED, Loeber RL, Codreanu SG, Ming X, Liebler DC et al. 2010. DNA–protein cross-linking by 1,2,3,4-diepoxybutane. J. Proteome Res. 9:4356–67
    [Google Scholar]
  47. 47.
    Macfie A, Hagan E, Zhitkovich A. 2010. Mechanism of DNA–protein cross-linking by chromium. Chem. Res. Toxicol. 23:341–47
    [Google Scholar]
  48. 48.
    Loeber RL, Michaelson-Richie ED, Codreanu SG, Liebler DC, Campbell CR, Tretyakova NY. 2009. Proteomic analysis of DNA–protein cross-linking by antitumor nitrogen mustards. Chem. Res. Toxicol. 22:1151–62
    [Google Scholar]
  49. 49.
    Chvalova K, Brabec V, Kasparkova J. 2007. Mechanism of the formation of DNA–protein cross-links by antitumor cisplatin. Nucleic Acids Res 35:1812–21
    [Google Scholar]
  50. 50.
    Ward JF. 1994. The complexity of DNA damage: relevance to biological consequences. Int. J. Radiat Biol. 66:427–32
    [Google Scholar]
  51. 51.
    Zhang H, Koch CJ, Wallen CA, Wheeler KT. 1995. Radiation-induced DNA damage in tumors and normal tissues. III. Oxygen dependence of the formation of strand breaks and DNA–protein crosslinks. Radiat. Res. 142:163–68
    [Google Scholar]
  52. 52.
    Fornace AJ Jr., Little JB. 1977. DNA crosslinking induced by X-rays and chemical agents. Biochim. Biophys. Acta Nucleic Acids Protein Synth. 477:343–55
    [Google Scholar]
  53. 53.
    Shoulkamy MI, Nakano T, Ohshima M, Hirayama R, Uzawa A et al. 2012. Detection of DNA–protein crosslinks (DPCs) by novel direct fluorescence labeling methods: distinct stabilities of aldehyde and radiation-induced DPCs. Nucleic Acids Res 40:e143
    [Google Scholar]
  54. 54.
    Christman JK. 2002. 5-Azacytidine and 5-aza-2′-deoxycytidine as inhibitors of DNA methylation: mechanistic studies and their implications for cancer therapy. Oncogene 21:5483–95
    [Google Scholar]
  55. 55.
    Maslov AY, Lee M, Gundry M, Gravina S, Strogonova N et al. 2012. 5-aza-2′-deoxycytidine-induced genome rearrangements are mediated by DNMT1. Oncogene 31:5172–79
    [Google Scholar]
  56. 56.
    Fenaux P, Mufti GJ, Hellstrom-Lindberg E, Santini V, Finelli C et al. 2009. Efficacy of azacitidine compared with that of conventional care regimens in the treatment of higher-risk myelodysplastic syndromes: a randomised, open-label, phase III study. Lancet Oncol 10:223–32
    [Google Scholar]
  57. 57.
    Daskalakis M, Nguyen TT, Nguyen C, Guldberg P, Kohler G et al. 2002. Demethylation of a hypermethylated P15/INK4B gene in patients with myelodysplastic syndrome by 5-Aza-2′-deoxycytidine (decitabine) treatment. Blood 100:2957–64
    [Google Scholar]
  58. 58.
    Pommier Y, Marchand C. 2011. Interfacial inhibitors: targeting macromolecular complexes. Nat. Rev. Drug Discov. 11:25–36
    [Google Scholar]
  59. 59.
    Pommier Y, Sun Y, Huang SN, Nitiss JL. 2016. Roles of eukaryotic topoisomerases in transcription, replication and genomic stability. Nat. Rev. Mol. Cell Biol. 17:703–21
    [Google Scholar]
  60. 60.
    Champoux JJ. 2001. DNA topoisomerases: structure, function, and mechanism. Annu. Rev. Biochem. 70:369–413
    [Google Scholar]
  61. 61.
    Liu LF, Desai SD, Li TK, Mao Y, Sun M, Sim SP. 2000. Mechanism of action of camptothecin. Ann. N. Y. Acad. Sci. 922:1–10
    [Google Scholar]
  62. 62.
    Sirikantaramas S, Yamazaki M, Saito K. 2008. Mutations in topoisomerase I as a self-resistance mechanism coevolved with the production of the anticancer alkaloid camptothecin in plants. PNAS 105:6782–86
    [Google Scholar]
  63. 63.
    Fujimori A, Harker WG, Kohlhagen G, Hoki Y, Pommier Y. 1995. Mutation at the catalytic site of topoisomerase I in CEM/C2, a human leukemia cell line resistant to camptothecin. Cancer Res 55:1339–46
    [Google Scholar]
  64. 64.
    Pourquier P, Ueng LM, Kohlhagen G, Mazumder A, Gupta M et al. 1997. Effects of uracil incorporation, DNA mismatches, and abasic sites on cleavage and religation activities of mammalian topoisomerase I. J. Biol. Chem. 272:7792–96
    [Google Scholar]
  65. 65.
    Pourquier P, Waltman JL, Urasaki Y, Loktionova NA, Pegg AE et al. 2001. Topoisomerase I-mediated cytotoxicity of N-methyl-N′-nitro-N-nitrosoguanidine: trapping of topoisomerase I by the O6-methylguanine. Cancer Res 61:53–58
    [Google Scholar]
  66. 66.
    Nitiss JL. 2009. Targeting DNA topoisomerase II in cancer chemotherapy. Nat. Rev. Cancer 9:338–50
    [Google Scholar]
  67. 67.
    Gothe HJ, Bouwman BAM, Gusmao EG, Piccinno R, Petrosino G et al. 2019. Spatial chromosome folding and active transcription drive DNA fragility and formation of oncogenic MLL translocations. Mol. Cell 75:267–83.e12
    [Google Scholar]
  68. 68.
    Canela A, Maman Y, Huang SN, Wutz G, Tang W et al. 2019. Topoisomerase II-induced chromosome breakage and translocation is determined by chromosome architecture and transcriptional activity. Mol. Cell 75:252–66.e8
    [Google Scholar]
  69. 69.
    Bernard P, Couturier M. 1992. Cell killing by the F plasmid CcdB protein involves poisoning of DNA-topoisomerase II complexes. J. Mol. Biol. 226:735–45
    [Google Scholar]
  70. 70.
    Bergerat A, de Massy B, Gadelle D, Varoutas PC, Nicolas A, Forterre P 1997. An atypical topoisomerase II from Archaea with implications for meiotic recombination. Nature 386:414–17
    [Google Scholar]
  71. 71.
    Neale MJ, Pan J, Keeney S. 2005. Endonucleolytic processing of covalent protein-linked DNA double-strand breaks. Nature 436:1053–57
    [Google Scholar]
  72. 72.
    Dheekollu J, Wiedmer A, Ayyanathan K, Deakyne JS, Messick TE, Lieberman PM. 2021. Cell-cycle-dependent EBNA1-DNA crosslinking promotes replication termination at oriP and viral episome maintenance. Cell 184:643–54.e13
    [Google Scholar]
  73. 73.
    Beard WA, Horton JK, Prasad R, Wilson SH 2019. Eukaryotic base excision repair: New approaches shine light on mechanism. Annu. Rev. Biochem. 88:137–62
    [Google Scholar]
  74. 74.
    Cortez D. 2019. Replication-coupled DNA repair. Mol. Cell 74:866–76
    [Google Scholar]
  75. 75.
    Thompson PS, Amidon KM, Mohni KN, Cortez D, Eichman BF 2019. Protection of abasic sites during DNA replication by a stable thiazolidine protein-DNA cross-link. Nat. Struct. Mol. Biol. 26:613–18
    [Google Scholar]
  76. 76.
    Halabelian L, Ravichandran M, Li Y, Zeng H, Rao A et al. 2019. Structural basis of HMCES interactions with abasic DNA and multivalent substrate recognition. Nat. Struct. Mol. Biol. 26:607–12
    [Google Scholar]
  77. 77.
    Wang N, Bao H, Chen L, Liu Y, Li Y et al. 2019. Molecular basis of abasic site sensing in single-stranded DNA by the SRAP domain of E. coli yedK. Nucleic Acids Res 47:10388–99
    [Google Scholar]
  78. 78.
    Aravind L, Anand S, Iyer LM. 2013. Novel autoproteolytic and DNA-damage sensing components in the bacterial SOS response and oxidized methylcytosine-induced eukaryotic DNA demethylation systems. Biol. Direct 8:20
    [Google Scholar]
  79. 79.
    Mehta KPM, Lovejoy CA, Zhao R, Heintzman DR, Cortez D 2020. HMCES maintains replication fork progression and prevents double-strand breaks in response to APOBEC deamination and abasic site formation. Cell Rep 31:107705
    [Google Scholar]
  80. 80.
    Biayna J, Garcia-Cao I, Alvarez MM, Salvadores M, Espinosa-Carrasco J et al. 2021. Loss of the abasic site sensor HMCES is synthetic lethal with the activity of the APOBEC3A cytosine deaminase in cancer cells. PLOS Biol 19:e3001176
    [Google Scholar]
  81. 81.
    Shukla V, Halabelian L, Balagere S, Samaniego-Castruita D, Feldman DE et al. 2020. HMCES functions in the alternative end-joining pathway of the DNA DSB repair during class switch recombination in B cells. Mol. Cell 77:384–94.e4
    [Google Scholar]
  82. 82.
    Kanellis P, Gagliardi M, Banath JP, Szilard RK, Nakada S et al. 2007. A screen for suppressors of gross chromosomal rearrangements identifies a conserved role for PLP in preventing DNA lesions. PLOS Genet 3:e134
    [Google Scholar]
  83. 83.
    Balakirev MY, Mullally JE, Favier A, Assard N, Sulpice E et al. 2015. Wss1 metalloprotease partners with Cdc48/Doa1 in processing genotoxic SUMO conjugates. eLife 4:e06763
    [Google Scholar]
  84. 84.
    Enderle J, Dorn A, Beying N, Trapp O, Puchta H. 2019. The protease WSS1A, the endonuclease MUS81, and the phosphodiesterase TDP1 are involved in independent pathways of DNA–protein crosslink repair in plants. Plant Cell 31:775–90
    [Google Scholar]
  85. 85.
    Homchan A, Sukted J, Mongkolsuk S, Jeruzalmi D, Matangkasombut O, Pakotiprapha D. 2020. Wss1 homolog from Candida albicans and its role in DNA–protein crosslink tolerance. Mol. Microbiol. 114:409–22
    [Google Scholar]
  86. 86.
    Noguchi C, Grothusen G, Anandarajan V, Martinez-Lage Garcia M, Terlecky D et al. 2017. Genetic controls of DNA damage avoidance in response to acetaldehyde in fission yeast. Cell Cycle 16:45–58
    [Google Scholar]
  87. 87.
    Reinking HK, Hofmann K, Stingele J. 2020. Function and evolution of the DNA-protein crosslink proteases Wss1 and SPRTN. DNA Repair 88:102822
    [Google Scholar]
  88. 88.
    Mosbech A, Gibbs-Seymour I, Kagias K, Thorslund T, Beli P et al. 2012. DVC1 (C1orf124) is a DNA damage–targeting p97 adaptor that promotes ubiquitin-dependent responses to replication blocks. Nat. Struct. Mol. Biol. 19:1084–92
    [Google Scholar]
  89. 89.
    Borgermann N, Ackermann L, Schwertman P, Hendriks IA, Thijssen K et al. 2019. SUMOylation promotes protective responses to DNA-protein crosslinks. EMBO J 38:e101496
    [Google Scholar]
  90. 90.
    Delabaere L, Orsi GA, Sapey-Triomphe L, Horard B, Couble P, Loppin B. 2014. The Spartan ortholog maternal haploid is required for paternal chromosome integrity in the Drosophila zygote. Curr. Biol. 24:2281–87
    [Google Scholar]
  91. 91.
    Ruijs MW, van Andel RN, Oshima J, Madan K, Nieuwint AW, Aalfs CM. 2003. Atypical progeroid syndrome: an unknown helicase gene defect?. Am. J. Med. Genet. A 116A:295–99
    [Google Scholar]
  92. 92.
    Morocz M, Zsigmond E, Toth R, Enyedi MZ, Pinter L, Haracska L. 2017. DNA-dependent protease activity of human Spartan facilitates replication of DNA–protein crosslink-containing DNA. Nucleic Acids Res 45:3172–88
    [Google Scholar]
  93. 93.
    Li F, Raczynska JE, Chen Z, Yu H 2019. Structural insight into DNA-dependent activation of human metalloprotease Spartan. Cell Rep 26:3336–46.e4
    [Google Scholar]
  94. 94.
    Reinking HK, Kang HS, Gotz MJ, Li HY, Kieser A et al. 2020. DNA structure-specific cleavage of DNA-protein crosslinks by the SPRTN protease. Mol. Cell 80:102–13.e6Reveals that SPRTN achieves specificity by cleaving DPCs in a DNA structure–specific manner.
    [Google Scholar]
  95. 95.
    Davis EJ, Lachaud C, Appleton P, Macartney TJ, Nathke I, Rouse J. 2012. DVC1 (C1orf124) recruits the p97 protein segregase to sites of DNA damage. Nat. Struct. Mol. Biol. 19:1093–100
    [Google Scholar]
  96. 96.
    Fielden J, Wiseman K, Torrecilla I, Li S, Hume S et al. 2020. TEX264 coordinates p97- and SPRTN-mediated resolution of topoisomerase 1-DNA adducts. Nat. Commun. 11:1274
    [Google Scholar]
  97. 97.
    Zhao S, Kieser A, Li HY, Reinking HK, Weickert P et al. 2021. A ubiquitin switch controls autocatalytic inactivation of the DNA–protein crosslink repair protease SPRTN. Nucleic Acids Res 49:902–15
    [Google Scholar]
  98. 98.
    Centore RC, Yazinski SA, Tse A, Zou L. 2012. Spartan/C1orf124, a reader of PCNA ubiquitylation and a regulator of UV-induced DNA damage response. Mol. Cell 46:625–35
    [Google Scholar]
  99. 99.
    Perry M, Biegert M, Kollala SS, Mallard H, Su G et al. 2021. USP11 mediates repair of DNA–protein cross-links by deubiquitinating SPRTN metalloprotease. J. Biol. Chem. 296:100396
    [Google Scholar]
  100. 100.
    Huang J, Zhou Q, Gao M, Nowsheen S, Zhao F et al. 2020. Tandem deubiquitination and acetylation of SPRTN promotes DNA-protein crosslink repair and protects against aging. Mol. Cell 79:824–35.e5
    [Google Scholar]
  101. 101.
    Ruggiano A, Ramadan K. 2021. DNA–protein crosslink proteases in genome stability. Commun. Biol. 4:11
    [Google Scholar]
  102. 102.
    Carmell MA, Dokshin GA, Skaletsky H, Hu YC, van Wolfswinkel JC et al. 2016. A widely employed germ cell marker is an ancient disordered protein with reproductive functions in diverse eukaryotes. eLife 5:e19993
    [Google Scholar]
  103. 103.
    Bhargava V, Goldstein CD, Russell L, Xu L, Ahmed M et al. 2020. GCNA preserves genome integrity and fertility across species. Dev. Cell 52:38–52.e10
    [Google Scholar]
  104. 104.
    Dokshin GA, Davis GM, Sawle AD, Eldridge MD, Nicholls PK et al. 2020. GCNA interacts with Spartan and Topoisomerase II to regulate genome stability. Dev. Cell 52:53–68.e6
    [Google Scholar]
  105. 105.
    Kojima Y, Machida Y, Palani S, Caulfield TR, Radisky ES et al. 2020. FAM111A protects replication forks from protein obstacles via its trypsin-like domain. Nat. Commun. 11:1318
    [Google Scholar]
  106. 106.
    Fine DA, Rozenblatt-Rosen O, Padi M, Korkhin A, James RL et al. 2012. Identification of FAM111A as an SV40 host range restriction and adenovirus helper factor. PLOS Pathog 8:e1002949
    [Google Scholar]
  107. 107.
    Unger S, Gorna MW, Le Bechec A, Do Vale-Pereira S, Bedeschi MF et al. 2013. FAM111A mutations result in hypoparathyroidism and impaired skeletal development. Am. J. Hum. Genet. 92:990–95
    [Google Scholar]
  108. 108.
    Nie M, Oravcova M, Jami-Alahmadi Y, Wohlschlegel JA, Lazzerini-Denchi E, Boddy MN. 2021. FAM111A induces nuclear dysfunction in disease and viral restriction. EMBO Rep 22:e50803
    [Google Scholar]
  109. 109.
    Hoffmann S, Pentakota S, Mund A, Haahr P, Coscia F et al. 2020. FAM111 protease activity undermines cellular fitness and is amplified by gain-of-function mutations in human disease. EMBO Rep 21:e50662
    [Google Scholar]
  110. 110.
    Serbyn N, Noireterre A, Bagdiul I, Plank M, Michel AH et al. 2020. The aspartic protease Ddi1 contributes to DNA-protein crosslink repair in yeast. Mol. Cell 77:1066–79.e9
    [Google Scholar]
  111. 111.
    Svoboda M, Konvalinka J, Trempe JF, Grantz Saskova K. 2019. The yeast proteases Ddi1 and Wss1 are both involved in the DNA replication stress response. DNA Repair 80:45–51
    [Google Scholar]
  112. 112.
    Yip MCJ, Bodnar NO, Rapoport TA. 2020. Ddi1 is a ubiquitin-dependent protease. PNAS 117:7776–81
    [Google Scholar]
  113. 113.
    Dirac-Svejstrup AB, Walker J, Faull P, Encheva V, Akimov V et al. 2020. DDI2 is a ubiquitin-directed endoprotease responsible for cleavage of transcription factor NRF1. Mol. Cell 79:332–41.e7
    [Google Scholar]
  114. 114.
    Lehrbach NJ, Ruvkun G. 2016. Proteasome dysfunction triggers activation of SKN-1A/Nrf1 by the aspartic protease DDI-1. eLife 5:e17721
    [Google Scholar]
  115. 115.
    Koizumi S, Irie T, Hirayama S, Sakurai Y, Yashiroda H et al. 2016. The aspartyl protease DDI2 activates Nrf1 to compensate for proteasome dysfunction. eLife 5:e18357
    [Google Scholar]
  116. 116.
    Lin CP, Ban Y, Lyu YL, Desai SD, Liu LF. 2008. A ubiquitin-proteasome pathway for the repair of topoisomerase I-DNA covalent complexes. J. Biol. Chem. 283:21074–83
    [Google Scholar]
  117. 117.
    Sun Y, Miller Jenkins LM, Su YP, Nitiss KC, Nitiss JL, Pommier Y 2020. A conserved SUMO pathway repairs topoisomerase DNA-protein cross-links by engaging ubiquitin-mediated proteasomal degradation. Sci. Adv. 6:eaba6290
    [Google Scholar]
  118. 118.
    Mao Y, Desai SD, Ting CY, Hwang J, Liu LF. 2001. 26 S proteasome-mediated degradation of topoisomerase II cleavable complexes. J. Biol. Chem. 276:40652–58
    [Google Scholar]
  119. 119.
    Baker DJ, Wuenschell G, Xia L, Termini J, Bates SE et al. 2007. Nucleotide excision repair eliminates unique DNA-protein cross-links from mammalian cells. J. Biol. Chem. 282:22592–604
    [Google Scholar]
  120. 120.
    Quievryn G, Zhitkovich A. 2000. Loss of DNA–protein crosslinks from formaldehyde-exposed cells occurs through spontaneous hydrolysis and an active repair process linked to proteosome function. Carcinogenesis 21:1573–80
    [Google Scholar]
  121. 121.
    Larsen NB, Gao AO, Sparks JL, Gallina I, Wu RA et al. 2019. Replication-coupled DNA-protein crosslink repair by SPRTN and the proteasome in Xenopus egg extracts. Mol. Cell 73:574–88.e7Shows parallel replication-coupled DPC proteolysis by SPRTN and the proteasome.
    [Google Scholar]
  122. 122.
    Sparks JL, Chistol G, Gao AO, Raschle M, Larsen NB et al. 2019. The CMG helicase bypasses DNA-protein cross-links to facilitate their repair. Cell 176:167–81.e21Describes the mechanism of DPC bypass by the CMG helicase.
    [Google Scholar]
  123. 123.
    Gallina I, Hendriks IA, Hoffmann S, Larsen NB, Johansen J et al. 2021. The ubiquitin ligase RFWD3 is required for translesion DNA synthesis. Mol. Cell 81:442–58.e9
    [Google Scholar]
  124. 124.
    Vannier JB, Sarek G, Boulton SJ. 2014. RTEL1: functions of a disease-associated helicase. Trends Cell Biol 24:416–25
    [Google Scholar]
  125. 125.
    Elia AE, Wang DC, Willis NA, Boardman AP, Hajdu I et al. 2015. RFWD3-dependent ubiquitination of RPA regulates repair at stalled replication forks. Mol. Cell 60:280–93
    [Google Scholar]
  126. 126.
    Feeney L, Munoz IM, Lachaud C, Toth R, Appleton PL et al. 2017. RPA-mediated recruitment of the E3 ligase RFWD3 is vital for interstrand crosslink repair and human health. Mol. Cell 66:610–21.e4
    [Google Scholar]
  127. 127.
    Inano S, Sato K, Katsuki Y, Kobayashi W, Tanaka H et al. 2017. RFWD3-mediated ubiquitination promotes timely removal of both RPA and RAD51 from DNA damage sites to facilitate homologous recombination. Mol. Cell 66:622–34.e8
    [Google Scholar]
  128. 128.
    Knies K, Inano S, Ramirez MJ, Ishiai M, Surralles J et al. 2017. Biallelic mutations in the ubiquitin ligase RFWD3 cause Fanconi anemia. J. Clin. Investig. 127:3013–27
    [Google Scholar]
  129. 129.
    Sale JE. 2013. Translesion DNA synthesis and mutagenesis in eukaryotes. Cold Spring Harb. Perspect. Biol. 5:a012708
    [Google Scholar]
  130. 130.
    Pommier Y, Huang SY, Gao R, Das BB, Murai J, Marchand C. 2014. Tyrosyl-DNA-phosphodiesterases (TDP1 and TDP2). DNA Repair 19:114–29
    [Google Scholar]
  131. 131.
    Pouliot JJ, Yao KC, Robertson CA, Nash HA. 1999. Yeast gene for a Tyr-DNA phosphodiesterase that repairs topoisomerase I complexes. Science 286:552–55Identifies Tdp1 as the enzyme releasing covalent Top1 DPCs in yeast.
    [Google Scholar]
  132. 132.
    Yang SW, Burgin AB Jr., Huizenga BN, Robertson CA, Yao KC, Nash HA. 1996. A eukaryotic enzyme that can disjoin dead-end covalent complexes between DNA and type I topoisomerases. PNAS 93:11534–39Identifies Tdp1 as the enzyme releasing covalent Top1 DPCs in yeast.
    [Google Scholar]
  133. 133.
    Takashima H, Boerkoel CF, John J, Saifi GM, Salih MA et al. 2002. Mutation of TDP1, encoding a topoisomerase I-dependent DNA damage repair enzyme, in spinocerebellar ataxia with axonal neuropathy. Nat. Genet. 32:267–72
    [Google Scholar]
  134. 134.
    El-Khamisy SF, Saifi GM, Weinfeld M, Johansson F, Helleday T et al. 2005. Defective DNA single-strand break repair in spinocerebellar ataxia with axonal neuropathy-1. Nature 434:108–13
    [Google Scholar]
  135. 135.
    Interthal H, Chen HJ, Kehl-Fie TE, Zotzmann J, Leppard JB, Champoux JJ. 2005. SCAN1 mutant Tdp1 accumulates the enzyme–DNA intermediate and causes camptothecin hypersensitivity. EMBO J 24:2224–33
    [Google Scholar]
  136. 136.
    Interthal H, Champoux JJ. 2011. Effects of DNA and protein size on substrate cleavage by human tyrosyl-DNA phosphodiesterase 1. Biochem. J. 436:559–66
    [Google Scholar]
  137. 137.
    Steinacher R, Osman F, Lorenz A, Bryer C, Whitby MC 2013. Slx8 removes Pli1-dependent protein-SUMO conjugates including SUMOylated topoisomerase I to promote genome stability. PLOS ONE 8:e71960
    [Google Scholar]
  138. 138.
    Liu JCY, Kuhbacher U, Larsen NB, Borgermann N, Garvanska DH et al. 2021. Mechanism and function of DNA replication-independent DNA-protein crosslink repair via the SUMO-RNF4 pathway. EMBO J 40:e107413
    [Google Scholar]
  139. 139.
    Inamdar KV, Pouliot JJ, Zhou T, Lees-Miller SP, Rasouli-Nia A, Povirk LF 2002. Conversion of phosphoglycolate to phosphate termini on 3′ overhangs of DNA double strand breaks by the human tyrosyl-DNA phosphodiesterase hTdp1. J. Biol. Chem. 277:27162–68
    [Google Scholar]
  140. 140.
    Zhou T, Akopiants K, Mohapatra S, Lin PS, Valerie K et al. 2009. Tyrosyl-DNA phosphodiesterase and the repair of 3′-phosphoglycolate-terminated DNA double-strand breaks. DNA Repair 8:901–11
    [Google Scholar]
  141. 141.
    El-Khamisy SF. 2011. To live or to die: a matter of processing damaged DNA termini in neurons. EMBO Mol. Med. 3:78–88
    [Google Scholar]
  142. 142.
    Cortes Ledesma F, El Khamisy SF, Zuma MC, Osborn K, Caldecott KW 2009. A human 5′-tyrosyl DNA phosphodiesterase that repairs topoisomerase-mediated DNA damage. Nature 461:674–78Identifies TDP2 as a specific DNA repair enzyme for the repair of TOP2 DPCs.
    [Google Scholar]
  143. 143.
    Schellenberg MJ, Appel CD, Adhikari S, Robertson PD, Ramsden DA, Williams RS. 2012. Mechanism of repair of 5′-topoisomerase II-DNA adducts by mammalian tyrosyl-DNA phosphodiesterase 2. Nat. Struct. Mol. Biol. 19:1363–71
    [Google Scholar]
  144. 144.
    Wei Y, Diao LX, Lu S, Wang HT, Suo F et al. 2017. SUMO-targeted DNA translocase Rrp2 protects the genome from Top2-induced DNA damage. Mol. Cell 66:581–96.e6
    [Google Scholar]
  145. 145.
    Gao R, Schellenberg MJ, Huang SY, Abdelmalak M, Marchand C et al. 2014. Proteolytic degradation of topoisomerase II (Top2) enables the processing of Top2·DNA and Top2·RNA covalent complexes by tyrosyl-DNA-phosphodiesterase 2 (TDP2). J. Biol. Chem. 289:17960–69
    [Google Scholar]
  146. 146.
    Schellenberg MJ, Lieberman JA, Herrero-Ruiz A, Butler LR, Williams JG et al. 2017. ZATT (ZNF451)-mediated resolution of topoisomerase 2 DNA-protein cross-links. Science 357:1412–16
    [Google Scholar]
  147. 147.
    Gomez-Herreros F, Romero-Granados R, Zeng Z, Alvarez-Quilon A, Quintero C et al. 2013. TDP2-dependent non-homologous end-joining protects against topoisomerase II-induced DNA breaks and genome instability in cells and in vivo. PLOS Genet 9:e1003226
    [Google Scholar]
  148. 148.
    Paull TT. 2018. 20 years of Mre11 biology: no end in sight. Mol. Cell 71:419–27
    [Google Scholar]
  149. 149.
    Stohr BA, Kreuzer KN. 2001. Repair of topoisomerase-mediated DNA damage in bacteriophage T4. Genetics 158:19–28
    [Google Scholar]
  150. 150.
    Connelly JC, de Leau ES, Leach DR. 2003. Nucleolytic processing of a protein-bound DNA end by the E. coli SbcCD (MR) complex. DNA Repair 2:795–807
    [Google Scholar]
  151. 151.
    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]
  152. 152.
    Cannavo E, Cejka P. 2014. Sae2 promotes dsDNA endonuclease activity within Mre11-Rad50-Xrs2 to resect DNA breaks. Nature 514:122–25
    [Google Scholar]
  153. 153.
    Saathoff JH, Kashammer L, Lammens K, Byrne RT, Hopfner KP. 2018. The bacterial Mre11-Rad50 homolog SbcCD cleaves opposing strands of DNA by two chemically distinct nuclease reactions. Nucleic Acids Res 46:11303–14
    [Google Scholar]
  154. 154.
    Deshpande RA, Lee JH, 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]
  155. 155.
    Hartsuiker E, Mizuno K, Molnar M, Kohli J, Ohta K, Carr AM. 2009. Ctp1CtIP and Rad32Mre11 nuclease activity are required for Rec12Spo11 removal, but Rec12Spo11 removal is dispensable for other MRN-dependent meiotic functions. Mol. Cell. Biol. 29:1671–81
    [Google Scholar]
  156. 156.
    Hartsuiker E, Neale MJ, Carr AM. 2009. Distinct requirements for the Rad32(Mre11) nuclease and Ctp1(CtIP) in the removal of covalently bound topoisomerase I and II from DNA. Mol. Cell 33:117–23
    [Google Scholar]
  157. 157.
    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]
  158. 158.
    Hoa NN, Shimizu T, Zhou ZW, Wang ZQ, Deshpande RA et al. 2016. Mre11 is essential for the removal of lethal Topoisomerase 2 covalent cleavage complexes. Mol. Cell 64:580–92
    [Google Scholar]
  159. 159.
    Kashammer L, Saathoff JH, Lammens K, Gut F, Bartho J et al. 2019. Mechanism of DNA end sensing and processing by the Mre11-Rad50 complex. Mol. Cell 76:382–94.e6Describes the mechanism of DNA end detection by the MR complex.
    [Google Scholar]
  160. 160.
    Nakano T, Morishita S, Katafuchi A, Matsubara M, Horikawa Y et al. 2007. Nucleotide excision repair and homologous recombination systems commit differentially to the repair of DNA-protein crosslinks. Mol. Cell 28:147–58
    [Google Scholar]
  161. 161.
    de Graaf B, Clore A, McCullough AK. 2009. Cellular pathways for DNA repair and damage tolerance of formaldehyde-induced DNA-protein crosslinks. DNA Repair 8:1207–14
    [Google Scholar]
  162. 162.
    Nakano T, Katafuchi A, Matsubara M, Terato H, Tsuboi T et al. 2009. Homologous recombination but not nucleotide excision repair plays a pivotal role in tolerance of DNA-protein cross-links in mammalian cells. J. Biol. Chem. 284:27065–76
    [Google Scholar]
  163. 163.
    Serbyn N, Bagdiul I, Noireterre A, Michel AH, Suhandynata RT et al. 2021. SUMO orchestrates multiple alternative DNA-protein crosslink repair pathways. Cell Rep. 37:110034
    [Google Scholar]
  164. 164.
    Ruggiano A, Vaz B, Kilgas S, Popović M, Rodriguez-Berriguete G et al. 2021. The protease SPRTN and SUMOylation coordinate DNA-protein crosslink repair to prevent genome instability. Cell Rep. 37:110080
    [Google Scholar]
/content/journals/10.1146/annurev-biochem-032620-105820
Loading
/content/journals/10.1146/annurev-biochem-032620-105820
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