Skip to main content

Thank you for visiting nature.com. You are using a browser version with limited support for CSS. To obtain the best experience, we recommend you use a more up to date browser (or turn off compatibility mode in Internet Explorer). In the meantime, to ensure continued support, we are displaying the site without styles and JavaScript.

  • Review Article
  • Published:

Human topoisomerases and their roles in genome stability and organization

Abstract

Human topoisomerases comprise a family of six enzymes: two type IB (TOP1 and mitochondrial TOP1 (TOP1MT), two type IIA (TOP2A and TOP2B) and two type IA (TOP3A and TOP3B) topoisomerases. In this Review, we discuss their biochemistry and their roles in transcription, DNA replication and chromatin remodelling, and highlight the recent progress made in understanding TOP3A and TOP3B. Because of recent advances in elucidating the high-order organization of the genome through chromatin loops and topologically associating domains (TADs), we integrate the functions of topoisomerases with genome organization. We also discuss the physiological and pathological formation of irreversible topoisomerase cleavage complexes (TOPccs) as they generate topoisomerase DNA–protein crosslinks (TOP-DPCs) coupled with DNA breaks. We discuss the expanding number of redundant pathways that repair TOP-DPCs, and the defects in those pathways, which are increasingly recognized as source of genomic damage leading to neurological diseases and cancer.

This is a preview of subscription content, access via your institution

Access options

Buy this article

Prices may be subject to local taxes which are calculated during checkout

Fig. 1: Topological problems solved by human topoisomerases.
Fig. 2: Functions of topoisomerases in transcription.
Fig. 3: Functions of topoisomerases in genome organization.
Fig. 4: Genotoxic and pathogenic topoisomerase lesions.
Fig. 5: Topoisomerase-induced mutagenesis and recombination events.
Fig. 6: Main repair pathways for trapped topoisomerases in humans.

Similar content being viewed by others

References

  1. Pommier, Y., Sun, Y., Huang, S. N. & Nitiss, J. L. Roles of eukaryotic topoisomerases in transcription, replication and genomic stability. Nat. Rev. Mol. Cell Biol. 17, 703–721 (2016). This prior review complements the current Review with detailed biochemical and pharmacological information.

    Article  CAS  PubMed  Google Scholar 

  2. Saha, S. et al. DNA and RNA cleavage complexes and repair pathway for TOP3B RNA- and DNA–protein crosslinks. Cell Rep. 33, 108569 (2020). This work presents the first evidence for cellular TOP3B RNA cleavage complexes and repair pathways.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  3. Capranico, G., Marinello, J. & Chillemi, G. Type I DNA topoisomerases. J. Med. Chem. 60, 2169–2192 (2017). This work presents a detailed overview of the type I topoisomerases, which complements the present Review.

    Article  CAS  PubMed  Google Scholar 

  4. Lee, S. K. & Wang, W. Roles of topoisomerases in heterochromatin, aging, and diseases. Genes 10, 884 (2019).

    Article  CAS  PubMed Central  Google Scholar 

  5. Bjornsti, M. A. & Kaufmann, S. H. Topoisomerases and cancer chemotherapy: recent advances and unanswered questions. F1000Res. 8, 1704 (2019).

    Article  CAS  Google Scholar 

  6. Bizard, A. H. & Hickson, I. D. The many lives of type IA topoisomerases. J. Biol. Chem. 295, 7138–7153 (2020). This comprehensive review complements the present Review.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  7. Joshi, R. S., Nikolaou, C. & Roca, J. Structure and chromosomal organization of yeast genes regulated by topoisomerase II. Int. J. Mol. Sci. 19, 134 (2018).

    Article  PubMed Central  Google Scholar 

  8. Lee, J. H. & Berger, J. M. Cell cycle-dependent control and roles of DNA topoisomerase II. Genes 10, 859 (2019).

    Article  CAS  PubMed Central  Google Scholar 

  9. Dyson, S., Segura, J., Martinez-Garcia, B., Valdes, A. & Roca, J. Condensin minimizes topoisomerase II-mediated entanglements of DNA in vivo. EMBO J. 40, e105393 (2021). This article provides new insights into the coordination of SMCs and TOP2 in yeast.

    Article  CAS  PubMed  Google Scholar 

  10. Valdes, A., Segura, J., Dyson, S., Martinez-Garcia, B. & Roca, J. DNA knots occur in intracellular chromatin. Nucleic Acids Res. 46, 650–660 (2018).

    Article  CAS  PubMed  Google Scholar 

  11. Griffith, J. D. & Nash, H. A. Genetic rearrangement of DNA induces knots with a unique topology: implications for the mechanism of synapsis and crossing-over. Proc. Natl Acad. Sci. USA 82, 3124–3128 (1985).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  12. Nicholls, T. J. et al. Topoisomerase 3α is required for decatenation and segregation of human mtDNA. Mol. Cell 69, 9–23.e6 (2018).

    Article  CAS  PubMed  Google Scholar 

  13. Jiang, W. et al. Predominant cellular mitochondrial dysfunction in the TOP3A gene-caused Bloom syndrome-like disorder. Biochim. Biophys. Acta 1867, 166106 (2021).

    Article  CAS  Google Scholar 

  14. Ahmad, M. et al. RNA topoisomerase is prevalent in all domains of life and associates with polyribosomes in animals. Nucleic Acids Res. 44, 6335–6349 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  15. Ahmad, M., Xu, D. & Wang, W. Type IA topoisomerases can be “magicians” for both DNA and RNA in all domains of life. RNA Biol. 14, 854–864 (2017).

    Article  PubMed  PubMed Central  Google Scholar 

  16. Nitiss, J. L. DNA topoisomerase II and its growing repertoire of biological functions. Nat. Rev. Cancer 9, 327–337 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  17. Vann, K. R., Oviatt, A. A. & Osheroff, N. Topoisomerase II poisons: converting essential enzymes into molecular scissors. Biochemistry 60, 1630–1641 (2021). This work presents a complementary review of TOP2 in oncology.

    Article  CAS  PubMed  Google Scholar 

  18. Champoux, J. J. DNA topoisomerases: structure, function, and mechanism. Annu. Rev. Biochem. 70, 369–413 (2001).

    Article  CAS  PubMed  Google Scholar 

  19. Yang, W. Topoisomerases and site-specific recombinases: similarities in structure and mechanism. Crit. Rev. Biochem. Mol. Biol. 45, 520–534 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  20. Seol, Y., Zhang, H., Pommier, Y. & Neuman, K. C. A kinetic clutch governs religation by type IB topoisomerases and determines camptothecin sensitivity. Proc. Natl Acad. Sci. USA 109, 16125–16130 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  21. Stewart, L., Redinbo, M. R., Qiu, X., Hol, W. G. & Champoux, J. J. A model for the mechanism of human topoisomerase I. Science 279, 1534–1541 (1998).

    Article  CAS  PubMed  Google Scholar 

  22. Salceda, J., Fernandez, X. & Roca, J. Topoisomerase II, not topoisomerase I, is the proficient relaxase of nucleosomal DNA. EMBO J. 25, 2575–2583 (2006).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  23. Champoux, J. J. Evidence for an intermediate with a single-strand break in the reaction catalyzed by the DNA untwisting enzyme. Proc. Natl. Acad. Sci. USA 73, 3488–3491 (1976).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  24. Woessner, R. D., Mattern, M. R., Mirabelli, C. K., Johnson, R. K. & Drake, F. H. Proliferation- and cell cycle-dependent differences in expression of the 170 kilodalton and 180 kilodalton forms of topoisomerase II in NIH-3T3 cells. Cell Growth Differ. 2, 209–214 (1991).

    CAS  PubMed  Google Scholar 

  25. McClendon, A. K., Rodriguez, A. C. & Osheroff, N. Human topoisomerase IIα rapidly relaxes positively supercoiled DNA: implications for enzyme action ahead of replication forks. J. Biol. Chem. 280, 39337–39345 (2005).

    Article  CAS  PubMed  Google Scholar 

  26. Stoll, G. et al. Deletion of TOP3β, a component of FMRP-containing mRNPs, contributes to neurodevelopmental disorders. Nat. Neurosci. 16, 1228–1237 (2013). This article provides the first evidence that TOP3B genetic defects lead to neurological disorder and that TOP3B acts as a dual RNA and DNA topoisomerase.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  27. Yang, Y. et al. Arginine methylation facilitates the recruitment of TOP3B to chromatin to prevent R loop accumulation. Mol. Cell 53, 484–497 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  28. Wright, W. D., Shah, S. S. & Heyer, W. D. Homologous recombination and the repair of DNA double-strand breaks. J. Biol. Chem. 293, 10524–10535 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  29. Hoadley, K. A. et al. Defining the molecular interface that connects the Fanconi anemia protein FANCM to the Bloom syndrome dissolvasome. Proc. Natl Acad. Sci. USA 109, 4437–4442 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  30. Siaw, G. E., Liu, I. F., Lin, P. Y., Been, M. D. & Hsieh, T. S. DNA and RNA topoisomerase activities of Top3β are promoted by mediator protein Tudor domain-containing protein 3. Proc. Natl Acad. Sci. USA 113, E5544–5551 (2016).

    Article  PubMed  PubMed Central  Google Scholar 

  31. Thomas, A. & Pommier, Y. Targeting topoisomerase I in the era of precision medicine. Clin. Cancer Res. 25, 6581–6589 (2019). This article provides an up-to-date review of TOP1 inhibitors approved and in clinical development for cancer treatments.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  32. Nitiss, J. L. Targeting DNA topoisomerase II in cancer chemotherapy. Nat. Rev. Cancer 9, 338–350 (2009). This article provides an extensive review of topoisomerase inhibitors.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  33. Thakurela, S. et al. Gene regulation and priming by topoisomerase IIα in embryonic stem cells. Nat. Commun. 4, 2478 (2013).

    Article  PubMed  Google Scholar 

  34. Austin, C. A. et al. TOP2B: the first thirty years. Int. J. Mol. Sci. 19, 2765 (2018). This excellent article reviews the discovery and biology of TOP2B.

    Article  PubMed Central  Google Scholar 

  35. Pommier, Y. DNA Topoisomerases and Cancer (Springer, 2012).

  36. Rajapakse, V. N. et al. CellMinerCDB for integrative cross-database genomics and pharmacogenomics analyses of cancer cell lines. iScience 10, 247–264 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  37. Keszthelyi, A., Minchell, N. E. & Baxter, J. The causes and consequences of topological stress during DNA replication. Genes 7, 134 (2016).

    Article  PubMed Central  Google Scholar 

  38. Capranico, G., Jaxel, C., Roberge, M., Kohn, K. W. & Pommier, Y. Nucleosome positioning as a critical determinant for the DNA cleavage sites of mammalian DNA topoisomerase II in reconstituted simian virus 40 chromatin. Nucleic Acids Res. 18, 4553–4559 (1990).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  39. Le, T. T. et al. Synergistic coordination of chromatin torsional mechanics and topoisomerase activity. Cell 179, 619–631.e15 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  40. Servettaz, A. et al. Selective oxidation of DNA topoisomerase 1 induces systemic sclerosis in the mouse. J. Immunol. 182, 5855–5864 (2009).

    Article  CAS  PubMed  Google Scholar 

  41. Larcher, M. V. & Pasero, P. Top1 and Top2 promote replication fork arrest at a programmed pause site. Genes Dev. 34, 1–3 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  42. Lee, C. M., Wang, G., Pertsinidis, A. & Marians, K. J. Topoisomerase III acts at the replication fork to remove precatenanes. J. Bacteriol. 201, e00563-18 (2019).

    Article  PubMed  PubMed Central  Google Scholar 

  43. Schalbetter, S. A., Mansoubi, S., Chambers, A. L., Downs, J. A. & Baxter, J. Fork rotation and DNA precatenation are restricted during DNA replication to prevent chromosomal instability. Proc. Natl Acad. Sci. USA 112, E4565–4570 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  44. Minchell, N. E., Keszthelyi, A. & Baxter, J. Cohesin causes replicative DNA damage by trapping DNA topological stress. Mol. Cell 78, 739–751.e8 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  45. Boteva, L. et al. Common fragile sites are characterized by faulty condensin loading after replication stress. Cell Rep. 32, 108177 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  46. Sarni, D. et al. 3D genome organization contributes to genome instability at fragile sites. Nat. Commun. 11, 3613 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  47. Sima, J. et al. Identifying cis elements for spatiotemporal control of mammalian DNA replication. Cell 176, 816–830.e18 (2019).

    Article  CAS  PubMed  Google Scholar 

  48. Nitiss, J. L. DNA topoisomerase II and its growing repertoire of biological functions. Nat. Rev. Cancer 9, 327–337 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  49. Dewar, J. M., Budzowska, M. & Walter, J. C. The mechanism of DNA replication termination in vertebrates. Nature 525, 345–350 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  50. Shorrocks, A. K. et al. The Bloom syndrome complex senses RPA-coated single-stranded DNA to restart stalled replication forks. Nat. Commun. 12, 585 (2021).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  51. Davies, S. L., North, P. S. & Hickson, I. D. Role for BLM in replication-fork restart and suppression of origin firing after replicative stress. Nat. Struct. Mol. Biol. 14, 677–679 (2007).

    Article  CAS  PubMed  Google Scholar 

  52. Chan, K. L., North, P. S. & Hickson, I. D. BLM is required for faithful chromosome segregation and its localization defines a class of ultrafine anaphase bridges. EMBO J. 26, 3397–3409 (2007).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  53. Bizard, A. H. et al. PICH and TOP3A cooperate to induce positive DNA supercoiling. Nat. Struct. Mol. Biol. 26, 267–274 (2019). This article reveals a new function of TOP3A as reverse gyrase.

    Article  CAS  PubMed  Google Scholar 

  54. Liu, L. F. & Wang, J. C. Supercoiling of the DNA template during transcription. Proc. Natl Acad. Sci. USA 84, 7024–7027 (1987).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  55. Ma, J. & Wang, M. D. DNA supercoiling during transcription. Biophys. Rev. 8, 75–87 (2016). This article provides evidence for the selective activity of yeast TOP2 positive supercoils using single-molecule techniques.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  56. Baranello, L. et al. RNA polymerase II regulates topoisomerase 1 activity to favor efficient transcription. Cell 165, 357–371 (2016). This article provides evidence and a model for the role of TOP1 and BRD4 in transcription regulation.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  57. Racko, D., Benedetti, F., Dorier, J. & Stasiak, A. Transcription-induced supercoiling as the driving force of chromatin loop extrusion during formation of TADs in interphase chromosomes. Nucleic Acids Res. 46, 1648–1660 (2018). This article provides a model for the role of TOP1 and supercoiling in generating loop extrusion.

    Article  CAS  PubMed  Google Scholar 

  58. Kouzine, F., Levens, D. & Baranello, L. DNA topology and transcription. Nucleus 5, 195–202 (2014).

    Article  PubMed  PubMed Central  Google Scholar 

  59. Naughton, C. et al. Transcription forms and remodels supercoiling domains unfolding large-scale chromatin structures. Nat. Struct. Mol. Biol. 20, 387–395 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  60. Naughton, C., Corless, S. & Gilbert, N. Divergent RNA transcription: a role in promoter unwinding? Transcription 4, 162–166 (2013).

    Article  PubMed  PubMed Central  Google Scholar 

  61. Teves, S. S. & Henikoff, S. Transcription-generated torsional stress destabilizes nucleosomes. Nat. Struct. Mol. Biol. 21, 88–94 (2014).

    Article  CAS  PubMed  Google Scholar 

  62. King, I. F. et al. Topoisomerases facilitate transcription of long genes linked to autism. Nature 501, 58–62 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  63. Meng, L. et al. Towards a therapy for Angelman syndrome by targeting a long non-coding RNA. Nature 518, 409–412 (2015).

    Article  CAS  PubMed  Google Scholar 

  64. Rialdi, A. et al. Topoisomerase 1 inhibition suppresses inflammatory genes and protects from death by inflammation. Science 352, aad7993 (2016).

    Article  PubMed  PubMed Central  Google Scholar 

  65. Fragola, G. et al. Deletion of topoisomerase 1 in excitatory neurons causes genomic instability and early onset neurodegeneration. Nat. Commun. 11, 1962 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  66. Ho, J. S. Y. et al. TOP1 inhibition therapy protects against SARS-CoV-2-induced lethal inflammation. Cell 184, 2618–2632.e17 (2021).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  67. Tuduri, S. et al. Topoisomerase I suppresses genomic instability by preventing interference between replication and transcription. Nat. Cell Biol. 11, 1315–1324 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  68. Garcia-Muse, T. & Aguilera, A. R loops: from physiological to pathological roles. Cell 179, 604–618 (2019).

    Article  CAS  PubMed  Google Scholar 

  69. Promonet, A. et al. Topoisomerase 1 prevents replication stress at R-loop-enriched transcription termination sites. Nat. Commun. 11, 3940 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  70. Miglietta, G., Russo, M. & Capranico, G. G-quadruplex–R-loop interactions and the mechanism of anticancer G-quadruplex binders. Nucleic Acids Res. 48, 11942–11957 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  71. Wells, R. D. Non-B DNA conformations, mutagenesis and disease. Trends Biochem. Sci. 32, 271–278 (2007).

    Article  CAS  PubMed  Google Scholar 

  72. Sordet, O. et al. Ataxia telangiectasia mutated activation by transcription- and topoisomerase I-induced DNA double-strand breaks. EMBO Rep. 10, 887–893 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  73. Cristini, A. et al. Dual processing of R-loops and topoisomerase I induces transcription-dependent DNA double-strand breaks. Cell Rep. 28, 3167–3181.e6 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  74. Sims, R. J. 3rd et al. The C-terminal domain of RNA polymerase II is modified by site-specific methylation. Science 332, 99–103 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  75. Madabhushi, R. et al. Activity-induced DNA breaks govern the expression of neuronal early-response genes. Cell 161, 1592–1605 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  76. Madabhushi, R. The roles of DNA topoisomerase IIβ in transcription. Int. J. Mol. Sci. 19, 1917 (2018).

    Article  PubMed Central  Google Scholar 

  77. Haffner, M. C. et al. Androgen-induced TOP2B-mediated double-strand breaks and prostate cancer gene rearrangements. Nat. Genet. 42, 668–675 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  78. Haffner, M. C., De Marzo, A. M., Meeker, A. K., Nelson, W. G. & Yegnasubramanian, S. Transcription-induced DNA double strand breaks: both oncogenic force and potential therapeutic target? Clin. Cancer Res. 17, 3858–3864 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  79. Hedayati, M. et al. Androgen deprivation followed by acute androgen stimulation selectively sensitizes AR-positive prostate cancer cells to ionizing radiation. Clin. Cancer Res. 22, 3310–3319 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  80. Morimoto, S. et al. Type II DNA topoisomerases cause spontaneous double-strand breaks in genomic DNA. Genes 10, 868 (2019).

    Article  CAS  PubMed Central  Google Scholar 

  81. McNamara, S., Wang, H., Hanna, N. & Miller, W. H. Jr. Topoisomerase IIβ negatively modulates retinoic acid receptor α function: a novel mechanism of retinoic acid resistance. Mol. Cell Biol. 28, 2066–2077 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  82. Perillo, B. et al. DNA oxidation as triggered by H3K9me2 demethylation drives estrogen-induced gene expression. Science 319, 202–206 (2008).

    Article  CAS  PubMed  Google Scholar 

  83. Ju, B. G. et al. A topoisomerase IIβ-mediated dsDNA break required for regulated transcription. Science 312, 1798–1802 (2006).

    Article  CAS  PubMed  Google Scholar 

  84. Calderwood, S. K. A critical role for topoisomerase IIb and DNA double strand breaks in transcription. Transcription 7, 75–83 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  85. Bunch, H. et al. Transcriptional elongation requires DNA break-induced signalling. Nat. Commun. 6, 10191 (2015).

    Article  CAS  PubMed  Google Scholar 

  86. Buchel, G. et al. Association with Aurora-A controls N-MYC-dependent promoter escape and pause release of RNA polymerase II during the cell cycle. Cell Rep. 21, 3483–3497 (2017).

    Article  PubMed  PubMed Central  Google Scholar 

  87. Herrero-Ruiz, A. et al. Topoisomerase IIα represses transcription by enforcing promoter-proximal pausing. Cell Rep. 35, 108977 (2021).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  88. Itou, J. et al. Estrogen induces mammary ductal dysplasia via the upregulation of myc expression in a DNA-repair-deficient condition. iScience 23, 100821 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  89. Davidson, I. F. & Peters, J. M. Genome folding through loop extrusion by SMC complexes. Nat. Rev. Mol. Cell Biol. 22, 445–464 (2021). This work reviews chromatin loops and TADs.

    Article  CAS  PubMed  Google Scholar 

  90. Lieberman-Aiden, E. et al. Comprehensive mapping of long-range interactions reveals folding principles of the human genome. Science 326, 289–293 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  91. Misteli, T. The self-organizing genome: principles of genome architecture and function. Cell 183, 28–45 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  92. Waldman, T. Emerging themes in cohesin cancer biology. Nat. Rev. Cancer 20, 504–515 (2020).

    Article  CAS  PubMed  Google Scholar 

  93. Uhlmann, F. SMC complexes: from DNA to chromosomes. Nat. Rev. Mol. Cell Biol. 17, 399–412 (2016). This work is a detailed review of SMC complexes.

    Article  CAS  PubMed  Google Scholar 

  94. Aragon, L. The Smc5/6 complex: new and old functions of the enigmatic long-distance relative. Annu. Rev. Genet. 52, 89–107 (2018).

    Article  CAS  PubMed  Google Scholar 

  95. Bonner, J. N. et al. Smc5/6 mediated sumoylation of the Sgs1–Top3–Rmi1 complex promotes removal of recombination intermediates. Cell Rep. 16, 368–378 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  96. Verver, D. E. et al. Non-SMC element 2 (NSMCE2) of the SMC5/6 complex helps to resolve topological stress. Int. J. Mol. Sci. 16, 368–378 (2016).

    Google Scholar 

  97. Potts, P. R., Porteus, M. H. & Yu, H. Human SMC5/6 complex promotes sister chromatid homologous recombination by recruiting the SMC1/3 cohesin complex to double-strand breaks. EMBO J. 25, 3377–3388 (2006).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  98. Vian, L. et al. The energetics and physiological impact of cohesin extrusion. Cell 175, 292–294 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  99. Canela, A. et al. Genome organization drives chromosome fragility. Cell 170, 507–521.e18 (2017). This article uses End-seq to map TOP2 sites with cohesin and CTCF.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  100. Canela, A. et al. Topoisomerase II-induced chromosome breakage and translocation is determined by chromosome architecture and transcriptional activity. Mol. Cell 75, 252–266 (2019). This article uses the End-seq method developed by the authors to map TOP2 sites and their processing to frank breaks and translocations.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  101. Gothe, H. J. et al. Spatial chromosome folding and active transcription drive DNA fragility and formation of oncogenic MLL translocations. Mol. Cell 75, 267–283.e12 (2019).

    Article  CAS  PubMed  Google Scholar 

  102. Stigler, J., Camdere, G. O., Koshland, D. E. & Greene, E. C. Single-molecule imaging reveals a collapsed conformational state for DNA-bound cohesin. Cell Rep. 15, 988–998 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  103. Schwarzer, W. et al. Two independent modes of chromatin organization revealed by cohesin removal. Nature 551, 51–56 (2017).

    Article  PubMed  PubMed Central  Google Scholar 

  104. Rao, S. S. et al. A 3D map of the human genome at kilobase resolution reveals principles of chromatin looping. Cell 159, 1665–1680 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  105. Martinez-Garcia, P. M. et al. Genome-wide prediction of topoisomerase IIβ binding by architectural factors and chromatin accessibility. PLoS Comput. Biol. 17, e1007814 (2021).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  106. Uuskula-Reimand, L. et al. Topoisomerase IIβ interacts with cohesin and CTCF at topological domain borders. Genome Biol. 17, 182 (2016).

    Article  PubMed  PubMed Central  Google Scholar 

  107. Manville, C. M. et al. Genome-wide ChIP–seq analysis of human TOP2B occupancy in MCF7 breast cancer epithelial cells. Biol. Open. 4, 1436–1447 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  108. Al Mahmud, M. R. et al. TDP2 suppresses genomic instability induced by androgens in the epithelial cells of prostate glands. Genes. Cell 25, 450–465 (2020). This article suggests the role of abortive TOP2ccs and defective repair as a source of prostate cancer.

    Article  CAS  Google Scholar 

  109. Andersson, R. et al. An atlas of active enhancers across human cell types and tissues. Nature 507, 455–461 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  110. Hirabayashi, S. et al. NET-CAGE characterizes the dynamics and topology of human transcribed cis-regulatory elements. Nat. Genet. 51, 1369–1379 (2019).

    Article  CAS  PubMed  Google Scholar 

  111. Li, W., Notani, D. & Rosenfeld, M. G. Enhancers as non-coding RNA transcription units: recent insights and future perspectives. Nat. Rev. Genet. 17, 207–223 (2016).

    Article  CAS  PubMed  Google Scholar 

  112. Tang, Z. et al. CTCF-mediated human 3D genome architecture reveals chromatin topology for transcription. Cell 163, 1611–1627 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  113. Smith, E. M., Lajoie, B. R., Jain, G. & Dekker, J. Invariant TAD boundaries constrain cell-type-specific looping interactions between promoters and distal elements around the CFTR locus. Am. J. Hum. Genet. 98, 185–201 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  114. Lin, S. J. & O’Connell, M. J. DNA topoisomerase II modulates acetyl-regulation of cohesin-mediated chromosome dynamics. Curr. Genet. 63, 923–930 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  115. Holm, C., Goto, T., Wang, J. C. & Bolstein, D. DNA topoisomerase II is required at the time of mitosis in yeast. Cell. 41, 553–563 (1985).

    Article  CAS  PubMed  Google Scholar 

  116. Ishida, R. et al. Inhibition of intracellular topoisomerase II by antitumor bis(2,6-dioxopiperazine) derivatives: mode of cell growth inhibition distinct from that of cleavable complex-forming type inhibitors. Cancer Res. 51, 4909–4916 (1991).

    CAS  PubMed  Google Scholar 

  117. Fielding, A. B. et al. The deubiquitylase USP15 regulates topoisomerase IIα to maintain genome integrity. Oncogene 37, 2326–2342 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  118. Nielsen, C. F. et al. PICH promotes sister chromatid disjunction and co-operates with topoisomerase II in mitosis. Nat. Commun. 6, 8962 (2015).

    Article  CAS  PubMed  Google Scholar 

  119. Shintomi, K., Takahashi, T. S. & Hirano, T. Reconstitution of mitotic chromatids with a minimum set of purified factors. Nat. Cell Biol. 17, 1014–1023 (2015).

    Article  CAS  PubMed  Google Scholar 

  120. Ono, T., Sakamoto, C., Nakao, M., Saitoh, N. & Hirano, T. Condensin II plays an essential role in reversible assembly of mitotic chromosomes in situ. Mol. Biol. Cell 28, 2875–2886 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  121. Piskadlo, E. & Oliveira, R. A. A topology-centric view on mitotic chromosome architecture. Int. J. Mol. Sci. 18, 2751 (2017).

    Article  PubMed Central  Google Scholar 

  122. Sen, N. et al. Physical proximity of sister chromatids promotes Top2-dependent intertwining. Mol. Cell 64, 134–147 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  123. Earnshaw, W. C., Halligan, B., Cooke, C. A., Heck, M. M. & Liu, L. F. Topoisomerase II is a structural component of mitotic chromosome scaffolds. J. Cell. Biol. 100, 1706–1715 (1985).

    Article  CAS  PubMed  Google Scholar 

  124. Gasser, S. M., Laroche, T., Falquet, J., Boy de la Tour, E. & Laemmli, U. K. Metaphase chromosome structure. Involvement of topoisomerase II. J. Mol. Biol. 188, 613–629 (1986).

    Article  CAS  PubMed  Google Scholar 

  125. Daniloski, Z., Bisht, K. K., McStay, B. & Smith, S. Resolution of human ribosomal DNA occurs in anaphase, dependent on tankyrase 1, condensin II, and topoisomerase IIα. Genes. Dev. 33, 276–281 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  126. Loe, T. K. et al. Telomere length heterogeneity in ALT cells is maintained by PML-dependent localization of the BTR complex to telomeres. Genes Dev. 34, 650–662 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  127. Sobinoff, A. P. et al. BLM and SLX4 play opposing roles in recombination-dependent replication at human telomeres. EMBO J. 36, 2907–2919 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  128. Temime-Smaali, N. et al. Topoisomerase IIIα is required for normal proliferation and telomere stability in alternative lengthening of telomeres. EMBO J. 27, 1513–1524 (2008).

    CAS  PubMed  PubMed Central  Google Scholar 

  129. Ye, J. et al. TRF2 and apollo cooperate with topoisomerase 2α to protect human telomeres from replicative damage. Cell 142, 230–242 (2010).

    Article  CAS  PubMed  Google Scholar 

  130. Zhang, T. et al. Looping-out mechanism for resolution of replicative stress at telomeres. EMBO Rep. 18, 1412–1428 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  131. Clarke, D. J. & Azuma, Y. Non-catalytic roles of the topoisomerase IIα C-terminal domain. Int. J. Mol. Sci. 18, 2438 (2017).

    Article  PubMed Central  Google Scholar 

  132. Deiss, K. et al. A genome-wide RNAi screen identifies the SMC5/6 complex as a non-redundant regulator of a Topo2a-dependent G2 arrest. Nucleic Acids Res. 47, 2906–2921 (2019).

    Article  CAS  PubMed  Google Scholar 

  133. Bedez, C. et al. Post-translational modifications in DNA topoisomerase 2α highlight the role of a eukaryote-specific residue in the ATPase domain. Sci. Rep. 8, 9272 (2018).

    Article  PubMed  PubMed Central  Google Scholar 

  134. Vanden Broeck, A. et al. Structural basis for allosteric regulation of human topoisomerase IIα. Nat. Commun. 12, 2962 (2021).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  135. Dykhuizen, E. C. et al. BAF complexes facilitate decatenation of DNA by topoisomerase IIα. Nature 497, 624–627 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  136. Miller, E. L. et al. TOP2 synergizes with BAF chromatin remodeling for both resolution and formation of facultative heterochromatin. Nat. Struct. Mol. Biol. 24, 344–352 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  137. Dalcher, D. et al. BAZ2A safeguards genome architecture of ground-state pluripotent stem cells. EMBO J. 39, e105606 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  138. Lee, S. K. et al. Topoisomerase 3β interacts with RNAi machinery to promote heterochromatin formation and transcriptional silencing in Drosophila. Nat. Commun. 9, 4946 (2018).

    Article  PubMed  PubMed Central  Google Scholar 

  139. Husain, A. et al. Chromatin remodeller SMARCA4 recruits topoisomerase 1 and suppresses transcription-associated genomic instability. Nat. Commun. 7, 10549 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  140. Vanden Broeck, A. et al. Structural basis for allosteric regulation of human topoisomerase IIα. Nat. Commun. 12, 2962 (2021).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  141. Li, T. K. et al. Activation of topoisomerase II-mediated excision of chromosomal DNA loops during oxidative stress. Genes Dev. 13, 1553–1560 (1999).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  142. Wang, H. et al. Stimulation of topoisomerase II-mediated DNA damage via a mechanism involving protein thiolation. Biochemistry 40, 3316–3323 (2001).

    Article  CAS  PubMed  Google Scholar 

  143. Xiao, H., Li, T. K., Yang, J. M. & Liu, L. F. Acidic pH induces topoisomerase II-mediated DNA damage. Proc. Natl Acad. Sci. USA 100, 5205–5210 (2003).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  144. Kim, N. et al. Mutagenic processing of ribonucleotides in DNA by yeast topoisomerase I. Science 332, 1561–1564 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  145. Sekiguchi, J. & Shuman, S. Site-specific ribonuclease activity of eukaryotic DNA topoisomerase I. Mol. Cell 1, 89–97 (1997).

    Article  CAS  PubMed  Google Scholar 

  146. Sparks, J. L. et al. RNase H2-initiated ribonucleotide excision repair. Mol. Cell 47, 980–986 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  147. Reijns, M. A. & Jackson, A. P. Ribonuclease H2 in health and disease. Biochemical Soc. Trans. 42, 717–725 (2014).

    Article  CAS  Google Scholar 

  148. Williams, J. S., Lujan, S. A. & Kunkel, T. A. Processing ribonucleotides incorporated during eukaryotic DNA replication. Nat. Rev. Mol. Cell Biol. 17, 350–363 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  149. Alvarez-Quilon, A. et al. Endogenous DNA 3′ blocks are vulnerabilities for BRCA1 and BRCA2 deficiency and are reversed by the APE2 nuclease. Mol. Cell 78, 1152–1165.e8 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  150. Lin, Y. et al. APE2 promotes DNA damage response pathway from a single-strand break. Nucleic Acids Res. 46, 2479–2494 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  151. Kim, N. & Jinks-Robertson, S. The Top1 paradox: friend and foe of the eukaryotic genome. DNA Repair. 56, 33–41 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  152. Huang, S. Y., Ghosh, S. & Pommier, Y. Topoisomerase I alone is sufficient to produce short DNA deletions and can also reverse nicks at ribonucleotide sites. J. Biol. Chem. 290, 14068–14076 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  153. Williams, J. S. et al. Topoisomerase 1-mediated removal of ribonucleotides from nascent leading-strand DNA. Mol. Cell 49, 1010–1015 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  154. Sparks, J. L. & Burgers, P. M. Error-free and mutagenic processing of topoisomerase 1-provoked damage at genomic ribonucleotides. EMBO J. 34, 1259–1269 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  155. Saha, L. K. et al. Topoisomerase I-driven repair of UV-induced damage in NER-deficient cells. Proc. Natl Acad. Sci. USA 117, 14412–14420 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  156. Pourquier, P., Jensen, A. D., Gong, S. S., Pommier, Y. & Rogler, C. E. Human DNA topoisomerase I-mediated cleavage and recombination of duck hepatitis B virus DNA in vitro. Nucleic Acids Res. 27, 1919–1925 (1999).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  157. Shuman, S. Polynucleotide ligase activity of eukaryotic topoisomerase I. Mol. Cell 1, 741–748 (1998).

    Article  CAS  PubMed  Google Scholar 

  158. Sekiguchi, J., Seeman, N. C. & Shuman, S. Resolution of Holliday junctions by eukaryotic DNA topoisomerase I. Proc. Natl Acad. Sci. USA 93, 785–789 (1996).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  159. Zhao, B. et al. Topoisomerase 1 cleavage complex enables pattern recognition and inflammation during senescence. Nat. Commun. 11, 908 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  160. Gravel, S., Chapman, J. R., Magill, C. & Jackson, S. P. DNA helicases Sgs1 and BLM promote DNA double-strand break resection. Genes Dev. 22, 2767–2772 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  161. Wu, L. & Hickson, I. D. The Bloom’s syndrome helicase suppresses crossing over during homologous recombination. Nature 426, 870–874 (2003).

    Article  CAS  PubMed  Google Scholar 

  162. Piazza, A. & Heyer, W. D. Homologous recombination and the formation of complex genomic rearrangements. Trends Cell Biol. 29, 135–149 (2019).

    Article  CAS  PubMed  Google Scholar 

  163. Morotomi-Yano, K., Saito, S., Adachi, N. & Yano, K. I. Dynamic behavior of DNA topoisomerase IIβ in response to DNA double-strand breaks. Sci. Rep. 8, 10344 (2018).

    Article  PubMed  PubMed Central  Google Scholar 

  164. Arnould, C. et al. Loop extrusion as a mechanism for formation of DNA damage repair foci. Nature 590, 660–665 (2021).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  165. Miao, Z. H. et al. Nonclassic functions of human topoisomerase I: genome-wide and pharmacologic analyses. Cancer Res. 67, 8752–8761 (2007).

    Article  CAS  PubMed  Google Scholar 

  166. Morham, S. G., Kluckman, K. D., Voulomanos, N. & Smithies, O. Targeted disruption of the mouse topoisomerase I gene by camptothecin selection. Mol. Cell. Biol. 16, 6804–6809 (1996).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  167. Bonner, W. M. et al. γH2AX and cancer. Nat. Rev. Cancer 8, 957–967 (2008). This landmark review is by William Bonner, who discovered γH2AX.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  168. Belotserkovskii, B. P., Tornaletti, S., D’Souza, A. D. & Hanawalt, P. C. R-loop generation during transcription: formation, processing and cellular outcomes. DNA Repair. 71, 69–81 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  169. Manzo, S. G. et al. DNA topoisomerase I differentially modulates R-loops across the human genome. Genome Biol. 19, 100–100 (2018).

    Article  PubMed  PubMed Central  Google Scholar 

  170. Hamperl, S., Bocek, M. J., Saldivar, J. C., Swigut, T. & Cimprich, K. A. Transcription–replication conflict orientation modulates R-loop levels and activates distinct DNA damage responses. Cell 170, 774–786.e19 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  171. Cristini, A., Gromak, N. & Sordet, O. Transcription-dependent DNA double-strand breaks and human disease. Mol. Cell Oncol. 7, 1691905 (2020).

    Article  PubMed  PubMed Central  Google Scholar 

  172. Sordet, O., Nakamura, A. J., Redon, C. E. & Pommier, Y. DNA double-strand breaks and ATM activation by transcription-blocking DNA lesions. Cell Cycle 9, 274–278 (2010).

    Article  CAS  PubMed  Google Scholar 

  173. Zhang, H. et al. Increased negative supercoiling of mtDNA in TOP1mt knockout mice and presence of topoisomerases IIα and IIβ in vertebrate mitochondria. Nucleic Acids Res. 42, 7259–7267 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  174. Khiati, S. et al. Lack of mitochondrial topoisomerase I (TOP1mt) impairs liver regeneration. Proc. Natl Acad. Sci. USA 112, 11282–11287 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  175. Rosa, I. D., Zhang, H., Khiati, S., Wu, X. & Pommier, Y. Transcription profiling suggests that mitochondrial topoisomerase IB acts as a topological barrier and regulator of mitochondrial DNA transcription. J. Biol. Chem. 292, 20162–20172 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  176. Baechler, S. A. et al. The mitochondrial type IB topoisomerase drives mitochondrial translation and carcinogenesis. Nat. Commun. 10, 83–83 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  177. Baechler, S. A., Dalla Rosa, I., Spinazzola, A. & Pommier, Y. Beyond the unwinding: role of TOP1MT in mitochondrial translation. Cell Cycle 18, 2377–2384 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  178. Zoppoli, G. et al. Coordinated regulation of mitochondrial topoisomerase IB with mitochondrial nuclear encoded genes and MYC. Nucleic Acids Res. 39, 6620–6632 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  179. Adachi, Y., Luke, M. & Laemmli, U. K. Chromosome assembly in vitro: topoisomerase II is required for condensation. Cell 64, 137–148 (1991).

    Article  CAS  PubMed  Google Scholar 

  180. Akimitsu, N. et al. Enforced cytokinesis without complete nuclear division in embryonic cells depleting the activity of DNA topoisomerase IIα. Genes Cell 8, 393–402 (2003).

    Article  CAS  Google Scholar 

  181. Tiwari, V. K. et al. Target genes of topoisomerase IIβ regulate neuronal survival and are defined by their chromatin state. Proc. Natl Acad. Sci. USA 109, E934–E943 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  182. Lyu, Y. L. & Wang, J. C. Aberrant lamination in the cerebral cortex of mouse embryos lacking DNA topoisomerase IIβ. Proc. Natl Acad. Sci. USA 100, 7123–7128 (2003).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  183. Yang, X., Li, W., Prescott, E. D., Burden, S. J. & Wang, J. C. DNA topoisomerase IIβ and neural development. Science 287, 131–134 (2000).

    Article  CAS  PubMed  Google Scholar 

  184. Li, W. & Wang, J. C. Mammalian DNA topoisomerase IIIα is essential in early embryogenesis. Proc. Natl. Acad. Sci. USA 95, 1010–1013 (1998).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  185. Plank, J. L., Chu, S. H., Pohlhaus, J. R., Wilson-Sali, T. & Hsieh, T. S. Drosophila melanogaster topoisomerase IIIα preferentially relaxes a positively or negatively supercoiled bubble substrate and is essential during development. J. Biol. Chem. 280, 3564–3573 (2005).

    Article  CAS  PubMed  Google Scholar 

  186. Wu, J., Feng, L. & Hsieh, T. S. Drosophila topo IIIα is required for the maintenance of mitochondrial genome and male germ-line stem cells. Proc. Natl Acad. Sci. USA 107, 6228–6233 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  187. Martin, C. A. et al. Mutations in TOP3A cause a Bloom syndrome-like disorder. Am. J. Hum. Genet. 103, 456 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  188. Hudson, D. F. et al. Loss of RMI2 increases genome instability and causes a Bloom-like syndrome. PLoS Genet. 12, e1006483 (2016).

    Article  PubMed  PubMed Central  Google Scholar 

  189. Manthei, K. A. & Keck, J. L. The BLM dissolvasome in DNA replication and repair. Cell Mol. Life Sci. 70, 4067–4084 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  190. Lee, S. H., Siaw, G. E. L., Willcox, S., Griffith, J. D. & Hsieh, T. S. Synthesis and dissolution of hemicatenanes by type IA DNA topoisomerases. Proc. Natl Acad. Sci. USA 110, E3587–3594 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  191. Tubbs, A. & Nussenzweig, A. Endogenous DNA damage as a source of genomic instability in cancer. Cell 168, 644–656 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  192. Deans, A. J. & West, S. C. FANCM connects the genome instability disorders Bloom’s syndrome and fanconi anemia. Mol. Cell 36, 943–953 (2009).

    CAS  Google Scholar 

  193. Tsai, H. Z., Lin, R. K. & Hsieh, T. S. Drosophila mitochondrial topoisomerase IIIα affects the aging process via maintenance of mitochondrial function and genome integrity. J. Biomed. Sci. 23, 38 (2016).

    Article  PubMed  PubMed Central  Google Scholar 

  194. Kwan, K. Y. & Wang, J. C. Mice lacking DNA topoisomerase IIIβ develop to maturity but show a reduced mean lifespan. Proc. Natl Acad. Sci. USA 98, 5717–5721 (2001). This article demonstrates the viability of Top3b-knockout mice and their phenotype.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  195. Kwan, K. Y., Moens, P. B. & Wang, J. C. Infertility and aneuploidy in mice lacking a type IA DNA topoisomerase IIIβ. Proc. Natl Acad. Sci. USA 100, 2526–2531 (2003).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  196. Kwan, K. Y. et al. Development of autoimmunity in mice lacking DNA topoisomerase 3β. Proc. Natl Acad. Sci. USA 104, 9242–9247 (2007).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  197. Xu, D. et al. Top3β is an RNA topoisomerase that works with fragile X syndrome protein to promote synapse formation. Nat. Neurosci. 16, 1238–1247 (2013). This article provides the first evidence that TOP3B acts as a dual RNA and DNA topoisomerase.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  198. Joo, Y. et al. Topoisomerase 3β knockout mice show transcriptional and behavioural impairments associated with neurogenesis and synaptic plasticity. Nat. Commun. 11, 3143 (2020). This article reveals the importance of TOP3B for neuronal functions.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  199. Zhang, T. et al. Loss of TOP3B leads to increased R-loop formation and genome instability. Open Biol. 9, 190222 (2019).

    Article  PubMed  PubMed Central  Google Scholar 

  200. Ahmad, M. et al. Topoisomerase 3β is the major topoisomerase for mRNAs and linked to neurodevelopment and mental dysfunction. Nucleic Acids Res. 45, 2704–2713 (2017).

    CAS  PubMed  Google Scholar 

  201. Sun, Y., Saha, L. K., Saha, S., Jo, U. & Pommier, Y. Debulking of topoisomerase DNA–protein crosslinks (TOP-DPC) by the proteasome, non-proteasomal and non-proteolytic pathways. DNA Repair 94, 102926 (2020).

    Article  CAS  PubMed  Google Scholar 

  202. Dalla Rosa, I. et al. Mapping topoisomerase sites in mitochondrial DNA with a poisonous mitochondrial topoisomerase I (Top1mt). J. Biol. Chem. 289, 18595–18602 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  203. Stantial, N. et al. Trapped topoisomerase II initiates formation of de novo duplications via the nonhomologous end-joining pathway in yeast. Proc. Natl Acad. Sci. USA 117, 26876–26884 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  204. Pommier, Y. Drugging topoisomerases: lessons and challenges. ACS Chem. Biol. 8, 82–95 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  205. Hsiang, Y. H., Hertzberg, R., Hecht, S. & Liu, L. F. Camptothecin induces protein-linked DNA breaks via mammalian DNA topoisomerase I. J. Biol. Chem. 260, 14873–14878 (1985).

    Article  CAS  PubMed  Google Scholar 

  206. Pommier, Y. & Marchand, C. Interfacial inhibitors: targeting macromolecular complexes. Nat. Rev. Drug Discov. 11, 25–36 (2011). This review describes the landmark discovery of topoisomerase inhibitors and interfacial inhibitors and a paradigm for pharmacology.

    Article  PubMed  PubMed Central  Google Scholar 

  207. Pommier, Y. et al. Tyrosyl-DNA-phosphodiesterases (TDP1 and TDP2). DNA Repair. 19, 114–129 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  208. Comeaux, E. Q. & van Waardenburg, R. C. Tyrosyl-DNA phosphodiesterase I resolves both naturally and chemically induced DNA adducts and its potential as a therapeutic target. Drug Metab. Rev. 46, 494–507 (2014).

    Article  CAS  PubMed  Google Scholar 

  209. Pourquier, P. et al. Gemcitabine (2′,2′-difluoro-2′-deoxycytidine), an antimetabolite that poisons topoisomerase I. Clin. Cancer Res. 8, 2499–2504 (2002).

    CAS  PubMed  Google Scholar 

  210. Pourquier, P. et al. Induction of topoisomerase I cleavage complexes by 1-β-d-arabinofuranosylcytosine (ara-C) in vitro and in ara-C-treated cells. Proc. Natl Acad. Sci. USA 97, 1885–1890 (2000).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  211. Huang, S. N., Williams, J. S., Arana, M. E., Kunkel, T. A. & Pommier, Y. Topoisomerase I-mediated cleavage at unrepaired ribonucleotides generates DNA double-strand breaks. EMBO J. 36, 361–373 (2017).

    Article  CAS  PubMed  Google Scholar 

  212. Zimmermann, M. et al. CRISPR screens identify genomic ribonucleotides as a source of PARP-trapping lesions. Nature 559, 285–289 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  213. Pourquier, P. et al. Trapping of mammalian topoisomerase I and recombinations induced by damaged DNA containing nicks or gaps: importance of DNA end phosphorylation and camptothecin effects. J. Biol. Chem. 272, 26441–26447 (1997).

    Article  CAS  PubMed  Google Scholar 

  214. Pourquier, P. et al. Induction of reversible complexes between eukaryotic DNA topoisomerase I and DNA-containing oxidative base damages: 7,8-dihydro-8-oxoguanine and 5-hydroxycytosine. J. Biol. Chem. 274, 8516–8523 (1999).

    Article  CAS  PubMed  Google Scholar 

  215. Pourquier, P. et al. Effects of uracil incorporation, DNA mismatches, and abasic sites on cleavage and religation activities of mammalian topoisomerase I. J. Biol. Chem. 272, 7792–7796 (1997).

    Article  CAS  PubMed  Google Scholar 

  216. Pourquier, P. et al. Topoisomerase I-mediated cytotoxicity of N-methyl-N′-nitro-N-nitrosoguanidine: trapping of topoisomerase I by the O6-methylguanine. Cancer Res. 61, 53–58 (2001).

    CAS  PubMed  Google Scholar 

  217. Van Waardenburg, R. C. A. M. et al. Platinated DNA adducts enhance poisoning of DNA topoisomerase I by camptothecin. J. Biol. Chem. 279, 54502–54509 (2004).

    Article  PubMed  Google Scholar 

  218. Reijns, M. A. et al. Enzymatic removal of ribonucleotides from DNA is essential for mammalian genome integrity and development. Cell 149, 1008–1022 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  219. Pommier, Y., Jenkins, J., Kohlhagen, G. & Leteurtre, F. DNA recombinase activity of eukaryotic DNA topoisomerase I; effects of camptothecin and other inhibitors. Mutat. Res. 337, 135–145 (1995).

    Article  CAS  PubMed  Google Scholar 

  220. Shuman, S. Vaccinia DNA topoisomerase I promotes illegitimate recombination in Escherichia coli. Proc. Natl Acad. Sci. USA 86, 3489–3493 (1989).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  221. Strumberg, D. et al. Conversion of topoisomerase I cleavage complexes on the leading strand of ribosomal DNA into 5′-phosphorylated DNA double-strand breaks by replication runoff. Mol. Cell. Biol. 20, 3977–3987 (2000).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  222. Maede, Y. et al. Differential and common DNA repair pathways for topoisomerase I- and II-targeted drugs in a genetic DT40 repair cell screen panel. Mol. Cancer Ther. 13, 214–220 (2014).

    Article  CAS  PubMed  Google Scholar 

  223. Britton, S. et al. ATM antagonizes NHEJ proteins assembly and DNA-ends synapsis at single-ended DNA double strand breaks. Nucleic Acids Res. 48, 9710–9723 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  224. Nakamura, K. et al. Proteome dynamics at broken replication forks reveal a distinct ATM-directed repair response suppressing DNA double-strand break ubiquitination. Mol. Cell (2021).

  225. Whelan, D. et al. Super-resolution visualization of distinct stalled and broken replication fork structures. PLoS Genetics 16, e1009256 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  226. Balmus, G. et al. ATM orchestrates the DNA-damage response to counter toxic non-homologous end-joining at broken replication forks. Nat. Commun. 10, 87 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  227. Muslimovic, A., Nystrom, S., Gao, Y. & Hammarsten, O. Numerical analysis of etoposide induced DNA breaks. PLoS ONE 4, e5859 (2009).

    Article  PubMed  PubMed Central  Google Scholar 

  228. Cowell, I. G. & Austin, C. A. Mechanism of generation of therapy related leukemia in response to anti-topoisomerase II agents. Int. J. Environ. Res. Public Health 9, 2075–2091 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  229. Chen, G. L. et al. Nonintercalative antitumor drugs interfere with the breakage-reunion reaction of mammalian DNA topoisomerase II. J. Biol. Chem. 259, 13560–13566 (1984).

    Article  CAS  PubMed  Google Scholar 

  230. Tewey, K. M., Chen, G. L., Nelson, E. M. & Liu, L. F. Intercalative antitumor drugs interfere with the breakage-reunion reaction of mammalian DNA topoisomerase II. J. Biol. Chem. 259, 9182–9187 (1984).

    Article  CAS  PubMed  Google Scholar 

  231. Xiao, H. et al. The topoisomerase IIβ circular clamp arrests transcription and signals a 26S proteasome pathway. Proc. Natl Acad. Sci. USA 100, 3239–3244 (2003).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  232. Hsiang, Y. H. & Liu, L. F. Evidence for the reversibility of cellular DNA lesion induced by mammalian topoisomerase II poisons. J. Biol. Chem. 264, 9713–9715 (1989).

    Article  CAS  PubMed  Google Scholar 

  233. Karras, G. I. et al. HSP90 shapes the consequences of human genetic variation. Cell 168, 856–866.e12 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  234. Lai, C. H. et al. HSP-90 inhibitor ganetespib is synergistic with doxorubicin in small cell lung cancer. Oncogene 33, 4867–4876 (2014).

    Article  CAS  PubMed  Google Scholar 

  235. Kingma, P. S., Greider, C. A. & Osheroff, N. Spontaneous DNA lesions poison human topoisomerase IIα and stimulate cleavage proximal to leukemic 11q23 chromosomal breakpoints. Biochemistry 36, 5934–5939 (1997).

    Article  CAS  PubMed  Google Scholar 

  236. Cline, S. D., Jones, W. R., Stone, M. P. & Osheroff, N. DNA abasic lesions in a different light: solution structure of an endogenous topoisomerase II poison. Biochemistry 38, 15500–15507 (1999).

    Article  CAS  PubMed  Google Scholar 

  237. Bigioni, M. et al. Position-specific effects of base mismatch on mammalian topoisomerase II DNA cleaving activity. Biochemistry 35, 153–159 (1996).

    Article  CAS  PubMed  Google Scholar 

  238. Sabourin, M. & Osheroff, N. Sensitivity of human type II topoisomerases to DNA damage: stimulation of enzyme-mediated DNA cleavage by abasic, oxidized and alkylated lesions. Nucleic Acids Res. 28, 1947–1954 (2000).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  239. Szlachta, K. et al. Topoisomerase II contributes to DNA secondary structure-mediated double-stranded breaks. Nucleic Acids Res. 48, 6654–6671 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  240. Deweese, J. E. & Osheroff, N. The DNA cleavage reaction of topoisomerase II: wolf in sheep’s clothing. Nucleic Acids Res. 37, 738–748 (2009).

    Article  CAS  PubMed  Google Scholar 

  241. Lee, K. C. et al. MRE11 facilitates the removal of human topoisomerase II complexes from genomic DNA. Biol. Open. 1, 863–873 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  242. Hoa, N. N. et al. Mre11 is essential for the removal of lethal topoisomerase 2 covalent cleavage complexes. Mol. Cell 64, 580–592 (2016).

    Article  CAS  PubMed  Google Scholar 

  243. Sasanuma, H. et al. BRCA1 ensures genome integrity by eliminating estrogen-induced pathological topoisomerase II–DNA complexes. Proc. Natl Acad. Sci. USA 115, E10642–E10651 (2018). This article reveals the role of TOP2-DPCs as a source of breast cancers in patients deficient in BRCA1.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  244. Sun, Y. et al. Excision repair of topoisomerase DNA–protein crosslinks (TOP-DPC). DNA Repair. 89, 102837 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  245. Ruggiano, A. & Ramadan, K. DNA–protein crosslink proteases in genome stability. Commun. Biol. 4, 11 (2021).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  246. Patel, A. G. et al. Immunodetection of human topoisomerase I–DNA covalent complexes. Nucleic Acids Res. 44, 2816–2826 (2016).

    Article  PubMed  PubMed Central  Google Scholar 

  247. Maskey, R. S. et al. Spartan deficiency causes accumulation of topoisomerase 1 cleavage complexes and tumorigenesis. Nucleic Acids Res. 45, 4564–4576 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  248. Kojima, Y. et al. FAM111A protects replication forks from protein obstacles via its trypsin-like domain. Nat. Commun. 11, 1318 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  249. Sun, Y. et al. A conserved SUMO pathway repairs topoisomerase DNA–protein cross-links by engaging ubiquitin-mediated proteasomal degradation. Sci. Adv. 6, eaba6290 (2020). Together with Sun, Y. et al. (2020) (ref. 201), this review provides details and references for the redundant pathways that repair abortive TOPccs, complementary to the current Review.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  250. Katyal, S. et al. Aberrant topoisomerase-1 DNA lesions are pathogenic in neurodegenerative genome instability syndromes. Nat. Neurosci. 17, 813–21 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  251. Desai, S. D., Liu, L. F., Vazquez-Abad, D. & D’Arpa, P. Ubiquitin-dependent destruction of topoisomerase I is stimulated by the antitumor drug camptothecin. J. Biol. Chem. 272, 24159–24164 (1997).

    Article  CAS  PubMed  Google Scholar 

  252. Mao, Y., Desai, S. D., Ting, C. Y., Hwang, J. L. & Liu, L. F. 26S proteasome-mediated degradation of topoisomerase II cleavable complexes. J. Biol. Chem. 276, 40652–40658 (2001).

    Article  CAS  PubMed  Google Scholar 

  253. Tsuda, M. et al. Tyrosyl-DNA phosphodiesterase 2 (TDP2) repairs topoisomerase 1 DNA–protein crosslinks and 3-blocking lesions in the absence of tyrosyl-DNA phosphodiesterase 1 (TDP1). DNA Repair. 91-92, 102849–102849 (2020).

    Article  CAS  PubMed  Google Scholar 

  254. Kojima, Y. & Machida, Y. J. DNA–protein crosslinks from environmental exposure: mechanisms of formation and repair. Env. Mol. Mutagen. 61, 716–729 (2020).

    Article  CAS  Google Scholar 

  255. Dokshin, G. A. et al. GCNA interacts with spartan and topoisomerase II to regulate genome stability. Dev. Cell 52, 53–68.e6 (2020).

    Article  CAS  PubMed  Google Scholar 

  256. Serbyn, N. et al. The aspartic protease Ddi1 contributes to DNA–protein crosslink repair in yeast. Mol. Cell 77, 1066–1079 (2020).

    Article  CAS  PubMed  Google Scholar 

  257. Schellenberg, M. J. et al. ZATT (ZNF451)-mediated resolution of topoisomerase 2 DNA–protein cross-links. Science 357, 1412–1416 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  258. Schellenberg, M. J. et al. Ubiquitin stimulated reversal of topoisomerase 2 DNA–protein crosslinks by TDP2. Nucleic Acids Res. 48, 6310–6325 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  259. Riccio, A. A., Schellenberg, M. J. & Williams, R. S. Molecular mechanisms of topoisomerase 2 DNA–protein crosslink resolution. Cell Mol. Life Sci. 77, 81–91 (2020).

    Article  CAS  PubMed  Google Scholar 

  260. Yang, S.-W. et al. A eukaryotic enzyme that can disjoin dead-end covalent complexes between DNA and type I topoisomerases. Proc. Natl. Acad. Sci. USA 93, 11534–11539 (1996).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  261. Interthal, H., Chen, H. J. & Champoux, J. J. Human Tdp1 cleaves a broad spectrum of substrates including phosphoamide linkages. J. Biol. Chem. 280, 36518–36528 (2005).

    Article  CAS  PubMed  Google Scholar 

  262. Nitiss, K. C., Malik, M., He, X., White, S. W. & Nitiss, J. L. Tyrosyl-DNA phosphodiesterase (Tdp1) participates in the repair of Top2-mediated DNA damage. Proc. Natl Acad. Sci. USA 103, 8953–8958 (2006).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  263. Gao, R., Huang, S. Y., Marchand, C. & Pommier, Y. Biochemical characterization of human tyrosyl-DNA phosphodiesterase 2 (TDP2/TTRAP): a Mg2+/Mn2+-dependent phosphodiesterase specific for the repair of topoisomerase cleavage complexes. J. Biol. Chem. 287, 30842–30852 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  264. Ledesma, F. C., El Khamisy, S. F., Zuma, M. C., Osborn, K. & Caldecott, K. W. A human 5′-tyrosyl DNA phosphodiesterase that repairs topoisomerase-mediated DNA damage. Nature 461, 674–678 (2009).

    Article  CAS  Google Scholar 

  265. Liao, C. et al. UCHL3 regulates topoisomerase-induced chromosomal break repair by controlling TDP1 proteostasis. Cell Rep. 23, 3352–3365 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  266. Hudson, J. J., Chiang, S. C., Wells, O. S., Rookyard, C. & El-Khamisy, S. F. SUMO modification of the neuroprotective protein TDP1 facilitates chromosomal single-strand break repair. Nat. Commun. 3, 733 (2012).

    Article  PubMed  Google Scholar 

  267. Das, B. B. et al. Optimal function of the DNA repair enzyme TDP1 requires its phosphorylation by ATM and/or DNA-PK. EMBO J. 28, 3667–3680 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  268. Das, B. B. et al. PARP1–TDP1 coupling for the repair of topoisomerase I-induced DNA damage. Nucleic Acids Res. 42, 4435–4449 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  269. Sun, Y. et al. PARylation prevents the proteasomal degradation of topoisomerase I DNA–protein crosslinks and induces their deubiquitylation. Nat. Commun. 12, 5010 (2021).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  270. Lee, K. C., Bramley, R. L., Cowell, I. G., Jackson, G. H. & Austin, C. A. Proteasomal inhibition potentiates drugs targeting DNA topoisomerase II. Biochem. Pharmacol. 103, 29–39 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  271. Pardo, B., Moriel-Carretero, M., Vicat, T., Aguilera, A. & Pasero, P. Homologous recombination and Mus81 promote replication completion in response to replication fork blockage. EMBO Rep. 21, e49367 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  272. Regairaz, M. et al. Mus81-mediated DNA cleavage resolves replication forks stalled by topoisomerase I–DNA complexes. J. Cell Biol. 195, 739–749 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  273. Ganguly, A. et al. Tdp1 processes chromate-induced single-strand DNA breaks that collapse replication forks. PLoS Genet. 14, e1007595 (2018).

    Article  PubMed  PubMed Central  Google Scholar 

  274. Kikuchi, K. et al. Structure-specific endonucleases Xpf and Mus81 play overlapping but essential roles in DNA repair by homologous recombination. Cancer Res. 73, 4362–4371 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  275. Zhang, Y. W. et al. Poly(ADP-ribose) polymerase and XPF-ERCC1 participate in distinct pathways for the repair of topoisomerase I-induced DNA damage in mammalian cells. Nucleic Acids Res. 39, 3607–3620 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  276. Sousa, F. G. et al. Alterations of DNA repair genes in the NCI-60 cell lines and their predictive value for anticancer drug activity. DNA Repair. 28, 107–115 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  277. Alagoz, M., Chiang, S.-C., Sharma, A. & El-Khamisy, S. F. ATM deficiency results in accumulation of DNA–topoisomerase I covalent intermediates in neural cells. PLoS ONE 8, e58239 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  278. Neale, M. J., Pan, J. & Keeney, S. Endonucleolytic processing of covalent protein-linked DNA double-strand breaks. Nature 436, 1053–1057 (2005).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  279. Sunter, N. J., Cowell, I. G., Willmore, E., Watters, G. P. & Austin, C. A. Role of topoisomerase II β in DNA damage response following IR and etoposide. J. Nucleic Acids 2010, 710589–710589 (2010).

    Article  PubMed  PubMed Central  Google Scholar 

  280. Gomez-Herreros, F. et al. 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 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  281. Akagawa, R. et al. UBC13-mediated ubiquitin signaling promotes removal of blocking adducts from DNA double-strand breaks. iScience 23, 101027–101027 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  282. Álvarez-Quilón, A. et al. ATM specifically mediates repair of double-strand breaks with blocked DNA ends. Nat. Commun. 5, 3347–3347 (2014).

    Article  PubMed  Google Scholar 

  283. Quennet, V., Beucher, A., Barton, O., Takeda, S. & Löbrich, M. CtIP and MRN promote non-homologous end-joining of etoposide-induced DNA double-strand breaks in G1. Nucleic Acids Res. 39, 2144–2152 (2011).

    Article  CAS  PubMed  Google Scholar 

  284. Zagnoli-Vieira, G. & Caldecott, K. W. TDP2, TOP2, and SUMO: what is ZATT about? Cell Res. 27, 1405–1406 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  285. Zhao, G. Y. et al. A critical role for the ubiquitin-conjugating enzyme Ubc13 in initiating homologous recombination. Mol. Cell 25, 663–675 (2007).

    Article  CAS  PubMed  Google Scholar 

  286. Allan, J. M. & Travis, L. B. Mechanisms of therapy-related carcinogenesis. Nat. Rev. Cancer 5, 943–955 (2005).

    Article  CAS  PubMed  Google Scholar 

  287. Leone, G., Fianchi, L., Pagano, L. & Voso, M. T. Incidence and susceptibility to therapy-related myeloid neoplasms. Chem. Biol. Interact. 184, 39–45 (2010).

    Article  CAS  PubMed  Google Scholar 

  288. Kayser, S. et al. The impact of therapy-related acute myeloid leukemia (AML) on outcome in 2853 adult patients with newly diagnosed AML. Blood 117, 2137–2145 (2011).

    Article  CAS  PubMed  Google Scholar 

  289. Mauritzson, N. et al. Pooled analysis of clinical and cytogenetic features in treatment-related and de novo adult acute myeloid leukemia and myelodysplastic syndromes based on a consecutive series of 761 patients analyzed 1976–1993 and on 5098 unselected cases reported in the literature 1974–2001. Leukemia 16, 2366–2378 (2002).

    Article  CAS  PubMed  Google Scholar 

  290. Chen, W. et al. Malignant transformation initiated by Mll-AF9: gene dosage and critical target cells. Cancer Cell 13, 432–440 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  291. Krivtsov, A. V. et al. Transformation from committed progenitor to leukaemia stem cell initiated by MLL-AF9. Nature 442, 818–822 (2006).

    Article  CAS  PubMed  Google Scholar 

  292. Cozzio, A. et al. Similar MLL-associated leukemias arising from self-renewing stem cells and short-lived myeloid progenitors. Genes Dev. 17, 3029–3035 (2003).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  293. Grimwade, D. & Enver, T. Acute promyelocytic leukemia: where does it stem from? Leukemia 18, 375–384 (2004).

    Article  CAS  PubMed  Google Scholar 

  294. Seita, J. & Weissman, I. L. Hematopoietic stem cell: self-renewal versus differentiation. Wiley Interdiscip. Rev. 2, 640–653 (2010).

    CAS  Google Scholar 

  295. Rowley, J. D. & Olney, H. J. International workshop on the relationship of prior therapy to balanced chromosome aberrations in therapy-related myelodysplastic syndromes and acute leukemia: overview report. Genes Chromosomes Cancer 33, 331–345 (2002).

    Article  PubMed  Google Scholar 

  296. Lieber, M. R. NHEJ and its backup pathways: relation to chromosomal translocations. Nat. Struct. Mol. Biol. 17, 393–395 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  297. Lieber, M. R. Mechanisms of human lymphoid chromosomal translocations. Nat. Rev. Cancer 16, 387–398 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  298. Meyer, C. et al. The MLL recombinome of acute leukemias in 2013. Leukemia 27, 2165–2176 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  299. Cowell, I. G. et al. Model for MLL translocations in therapy-related leukemia involving topoisomerase IIβ-mediated DNA strand breaks and gene proximity. Proc. Natl Acad. Sci. USA 109, 8989–8994 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  300. Felix, C. A. Secondary leukemias induced by topoisomerase-targeted drugs. Biochim. Biophys. Acta 1400, 233–255 (1998).

    Article  CAS  PubMed  Google Scholar 

  301. Libura, J., Slater, D. J., Felix, C. A. & Richardson, C. Therapy-related acute myeloid leukemia-like MLL rearrangements are induced by etoposide in primary human CD34+ cells and remain stable after clonal expansion. Blood 105, 2124–2131 (2005).

    Article  CAS  PubMed  Google Scholar 

  302. Hasan, S. K. et al. Molecular analysis of t(15;17) genomic breakpoints in secondary acute promyelocytic leukemia arising after treatment of multiple sclerosis. Blood 112, 3383–3390 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  303. Hasan, S. K. et al. Analysis of t(15;17) chromosomal breakpoint sequences in therapy-related versus de novo acute promyelocytic leukemia: association of DNA breaks with specific DNA motifs at PML and RARA loci. Genes Chromosomes Cancer 49, 726–732 (2010).

    Article  CAS  PubMed  Google Scholar 

  304. Mays, A. N. et al. Evidence for direct involvement of epirubicin in the formation of chromosomal translocations in t(15;17) therapy-related acute promyelocytic leukemia. Blood 115, 326–330 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  305. Mistry, A. R. et al. DNA topoisomerase II in therapy-related acute promyelocytic leukemia. N. Engl. J. Med. 352, 1529–1538 (2005).

    Article  CAS  PubMed  Google Scholar 

  306. Uuskula-Reimand, L. & Wilson, M. D. Break check: transcription-driven topoisomerase II collisions near chromatin loop anchors are hotspots for DNA damage and translocations. Mol. Cell 75, 203–205 (2019).

    Article  CAS  PubMed  Google Scholar 

  307. Scharf, S. et al. Transcription linked to recombination: a gene-internal promoter coincides with the recombination hot spot II of the human MLL gene. Oncogene 26, 1361–1371 (2007).

    Article  CAS  PubMed  Google Scholar 

  308. Smith, K. A., Cowell, I. G., Zhang, Y., Sondka, Z. & Austin, C. A. The role of topoisomerase IIβ on breakage and proximity of RUNX1 to partner alleles RUNX1T1 and EVI1. Genes Chromosomes Cancer 53, 117–128 (2014).

    Article  CAS  PubMed  Google Scholar 

  309. Ploski, J. E. & Aplan, P. D. Characterization of DNA fragmentation events caused by genotoxic and non-genotoxic agents. Mutat. Res. 473, 169–180 (2001).

    Article  CAS  PubMed  Google Scholar 

  310. Aplan, P. D., Chervinsky, D. S., Stanulla, M. & Burhans, W. C. Site-specific DNA cleavage within the MLL breakpoint cluster region induced by topoisomerase II inhibitors. Blood 87, 2649–2658 (1996).

    Article  CAS  PubMed  Google Scholar 

  311. Ng, A., Taylor, G. M. & Eden, O. B. Genotoxicity of etoposide: greater susceptibility of MLL than other target genes. Cancer Genet. Cytogenet. 164, 164–167 (2006).

    Article  CAS  PubMed  Google Scholar 

  312. Strissel, P. L., Strick, R., Rowley, J. D. & Zeleznik-Le, N. J. An in vivo topoisomerase II cleavage site and a DNase I hypersensitive site colocalize near exon 9 in the MLL breakpoint cluster region. Blood 92, 3793–3803 (1998).

    Article  CAS  PubMed  Google Scholar 

  313. Tang, H. L. et al. Cell survival, DNA damage, and oncogenic transformation after a transient and reversible apoptotic response. Mol. Biol. Cell 23, 2240–2252 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  314. Betti, C. J., Villalobos, M. J., Diaz, M. O. & Vaughan, A. T. M. Apoptotic triggers initiate translocations within the MLL gene involving the nonhomologous end joining repair system. Cancer Res. 61, 4550–4555 (2001).

    CAS  PubMed  Google Scholar 

  315. Betti, C. J. et al. Cleavage of the MLL gene by activators of apoptosis is independent of topoisomerase II activity. Leukemia 19, 2289–2295 (2005).

    Article  CAS  PubMed  Google Scholar 

  316. Sim, S.-P. & Liu, L. F. Nucleolytic cleavage of the mixed lineage leukemia breakpoint cluster region during apoptosis. J. Biol. Chem. 276, 31590–31595 (2001).

    Article  CAS  PubMed  Google Scholar 

  317. Hars, E. S., Lyu, Y. L., Lin, C.-P. & Liu, L. F. Role of apoptotic nuclease caspase-activated DNase in etoposide-induced treatment-related acute myelogenous leukemia. Cancer Res. 66, 8975–8979 (2006).

    Article  CAS  PubMed  Google Scholar 

  318. Wang, J. C. Cellular roles of DNA topoisomerases: a molecular perspective. Nat. Rev. Mol. Cell Biol. 3, 430–440 (2002). This work is a landmark review for topoisomerase discovery and biology by a pioneer in the topoisomerase field.

    Article  CAS  PubMed  Google Scholar 

  319. Lee, M. P., Brown, S. D., Chen, A. & Hsieh, T. S. DNA topoisomerase I is essential in Drosophila melanogaster. Proc. Natl Acad. Sci. USA 90, 6656–6660 (1993).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  320. Zhang, C. X., Chen, A. D., Gettel, N. J. & Hsieh, T. S. Essential functions of DNA topoisomerase I in Drosophila melanogaster. Dev. Biol. 222, 27–40 (2000).

    Article  CAS  PubMed  Google Scholar 

  321. Brill, S. J., DiNardo, S., Voelkel-Meiman, K. & Sternglanz, R. Need for DNA topoisomerase activity as a swivel for DNA replication for transcription of ribosomal RNA. Nature 326, 414–416 (1987).

    Article  CAS  PubMed  Google Scholar 

  322. Zhang, H., Seol, Y., Agama, K., Neuman, K. C. & Pommier, Y. Distribution bias and biochemical characterization of TOP1MT single nucleotide variants. Sci. Rep. 7, 8614 (2017).

    Article  PubMed  PubMed Central  Google Scholar 

  323. Chiang, S.-C. et al. Mitochondrial protein-linked DNA breaks perturb mitochondrial gene transcription and trigger free radical-induced DNA damage. Sci. Adv. 3, e1602506 (2017).

    Article  PubMed  PubMed Central  Google Scholar 

  324. Sobek, S. et al. Negative regulation of mitochondrial transcription by mitochondrial topoisomerase I. Nucleic Acids Res. 41, 9848–9857 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  325. Douarre, C. et al. Mitochondrial topoisomerase I is critical for mitochondrial integrity and cellular energy metabolism. PLoS ONE 7, e41094 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  326. Khiati, S. et al. Mitochondrial topoisomerase I (Top1mt) is a novel limiting factor of doxorubicin cardiotoxicity. Clin. Cancer Res. 20, 4873–4881 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  327. Morimoto, H. et al. An interplay of NOX1-derived ROS and oxygen determines the spermatogonial stem cell self-renewal efficiency under hypoxia. Genes Dev. 35, 250–260 (2021).

  328. Hoffmann, A., Heck, M. M., Bordwell, B. J., Rothfield, N. F. & Earnshaw, W. C. Human autoantibody to topoisomerase II. Exp. Cell Res. 180, 409–418 (1989).

    Article  CAS  PubMed  Google Scholar 

  329. Imai, H. et al. Autoantibody to DNA topoisomerase II in primary liver cancer. Clin. Cancer Res. 1, 417–424 (1995).

    CAS  PubMed  Google Scholar 

  330. Akimitsu, N. et al. Induction of apoptosis by depletion of DNA topoisomerase IIα in mammalian cells. Biochem. Biophys. Res. Commun. 307, 301–307 (2003).

    Article  CAS  PubMed  Google Scholar 

  331. Broderick, L. et al. Mutations in topoisomerase IIβ result in a B cell immunodeficiency. Nat. Commun. 10, 3644 (2019).

    Article  PubMed  PubMed Central  Google Scholar 

  332. Hiraide, T. et al. A de novo TOP2B variant associated with global developmental delay and autism spectrum disorder. Mol. Genet. Genom. Med. 8, e1145 (2020).

    CAS  Google Scholar 

  333. Lam, C. W., Yeung, W. L. & Law, C. Y. Global developmental delay and intellectual disability associated with a de novo TOP2B mutation. Clin. Chim. Acta 469, 63–68 (2017).

    Article  CAS  PubMed  Google Scholar 

  334. Dereuddre, S., Delaporte, C. & Jacquemin-Sablon, A. Role of topoisomerase IIβ in the resistance of 9-OH-ellipticine-resistant Chinese hamster fibroblasts to topoisomerase II inhibitors. Cancer Res. 57, 4301–4308 (1997).

    CAS  PubMed  Google Scholar 

  335. Edmond, M., Hanley, O. & Philippidou, P. Topoisomerase IIβ selectively regulates motor neuron identity and peripheral connectivity through Hox/Pbx-dependent transcriptional programs. eNeuro 4, ENEURO.0404-17.2017 (2017).

    Article  PubMed  PubMed Central  Google Scholar 

  336. Zhang, Y. L. et al. TOP2βeta is essential for ovarian follicles that are hypersensitive to chemotherapeutic drugs. Mol. Endocrinol. 27, 1678–1691 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  337. Navabpour, S., Rogers, J., McFadden, T. & Jarome, T. J. DNA double-strand breaks are a critical regulator of fear memory reconsolidation. Int. J. Mol. Sci. 21, 8995 (2020).

    Article  CAS  PubMed Central  Google Scholar 

  338. Chapman, J., Ng, Y. S. & Nicholls, T. J. The maintenance of mitochondrial DNA integrity and dynamics by mitochondrial membranes. Life 10, 164 (2020).

    Article  CAS  PubMed Central  Google Scholar 

  339. Herbaux, C. et al. TOP3A, a new partner gene fused to MLL in an adult patient with de novo acute myeloid leukaemia. Br. J. Haematol. 157, 128–131 (2012).

    Article  PubMed  Google Scholar 

  340. Wallis, J. W., Chrebet, G., Brodsky, G., Rolfe, M. & Rothstein, R. A hyper-recombination mutation in S. cerevisiae identifies a novel eukaryotic topoisomerase. Cell 58, 409–419 (1989).

    Article  CAS  PubMed  Google Scholar 

  341. Kim, R. A. & Wang, J. C. Identification of the yeast TOP3 gene product as a single strand-specific DNA topoisomerase. J. Biol. Chem. 267, 17178–17185 (1992).

    Article  CAS  PubMed  Google Scholar 

  342. Gangloff, S., MacDonald, J. P., Bendixen, C., Arthur, L. & Rothstein, R. The yeast type 1 topoisomerase Top3 interacts with Sgs1, a DNA helicase homolog: a potential eukaryotic reverse gyrase. Mol. Cell. Biol. 14, 8391–8398 (1994).

    CAS  PubMed  PubMed Central  Google Scholar 

  343. Rosato, M. et al. Combined cellomics and proteomics analysis reveals shared neuronal morphology and molecular pathway phenotypes for multiple schizophrenia risk genes. Mol. Psychiatry 26, 784–799 (2021).

    Article  PubMed  Google Scholar 

  344. El-Khamisy, S. F. et al. Defective DNA single-strand break repair in spinocerebellar ataxia with axonal neuropathy-1. Nature 434, 108–113 (2005).

    Article  CAS  PubMed  Google Scholar 

  345. Takashima, H. et al. Mutation of TDP1, encoding a topoisomerase I-dependent DNA damage repair enzyme, in spinocerebellar ataxia with axonal neuropathy. Nat. Genet. 32, 267–272 (2002).

    Article  CAS  PubMed  Google Scholar 

  346. Murai, J. et al. Tyrosyl-DNA phosphodiesterase 1 (TDP1) repairs DNA damage induced by topoisomerases I and II and base alkylation in vertebrate cells. J. Biol. Chem. 287, 12848–12857 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  347. Katyal, S. et al. TDP1 facilitates chromosomal single-strand break repair in neurons and is neuroprotective in vivo. EMBO J. 26, 4720–4731 (2007).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  348. Guo, D., Dexheimer, T. S., Pommier, Y. & Nash, H. A. Neuroprotection and repair of 3′-blocking DNA ends by glaikit (gkt) encoding Drosophila tyrosyl-DNA phosphodiesterase 1 (TDP1). Proc. Natl Acad. Sci. USA 111, 15816–15820 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  349. Pouliot, J. J., Robertson, C. A. & Nash, H. A. Pathways for repair of topoisomerase I covalent complexes in Saccharomyces cerevisiae. Genes. Cell 6, 677–687 (2001).

    Article  CAS  Google Scholar 

  350. Pouliot, J. J., Yao, K. C., Robertson, C. A. & Nash, H. A. Yeast gene for a Tyr-DNA phosphodiesterase that repairs topo I covalent complexes. Science 286, 552–555 (1999).

    Article  CAS  PubMed  Google Scholar 

  351. Zoghi, S. et al. A novel non-sense mutation in TDP2 causes spinocerebellar ataxia autosomal recessive 23 accompanied by bilateral upward gaze; report of a case and review of the literature. Eur. J. Med. Genet. 64, 104348 (2021).

    Article  PubMed  Google Scholar 

  352. Zeng, Z. et al. TDP2 promotes repair of topoisomerase I-mediated DNA damage in the absence of TDP1. Nucleic Acids Res. 40, 8371–8380 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  353. Maciejewski, S., Ullmer, W. & Semler, B. L. VPg unlinkase/TDP2 in cardiovirus infected cells: re-localization and proteolytic cleavage. Virology 516, 139–146 (2018).

    Article  CAS  PubMed  Google Scholar 

  354. Virgen-Slane, R. et al. An RNA virus hijacks an incognito function of a DNA repair enzyme. Proc. Natl Acad. Sci. USA 109, 14634–14639 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  355. Gomez-Herreros, F. et al. TDP2 protects transcription from abortive topoisomerase activity and is required for normal neural function. Nat. Genet. 46, 516–521 (2014).

    Article  CAS  PubMed  Google Scholar 

  356. Alvarez-Quilon, A. et al. Endogenous topoisomerase II-mediated DNA breaks drive thymic cancer predisposition linked to ATM deficiency. Nat. Commun. 11, 910 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  357. Bakx, J. A. M. et al. Duplex DNA and BLM regulate gate opening by the human TopoIIIα-RMI1-RMI2 complex. Nat. Commun. 13, 584 (2022).

  358. Long, B. H., Musial, S. T. & Brattain, M. G. Single- and double-strand DNA breakage and repair in human lung adenocarcinoma cells exposed to etoposide and teniposide. Cancer Res. 45, 3106–3112 (1985).

    CAS  PubMed  Google Scholar 

  359. Pommier, Y., Leo, E., Zhang, H. & Marchand, C. DNA topoisomerases and their poisoning by anticancer and antibacterial drugs. Chem. Biol. 17, 421–433 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  360. Pourquier, P. & Pommier, Y. Topoisomerase I-mediated DNA damage. Adv. Cancer Res. 80, 189–216 (2001).

    Article  CAS  PubMed  Google Scholar 

  361. Alexandrov, L. B. et al. The repertoire of mutational signatures in human cancer. Nature 578, 94–101 (2020).

  362. Reijns, M. A. M. et al. Signatures of TOP1 transcription-associated mutagenesis in cancer and germline. Nature https://doi.org/10.1038/s41586-022-04403-y (2022).

  363. Berti, M. et al. Human RECQ1 promotes restart of replication forks reversed by DNA topoisomerase I inhibition. Nat. Struct. Mol. Biol. 20, 347–354 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  364. Boot, A. et al. Recurrent mutations in topoisomerase IIα cause a previously undescribed mutator phenotype in human cancers. Proc. Natl Acad. Sci. USA 119, e2114024119, (2022).

  365. Zhang, H. F. et al. Cullin 3 promotes proteasomal degradation of the topoisomerase I–DNA covalent complex. Cancer Res. 64, 1114–1121 (2004).

    Article  CAS  PubMed  Google Scholar 

  366. Kerzendorfer, C. et al. Mutations in Cullin 4B result in a human syndrome associated with increased camptothecin-induced topoisomerase I-dependent DNA breaks. Hum. Mol. Genet. 19, 1324–1334 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  367. Falkenberg, M. & Gustafsson, C. M. Mammalian mitochondrial DNA replication and mechanisms of deletion formation. Crit. Rev. Biochem. Mol. Biol. 55, 509–524 (2020).

  368. Wang, Y., Lyu, Y. L. & Wang, J. C. Dual localization of human DNA topoisomerase IIIα to mitochondria and nucleus. Proc. Natl Acad. Sci. USA 99, 12114–12119 (2002).

  369. Zhang, H. et al. Human mitochondrial topoisomerase I. Proc. Natl Acad. Sci. USA 98, 10608–10613 (2001).

  370. Zhang, H., Meng, L. H., Zimonjic, D. B., Popescu, N. C. & Pommier, Y. Thirteen-exon-motif signature for vertebrate nuclear and mitochondrial type IB topoisomerases. Nucleic Acids Res. 32, 2087–2092 (2004).

  371. Low, R. L., Orton, S. & Friedman, D. B. A truncated form of DNA topoisomerase IIβ associates with the mtDNA genome in mammalian mitochondria. Eur. J. Biochem. 270, 4173–4186 (2003).

  372. Das, B. B., Dexheimer, T. S., Maddali, K. & Pommier, Y. Role of tyrosyl-DNA phosphodiesterase (TDP1) in mitochondria. Proc. Natl Acad. Sci. USA 107, 19790–19795 (2010).

  373. Huang, S. N. et al. Mitochondrial tyrosyl-DNA phosphodiesterase 2 and its TDP2(S) short isoform. EMBO Rep 19, e42139 (2018).

Download references

Acknowledgements

The authors’ understanding of topoisomerases stems from the dedication and constant discussions of members of the Laboratory of Molecular Pharmacology, Developmental Therapeutics Branch, Center for Cancer Research, the Intramural Program of the National Cancer Institute, National Institutes of Health (NIH) (Z01-BC-006161). They thank K. W. Kohn, who first hypothesized that doxorubicin and other anthracycline anticancer drugs act by forming protein-associated DNA breaks, which are now referred to as TOPccs. Y.P. and A.N. are supported by the Center for Cancer Research of the US National Cancer Institute (CCR-NCI) (Z01 BC 006161 and Z01 BC 006150 to Y.P.). This work was also supported by the Japan Society for the Promotion of Science (JSPS) KAKENHI (16H12595 and 16H06306) (to S.T.) and JSPS Core-to-Core Program, A. Advanced Research Networks (to S.T.). The authors thank X. Yang, postdoctoral fellow in the Pommier group, for his contribution to Supplementary Fig. 1. They are also grateful to the three reviewers who provided multiple suggestions and careful corrections.

Author information

Authors and Affiliations

Authors

Contributions

The authors contributed equally to all aspects of the article.

Corresponding author

Correspondence to Yves Pommier.

Ethics declarations

Competing interests

The authors declare no competing interests.

Peer review

Peer review information

Nature Reviews Molecular Cell Biology thanks the anonymous reviewers for their contribution to the peer review of this work.

Additional information

Publisher’s note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Supplementary information

Glossary

Tyrosyl-DNA phosphodiesterases

(TDPs). Referring to TDP1 and TDP2, enzymes that excise topoisomerase–DNA crosslinks by hydrolysing 3′-phosphodiester and 5′-phosphodiester bonds, respectively. TDP2 also excises TOP3–RNA crosslinks.

DNA supercoiling

The amount of DNA twist, which is the number of crossovers of the two strands across each other; writhe is a measure of the double helix winding around itself. Positive DNA supercoiling (Sc+) is defined by increased twist and/or writhe; negative DNA supercoiling (Sc) is the opposite.

Hemicatenanes

Crossovers of two strands of DNA originating from different DNA molecules.

Holliday junctions

Branched DNA structures consisting of four double-stranded arms joined together. Double Holliday junctions form hemicatenanes that are resolved by the Bloom syndrome protein (BLM)–topoisomerase 3A (TOP3A)–RecQ-mediated genome instability proteins (RMI1/2) (BTR) dissolvasome complex.

Topologically associating domains

(TADs). In interphase chromosomes, genomic regions in which interactions between loci are enriched compared with interactions with loci outside the domain.

Structural maintenance of chromosomes

(SMC). SMC complexes are ATPase complexes that tether and organize chromatin by forming chromatin loops and ensuring sister chromatid cohesion.

Homology-directed repair

(HDR). A replication-associated DNA double-strand break (DSB) repair pathway that uses a homologous sequence as a template for resynthesizing a missing DNA segment. Its classical form is homologous recombination and its common effector is RAD51.

γH2AX

Histone H2AX phosphorylated at Ser139; a sensitive biomarker of DNA double-strand breaks (DSBs).

TOPcc trapping

Stabilization of topoisomerase cleavage complexes (TOPccs) by inhibition of DNA end rejoining by a drug molecule bound at the interface of the DNA break and the enzyme or by DNA lesions that misalign the broken DNA ends, thereby preventing the release of the topoisomerase.

Non-homologous end joining

(NHEJ). The prominent DNA double-strand break (DSB) repair pathway, which rapidly joins adjacent DNA ends. Its main effectors are Ku70–Ku80 and DNA-dependent protein kinase (DNA-PK).

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Pommier, Y., Nussenzweig, A., Takeda, S. et al. Human topoisomerases and their roles in genome stability and organization. Nat Rev Mol Cell Biol 23, 407–427 (2022). https://doi.org/10.1038/s41580-022-00452-3

Download citation

  • Accepted:

  • Published:

  • Issue date:

  • DOI: https://doi.org/10.1038/s41580-022-00452-3

This article is cited by

Search

Quick links

Nature Briefing

Sign up for the Nature Briefing newsletter — what matters in science, free to your inbox daily.

Get the most important science stories of the day, free in your inbox. Sign up for Nature Briefing