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.

  • Article
  • Published:

A slipped-CAG DNA-binding small molecule induces trinucleotide-repeat contractions in vivo

Abstract

In many repeat diseases, such as Huntington’s disease (HD), ongoing repeat expansions in affected tissues contribute to disease onset, progression and severity. Inducing contractions of expanded repeats by exogenous agents is not yet possible. Traditional approaches would target proteins driving repeat mutations. Here we report a compound, naphthyridine-azaquinolone (NA), that specifically binds slipped-CAG DNA intermediates of expansion mutations, a previously unsuspected target. NA efficiently induces repeat contractions in HD patient cells as well as en masse contractions in medium spiny neurons of HD mouse striatum. Contractions are specific for the expanded allele, independently of DNA replication, require transcription across the coding CTG strand and arise by blocking repair of CAG slip-outs. NA-induced contractions depend on active expansions driven by MutSβ. NA injections in HD mouse striatum reduce mutant HTT protein aggregates, a biomarker of HD pathogenesis and severity. Repeat-structure-specific DNA ligands are a novel avenue to contract expanded repeats.

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: NA binds to long CAG slip-outs.
Fig. 2: NA specifically inhibits repair of long CAG slip-outs by human (HeLa) cell extracts.
Fig. 3: NA cellular distribution, non-toxicity and effects on repeat instability in HD patient cells.
Fig. 4: NA induces CAG contractions independent of proliferation, dependent on rCAG transcription.
Fig. 5: NA affects interaction of DNA repair proteins on long CAG slip-outs.
Fig. 6: NA induces CAG contractions in R6/2 mouse striatum.
Fig. 7: NA induces a reduction in mHTT aggregates in R6/2 mice.
Fig. 8: Schematic of the plausible mechanisms through which NA may induce contractions of expanded CAG tracts.

Similar content being viewed by others

Data availability

Raw sequencing data have been deposited at the Sequence Read Archive (SRA) under accession numbers SRR10532698, SRR10532697, SRR10532700 and SRR10532699. Source data for Figs. 18 and Extended Data Figs. 1 and 4 are provided online.

References

  1. Kennedy, L. et al. Dramatic tissue-specific mutation length increases are an early molecular event in Huntington disease pathogenesis. Hum. Mol. Genet. 12, 3359–3367 (2003).

    Article  CAS  PubMed  Google Scholar 

  2. Lopez Castel, A., Cleary, J. D. & Pearson, C. E. Repeat instability as the basis for human diseases and as a potential target for therapy. Nat. Rev. Mol. Cell Biol. 11, 165–170 (2010).

    Article  CAS  PubMed  Google Scholar 

  3. Mirkin, S. M. Expandable DNA repeats and human disease. Nature 447, 932–940 (2007).

    Article  CAS  PubMed  Google Scholar 

  4. Pearson, C. E., Nichol Edamura, K. & Cleary, J. D. Repeat instability: mechanisms of dynamic mutations. Nat. Rev. Genet. 6, 729–742 (2005).

    Article  CAS  PubMed  Google Scholar 

  5. Sathasivam, K., Amaechi, I., Mangiarini, L. & Bates, G. Identification of an HD patient with a (CAG)180 repeat expansion and the propagation of highly expanded CAG repeats in lambda phage. Hum. Genet. 99, 692–695 (1997).

    Article  CAS  PubMed  Google Scholar 

  6. Morales, F. et al. Somatic instability of the expanded CTG triplet repeat in myotonic dystrophy type 1 is a heritable quantitative trait and modifier of disease severity. Hum. Mol. Genet. 21, 3558–3567 (2012).

    Article  CAS  PubMed  Google Scholar 

  7. Swami, M. et al. Somatic expansion of the Huntington’s disease CAG repeat in the brain is associated with an earlier age of disease onset. Hum. Mol. Genet. 18, 3039–3047 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  8. Bettencourt, C. et al. DNA repair pathways underlie a common genetic mechanism modulating onset in polyglutamine diseases. Ann. Neurol. 79, 983–990 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  9. Genetic Modifiers of Huntington’s Disease Consortium Identification of genetic factors that modify clinical onset of Huntington’s disease. Cell 162, 516–526 (2015).

    Article  CAS  Google Scholar 

  10. Genetic Modifiers of Huntington’s Disease (GeM-HD) Consortium. CAG repeat not polyglutamine length determines timing of Huntington’s disease onset. Cell 178, 887–900.e14 (2019).

  11. Hensman Moss, D. J. et al. Identification of genetic variants associated with Huntington’s disease progression: a genome-wide association study. Lancet Neurol. 16, 701–711 (2017).

    Article  Google Scholar 

  12. Gusella, J. F. & MacDonald, M. E. Molecular genetics: unmasking polyglutamine triggers in neurodegenerative disease. Nat. Rev. Neurosci. 1, 109–115 (2000).

    Article  CAS  PubMed  Google Scholar 

  13. Rosenblatt, A. et al. Age, CAG repeat length, and clinical progression in Huntington’s disease. Mov. Disord. 27, 272–276 (2012).

    Article  PubMed  Google Scholar 

  14. Axford, M. M. et al. Detection of slipped-DNAs at the trinucleotide repeats of the myotonic dystrophy type I disease locus in patient tissues. PLoS Genet. 9, e1003866 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  15. Goula, A. V. et al. Stoichiometry of base excision repair proteins correlates with increased somatic CAG instability in striatum over cerebellum in Huntington’s disease transgenic mice. PLoS Genet. 5, e1000749 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  16. Lin, Y., Dent, S. Y., Wilson, J. H., Wells, R. D. & Napierala, M. R loops stimulate genetic instability of CTG·CAG repeats. Proc. Natl Acad. Sci. USA 107, 692–697 (2010).

    Article  PubMed  Google Scholar 

  17. Pearson, C. E. et al. Slipped-strand DNAs formed by long (CAG)·(CTG) repeats: slipped-out repeats and slip-out junctions. Nucleic Acids Res. 30, 4534–4547 (2002).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  18. Reddy, K. et al. Processing of double-R-loops in (CAG)·(CTG) and C9orf72 (GGGGCC)·(GGCCCC) repeats causes instability. Nucleic Acids Res. 42, 10473–10487 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  19. Schmidt, M. H. & Pearson, C. E. Disease-associated repeat instability and mismatch repair. DNA Repair 38, 117–126 (2016).

    Article  CAS  PubMed  Google Scholar 

  20. Tome, S. et al. Msh3 polymorphisms and protein levels affect CAG repeat instability in Huntington’s disease mice. PLoS Genet. 9, e1003280 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  21. Nakamori, M., Pearson, C. E. & Thornton, C. A. Bidirectional transcription stimulates expansion and contraction of expanded (CTG)·(CAG) repeats. Hum. Mol. Genet. 20, 580–588 (2011).

    Article  CAS  PubMed  Google Scholar 

  22. Lin, Y. & Wilson, J. H. Nucleotide excision repair, mismatch repair, and R-loops modulate convergent transcription-induced cell death and repeat instability. PLoS ONE 7, e46807 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  23. Panigrahi, G. B., Slean, M. M., Simard, J. P., Gileadi, O. & Pearson, C. E. Isolated short CTG/CAG DNA slip-outs are repaired efficiently by hMutSβ, but clustered slip-outs are poorly repaired. Proc. Natl Acad. Sci. USA 107, 12593–12598 (2010).

    Article  PubMed  PubMed Central  Google Scholar 

  24. Hagihara, M. & Nakatani, K. Inhibition of DNA replication by a d(CAG) repeat binding ligand. Nucleic Acids Symp. Ser. 50, 147–148 (2006).

    Article  CAS  Google Scholar 

  25. Hagihara, M., He, H. & Nakatani, K. Small molecule modulates hairpin structures in CAG trinucleotide repeats. ChemBioChem 12, 1686–1689 (2011).

    Article  CAS  PubMed  Google Scholar 

  26. Nakatani, K. et al. Small-molecule ligand induces nucleotide flipping in (CAG)n trinucleotide repeats. Nat. Chem. Biol. 1, 39–43 (2005).

    Article  CAS  PubMed  Google Scholar 

  27. Nielsen, P. E., Zhen, W. P., Henriksen, U. & Buchardt, O. Sequence-influenced interactions of oligoacridines with DNA detected by retarded gel electrophoretic migrations. Biochemistry 27, 67–73 (1988).

    Article  CAS  PubMed  Google Scholar 

  28. Pluciennik, A. et al. Extrahelical (CAG)/(CTG) triplet repeat elements support proliferating cell nuclear antigen loading and MutLα endonuclease activation. Proc. Natl Acad. Sci. USA 110, 12277–12282 (2013).

    Article  PubMed  PubMed Central  Google Scholar 

  29. Shelbourne, P. F. et al. Triplet repeat mutation length gains correlate with cell-type specific vulnerability in Huntington disease brain. Hum. Mol. Genet. 16, 1133–1142 (2007).

    Article  CAS  PubMed  Google Scholar 

  30. Silveira, I. et al. Trinucleotide repeats in 202 families with ataxia: a small expanded (CAG)n allele at the SCA17 locus. Arch. Neurol. 59, 623–629 (2002).

    Article  CAS  PubMed  Google Scholar 

  31. Sanchez-Contreras, M. & Cardozo-Pelaez, F. Age-related length variability of polymorphic CAG repeats. DNA Repair 49, 26–32 (2017).

    Article  CAS  PubMed  Google Scholar 

  32. Gao, R. et al. Instability of expanded CAG/CAA repeats in spinocerebellar ataxia type 17. Eur. J. Hum. Genet. 16, 215–222 (2008).

    Article  CAS  PubMed  Google Scholar 

  33. Gallon, R. et al. A sensitive and scalable microsatellite instability assay to diagnose constitutional mismatch repair deficiency by sequencing of peripheral blood leukocytes. Hum. Mutat. 40, 649–655 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  34. Keohavong, P., Xi, L. & Grant, S. G. Molecular analysis of mutations in the human HPRT gene. Methods Mol. Biol. 291, 161–170 (2005).

    CAS  PubMed  Google Scholar 

  35. Keohavong, P., Xi, L. & Grant, S. G. Molecular analysis of mutations in the human HPRT gene. Methods Mol. Biol. 1105, 291–301 (2014).

    Article  CAS  PubMed  Google Scholar 

  36. Albertini, R. J. et al. Mutagenicity monitoring following battlefield exposures: longitudinal study of HPRT mutations in Gulf War I veterans exposed to depleted uranium. Environ. Mol. Mutagen. 56, 581–593 (2015).

    Article  CAS  PubMed  Google Scholar 

  37. Nicklas, J. A. et al. Mutagenicity monitoring following battlefield exposures: molecular analysis of HPRT mutations in Gulf War I veterans exposed to depleted uranium. Environ. Mol. Mutagen. 56, 594–608 (2015).

    Article  CAS  PubMed  Google Scholar 

  38. Poon, S. L., McPherson, J. R., Tan, P., Teh, B. T. & Rozen, S. G. Mutation signatures of carcinogen exposure: genome-wide detection and new opportunities for cancer prevention. Genome Med. 6, 24 (2014).

    Article  PubMed  PubMed Central  Google Scholar 

  39. Behjati, S. et al. Mutational signatures of ionizing radiation in second malignancies. Nat. Commun. 7, 12605 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  40. Phillips, D. H. Mutational spectra and mutational signatures: insights into cancer aetiology and mechanisms of DNA damage and repair. DNA Repair 71, 6–11 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  41. Alexandrov, L. B. et al. The repertoire of mutational signatures in human cancer. Preprint at bioRxiv https://doi.org/10.1101/322859 (2019).

  42. Kucab, J. E. et al. A compendium of mutational signatures of environmental agents. Cell 177, 821–836.e16 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  43. Behjati, S. et al. Genome sequencing of normal cells reveals developmental lineages and mutational processes. Nature 513, 422–425 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  44. Rouhani, F. J. et al. Mutational history of a human cell lineage from somatic to induced pluripotent stem cells. PLoS Genet. 12, e1005932 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  45. Shlien, A. et al. Combined hereditary and somatic mutations of replication error repair genes result in rapid onset of ultra-hypermutated cancers. Nat. Genet. 47, 257–262 (2015).

    Article  CAS  PubMed  Google Scholar 

  46. Chalmers, Z. R. et al. Analysis of 100,000 human cancer genomes reveals the landscape of tumor mutational burden. Genome Med. 9, 34 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  47. Campbell, B. B. et al. Comprehensive analysis of hypermutation in human cancer. Cell 171, 1042–1056.e10 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  48. Hodel, K. P. et al. Explosive mutation accumulation triggered by heterozygous human Pol epsilon proofreading-deficiency is driven by suppression of mismatch repair. eLife 7, e32692 (2018).

    Article  PubMed  PubMed Central  Google Scholar 

  49. Rayner, E. et al. A panoply of errors: polymerase proofreading domain mutations in cancer. Nat. Rev. Cancer 16, 71–81 (2016).

    Article  CAS  PubMed  Google Scholar 

  50. Nakamori, M., Gourdon, G. & Thornton, C. A. Stabilization of expanded (CTG)·(CAG) repeats by antisense oligonucleotides. Mol. Ther. 19, 2222–2227 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  51. Su, X. A. & Freudenreich, C. H. Cytosine deamination and base excision repair cause R-loop-induced CAG repeat fragility and instability in Saccharomyces cerevisiae. Proc. Natl Acad. Sci. USA 114, E8392–E8401 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  52. Lin, Y., Hubert, L. Jr. & Wilson, J. H. Transcription destabilizes triplet repeats. Mol. Carcinog. 48, 350–361 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  53. Tomé, S. et al. MSH2 ATPase domain mutation affects CTG·CAG repeat instability in transgenic mice. PLoS Genet. 5, e1000482 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  54. McMurray, C. T. Hijacking of the mismatch repair system to cause CAG expansion and cell death in neurodegenerative disease. DNA Repair 7, 1121–1134 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  55. Morales, F. et al. A polymorphism in the MSH3 mismatch repair gene is associated with the levels of somatic instability of the expanded CTG repeat in the blood DNA of myotonic dystrophy type 1 patients. DNA Repair 40, 57–66 (2016).

    Article  CAS  PubMed  Google Scholar 

  56. Flower, M. et al. MSH3 modifies somatic instability and disease severity in Huntington’s and myotonic dystrophy type 1. Brain 142, 1876–1886 (2019).

    Article  PubMed Central  Google Scholar 

  57. Panigrahi, G. B., Lau, R., Montgomery, S. E., Leonard, M. R. & Pearson, C. E. Slipped (CTG)*(CAG) repeats can be correctly repaired, escape repair or undergo error-prone repair. Nat. Struct. Mol. Biol. 12, 654–662 (2005).

    Article  CAS  PubMed  Google Scholar 

  58. Zhang, T., Huang, J., Gu, L. & Li, G. M. In vitro repair of DNA hairpins containing various numbers of CAG/CTG trinucleotide repeats. DNA Repair 11, 201–209 (2012).

    Article  CAS  PubMed  Google Scholar 

  59. Lai, Y. et al. Crosstalk between MSH2–MSH3 and polβ promotes trinucleotide repeat expansion during base excision repair. Nat. Commun. 7, 12465 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  60. Tian, L. et al. Mismatch recognition protein MutSβ does not hijack (CAG)n hairpin repair in vitro. J. Biol. Chem. 284, 20452–20456 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  61. Nakatani, R., Nakamori, M., Fujimura, H., Mochizuki, H. & Takahashi, M. P. Large expansion of CTG·CAG repeats is exacerbated by MutSβ in human cells. Sci. Rep. 5, 11020 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  62. Chen, H., Lisby, M. & Symington, L. S. RPA coordinates DNA end resection and prevents formation of DNA hairpins. Mol. Cell 50, 589–600 (2013).

    Article  CAS  PubMed  Google Scholar 

  63. Nguyen, B. et al. Diffusion of human replication protein A along single-stranded DNA. J. Mol. Biol. 426, 3246–3261 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  64. Tsurimoto, T. & Stillman, B. Multiple replication factors augment DNA synthesis by the two eukaryotic DNA polymerases, alpha and delta. EMBO J. 8, 3883–3889 (1989).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  65. Tsurimoto, T. & Stillman, B. Replication factors required for SV40 DNA replication in vitro. I. DNA structure-specific recognition of a primer–template junction by eukaryotic DNA polymerases and their accessory proteins. J. Biol. Chem. 266, 1950–1960 (1991).

    CAS  PubMed  Google Scholar 

  66. Chan, N. L. et al. The Werner syndrome protein promotes CAG/CTG repeat stability by resolving large (CAG)n/(CTG)n hairpins. J. Biol. Chem. 287, 30151–30156 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  67. Callahan, J. L., Andrews, K. J., Zakian, V. A. & Freudenreich, C. H. Mutations in yeast replication proteins that increase CAG/CTG expansions also increase repeat fragility. Mol. Cell. Biol. 23, 7849–7860 (2003).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  68. Raji, N. S., Krishna, T. H. & Rao, K. S. DNA-polymerase α, β, Δ and ε activities in isolated neuronal and astroglial cell fractions from developing and aging rat cerebral cortex. Int. J. Dev. Neurosci. 20, 491–496 (2002).

    Article  CAS  PubMed  Google Scholar 

  69. Kovalenko, M. et al. Msh2 acts in medium-spiny striatal neurons as an enhancer of CAG instability and mutant huntingtin phenotypes in Huntington’s disease knock-in mice. PLoS ONE 7, e44273 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  70. Mangiarini, L. et al. Instability of highly expanded CAG repeats in mice transgenic for the Huntington’s disease mutation. Nat. Genet. 15, 197–200 (1997).

    Article  CAS  PubMed  Google Scholar 

  71. Chiang, C. et al. Complex reorganization and predominant non-homologous repair following chromosomal breakage in karyotypically balanced germline rearrangements and transgenic integration. Nat. Genet. 44, 390–397 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  72. Larson, E., Fyfe, I., Morton, A. J. & Monckton, D. G. Age-, tissue- and length-dependent bidirectional somatic CAG·CTG repeat instability in an allelic series of R6/2 Huntington disease mice. Neurobiol. Dis. 76, 98–111 (2015).

    Article  CAS  PubMed  Google Scholar 

  73. Kennedy, L. & Shelbourne, P. F. Dramatic mutation instability in HD mouse striatum: does polyglutamine load contribute to cell-specific vulnerability in Huntington’s disease? Hum. Mol. Genet. 9, 2539–2544 (2000).

    Article  CAS  PubMed  Google Scholar 

  74. Ishiguro, H. et al. Age-dependent and tissue-specific CAG repeat instability occurs in mouse knock-in for a mutant Huntington’s disease gene. J. Neurosci. Res. 65, 289–297 (2001).

    Article  CAS  PubMed  Google Scholar 

  75. Gonitel, R. et al. DNA instability in postmitotic neurons. Proc. Natl Acad. Sci. USA 105, 3467–3472 (2008).

    Article  PubMed  PubMed Central  Google Scholar 

  76. De Rooij, K. E., De Koning Gans, P. A., Roos, R. A., Van Ommen, G. J. & Den Dunnen, J. T. Somatic expansion of the (CAG)n repeat in Huntington disease brains. Hum. Genet. 95, 270–274 (1995).

    Article  PubMed  Google Scholar 

  77. Lee, J. M., Pinto, R. M., Gillis, T., St Claire, J. C. & Wheeler, V. C. Quantification of age-dependent somatic CAG repeat instability in Hdh CAG knock-in mice reveals different expansion dynamics in striatum and liver. PLoS ONE 6, e23647 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  78. Lee, J. M. et al. A novel approach to investigate tissue-specific trinucleotide repeat instability. BMC Syst. Biol. 4, 29 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  79. Wheeler, V. C. et al. Factors associated with HD CAG repeat instability in Huntington disease. J. Med. Genet. 44, 695–701 (2007).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  80. Higham, C. F., Morales, F., Cobbold, C. A., Haydon, D. T. & Monckton, D. G. High levels of somatic DNA diversity at the myotonic dystrophy type 1 locus are driven by ultra-frequent expansion and contraction mutations. Hum. Mol. Genet. 21, 2450–2463 (2012).

    Article  CAS  PubMed  Google Scholar 

  81. Veitch, N. J. et al. Inherited CAG.CTG allele length is a major modifier of somatic mutation length variability in Huntington disease. DNA Repair 6, 789–796 (2007).

    Article  CAS  PubMed  Google Scholar 

  82. Hornsby, P. J. & Didenko, V. V. In situ ligation: a decade and a half of experience. Methods Mol. Biol. 682, 49–63 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  83. Majtnerova, P. & Rousar, T. An overview of apoptosis assays detecting DNA fragmentation. Mol. Biol. Rep. 45, 1469–1478 (2018).

    Article  CAS  PubMed  Google Scholar 

  84. Iannicola, C. et al. Early alterations in gene expression and cell morphology in a mouse model of Huntington’s disease. J. Neurochem. 75, 830–839 (2000).

    Article  CAS  PubMed  Google Scholar 

  85. Turmaine, M. et al. Nonapoptotic neurodegeneration in a transgenic mouse model of Huntington’s disease. Proc. Natl Acad. Sci. USA 97, 8093–8097 (2000).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  86. Yu, Z. X. et al. Mutant huntingtin causes context-dependent neurodegeneration in mice with Huntington’s disease. J. Neurosci. 23, 2193–2202 (2003).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  87. DiFiglia, M. et al. Aggregation of huntingtin in neuronal intranuclear inclusions and dystrophic neurites in brain. Science 277, 1990–1993 (1997).

    Article  CAS  PubMed  Google Scholar 

  88. Li, S. H. & Li, X. J. Aggregation of N-terminal huntingtin is dependent on the length of its glutamine repeats. Hum. Mol. Genet. 7, 777–782 (1998).

    Article  CAS  PubMed  Google Scholar 

  89. Becher, M. W. et al. Intranuclear neuronal inclusions in Huntington’s disease and dentatorubral and pallidoluysian atrophy: correlation between the density of inclusions and IT15 CAG triplet repeat length. Neurobiol. Dis. 4, 387–397 (1998).

    Article  CAS  PubMed  Google Scholar 

  90. Li, H. et al. Ultrastructural localization and progressive formation of neuropil aggregates in Huntington’s disease transgenic mice. Hum. Mol. Genet. 8, 1227–1236 (1999).

    Article  CAS  PubMed  Google Scholar 

  91. Li, H., Li, S. H., Johnston, H., Shelbourne, P. F. & Li, X. J. Amino-terminal fragments of mutant huntingtin show selective accumulation in striatal neurons and synaptic toxicity. Nat. Genet. 25, 385–389 (2000).

    Article  CAS  PubMed  Google Scholar 

  92. Carty, N. et al. Characterization of HTT inclusion size, location, and timing in the zQ175 mouse model of Huntington’s disease: an in vivo high-content imaging study. PLoS ONE 10, e0123527 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  93. Kaytor, M. D., Wilkinson, K. D. & Warren, S. T. Modulating huntingtin half-life alters polyglutamine-dependent aggregate formation and cell toxicity. J. Neurochem. 89, 962–973 (2004).

    Article  CAS  PubMed  Google Scholar 

  94. Coufal, M. et al. Discovery of a novel small-molecule targeting selective clearance of mutant huntingtin fragments. J. Biomol. Screen. 12, 351–360 (2007).

    Article  CAS  PubMed  Google Scholar 

  95. Chopra, V. et al. A small-molecule therapeutic lead for Huntington’s disease: preclinical pharmacology and efficacy of C2-8 in the R6/2 transgenic mouse. Proc. Natl Acad. Sci. USA 104, 16685–16689 (2007).

    Article  PubMed  PubMed Central  Google Scholar 

  96. Butler, D. C. & Messer, A. Bifunctional anti-huntingtin proteasome-directed intrabodies mediate efficient degradation of mutant huntingtin exon 1 protein fragments. PLoS ONE 6, e29199 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  97. Perucho, J. et al. Striatal infusion of glial conditioned medium diminishes huntingtin pathology in r6/1 mice. PLoS ONE 8, e73120 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  98. Tsvetkov, A. S. et al. Proteostasis of polyglutamine varies among neurons and predicts neurodegeneration. Nat. Chem. Biol. 9, 586–592 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  99. Penney, J. B. Jr, Vonsattel, J. P., MacDonald, M. E., Gusella, J. F. & Myers, R. H. CAG repeat number governs the development rate of pathology in Huntington’s disease. Ann. Neurol. 41, 689–692 (1997).

    Article  PubMed  Google Scholar 

  100. Wheeler, V. C. et al. Long glutamine tracts cause nuclear localization of a novel form of huntingtin in medium spiny striatal neurons in HdhQ92 and HdhQ111 knock-in mice. Hum. Mol. Genet. 9, 503–513 (2000).

    Article  CAS  PubMed  Google Scholar 

  101. Rosenblatt, A. et al. Does CAG repeat number predict the rate of pathological changes in Huntington’s disease? Ann. Neurol. 44, 708–709 (1998).

    Article  CAS  PubMed  Google Scholar 

  102. Wild, E. J. & Tabrizi, S. J. Therapies targeting DNA and RNA in Huntington’s disease. Lancet Neurol. 16, 837–847 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  103. Dabrowska, M., Juzwa, W., Krzyzosiak, W. J. & Olejniczak, M. Precise excision of the CAG tract from the Huntingtin gene by Cas9 nickases. Front. Neurosci. https://doi.org/10.3389/fnins.2018.00075 (2018).

  104. Shin, J. W. et al. Permanent inactivation of Huntington’s disease mutation by personalized allele-specific CRISPR/Cas9. Hum. Mol. Genet. 25, 4566–4576 (2016).

    CAS  PubMed  PubMed Central  Google Scholar 

  105. Monteys, A. M., Ebanks, S. A., Keiser, M. S. & Davidson, B. L. CRISPR/Cas9 editing of the mutant huntingtin allele in vitro and in vivo. Mol. Ther. 25, 12–23 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  106. Cinesi, C., Aeschbach, L., Yang, B. & Dion, V. Contracting CAG/CTG repeats using the CRISPR–Cas9 nickase. Nat. Commun. 7, 13272 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  107. Suelves, N., Kirkham-McCarthy, L., Lahue, R. S. & Gines, S. A selective inhibitor of histone deacetylase 3 prevents cognitive deficits and suppresses striatal CAG repeat expansions in Huntington’s disease mice. Sci. Rep. 7, 6082 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  108. Eisenstein, M. CRISPR takes on Huntington’s disease. Nature 557, S42–S43 (2018).

    Article  CAS  PubMed  Google Scholar 

  109. Martins, S. et al. Modifiers of (CAG)n instability in Machado–Joseph disease (MJD/SCA3) transmissions: an association study with DNA replication, repair and recombination genes. Hum. Genet. 133, 1311–1318 (2014).

    Article  CAS  PubMed  Google Scholar 

  110. Guo, J., Gu, L., Leffak, M. & Li, G. M. MutSβ promotes trinucleotide repeat expansion by recruiting DNA polymerase β to nascent (CAG)n or (CTG)n hairpins for error-prone DNA synthesis. Cell Res. 26, 775–786 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  111. Hanawalt, P. C. & Spivak, G. Transcription-coupled DNA repair: two decades of progress and surprises. Nat. Rev. Mol. Cell Biol. 9, 958–970 (2008).

    Article  CAS  PubMed  Google Scholar 

  112. Hou, C., Chan, N. L., Gu, L. & Li, G. M. Incision-dependent and error-free repair of (CAG)n/(CTG)n hairpins in human cell extracts. Nat. Struct. Mol. Biol. 16, 869–875 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  113. Chan, N. L. et al. Coordinated processing of 3′ slipped (CAG)n/(CTG)n hairpins by DNA polymerases β and Δ preferentially induces repeat expansions. J. Biol. Chem. 288, 15015–15022 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  114. Pinto, R. M. et al. Mismatch repair genes Mlh1 and Mlh3 modify CAG instability in Huntington’s disease mice: genome-wide and candidate approaches. PLoS Genet. 9, e1003930 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  115. Wheeler, V. C. et al. Mismatch repair gene Msh2 modifies the timing of early disease in Hdh Q111 striatum. Hum. Mol. Genet. 12, 273–281 (2003).

    Article  CAS  PubMed  Google Scholar 

  116. Strobel, S. A., Doucette-Stamm, L. A., Riba, L., Housman, D. E. & Dervan, P. B. Site-specific cleavage of human chromosome 4 mediated by triple-helix formation. Science 254, 1639–1642 (1991).

    Article  CAS  PubMed  Google Scholar 

  117. Mittelman, D. et al. Zinc-finger directed double-strand breaks within CAG repeat tracts promote repeat instability in human cells. Proc. Natl Acad. Sci. USA 106, 9607–9612 (2009).

    Article  PubMed  PubMed Central  Google Scholar 

  118. Zeitler, B. et al. Allele-selective transcriptional repression of mutant HTT for the treatment of Huntington’s disease. Nat. Med. 25, 1131–1142 (2019).

    Article  CAS  PubMed  Google Scholar 

  119. Mosbach, V., Poggi, L. & Richard, G. F. Trinucleotide repeat instability during double-strand break repair: from mechanisms to gene therapy. Curr. Genet. 65, 17–28 (2019).

    Article  CAS  PubMed  Google Scholar 

  120. Malankhanova, T. B., Malakhova, A. A., Medvedev, S. P. & Zakian, S. M. Modern genome editing technologies in Huntington’s disease research. J. Huntington’s Dis. 6, 19–31 (2017).

    Article  Google Scholar 

  121. Babacic, H., Mehta, A., Merkel, O. & Schoser, B. CRISPR-cas gene-editing as plausible treatment of neuromuscular and nucleotide-repeat-expansion diseases: a systematic review. PLoS ONE 14, e0212198 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  122. Gomes-Pereira, M. & Monckton, D. G. Chemical modifiers of unstable expanded simple sequence repeats: what goes up, could come down. Mutat. Res. 598, 15–34 (2006).

    Article  CAS  PubMed  Google Scholar 

  123. Pineiro, E. et al. Mutagenic stress modulates the dynamics of CTG repeat instability associated with myotonic dystrophy type 1. Nucleic Acids Res. 31, 6733–6740 (2003).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  124. Budworth, H. et al. Suppression of somatic expansion delays the onset of pathophysiology in a mouse model of Huntington’s disease. PLoS Genet. 11, e1005267 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  125. Gottesfeld, J. M., Neely, L., Trauger, J. W., Baird, E. E. & Dervan, P. B. Regulation of gene expression by small molecules. Nature 387, 202–205 (1997).

    Article  CAS  PubMed  Google Scholar 

  126. Leontieva, O. V. & Blagosklonny, M. V. CDK4/6-inhibiting drug substitutes for p21 and p16 in senescence: duration of cell cycle arrest and MTOR activity determine geroconversion. Cell Cycle 12, 3063–3069 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  127. Nakamori, M., Sobczak, K., Moxley, R. T. & Thornton, C. A. Scaled-down genetic analysis of myotonic dystrophy type 1 and type 2. Neuromuscul. Disord. 19, 759–762 (2009).

    Article  PubMed  PubMed Central  Google Scholar 

  128. Brook, J. D. et al. Molecular basis of myotonic dystrophy: expansion of a trinucleotide (CTG) repeat at the 3′ end of a transcript encoding a protein kinase family member. Cell 69, 385 (1992).

    CAS  PubMed  Google Scholar 

  129. Dietmaier, W. et al. Diagnostic microsatellite instability: definition and correlation with mismatch repair protein expression. Cancer Res. 57, 4749–4756 (1997).

    CAS  PubMed  Google Scholar 

  130. Kabbarah, O. et al. A panel of repeat markers for detection of microsatellite instability in murine tumors. Mol. Carcinog. 38, 155–159 (2003).

    Article  CAS  PubMed  Google Scholar 

  131. Koob, M. D. et al. An untranslated CTG expansion causes a novel form of spinocerebellar ataxia (SCA8). Nat. Genet. 21, 379–384 (1999).

    Article  CAS  PubMed  Google Scholar 

  132. Kremer, B. et al. Sex-dependent mechanisms for expansions and contractions of the CAG repeat on affected Huntington disease chromosomes. Am. J. Hum. Genet. 57, 343–350 (1995).

    CAS  PubMed  PubMed Central  Google Scholar 

  133. Cleary, J. D., Nichol, K., Wang, Y. H. & Pearson, C. E. Evidence of cis-acting factors in replication-mediated trinucleotide repeat instability in primate cells. Nat. Genet. 31, 37–46 (2002).

    Article  CAS  PubMed  Google Scholar 

  134. Panigrahi, G. B., Cleary, J. D. & Pearson, C. E. In vitro (CTG)*(CAG) expansions and deletions by human cell extracts. J. Biol. Chem. 277, 13926–13934 (2002).

    Article  CAS  PubMed  Google Scholar 

  135. Reddy, K. et al. Determinants of R-loop formation at convergent bidirectionally transcribed trinucleotide repeats. Nucleic Acids Res. 39, 1749–1762 (2011).

    Article  CAS  PubMed  Google Scholar 

  136. Binz, S. K., Dickson, A. M., Haring, S. J. & Wold, M. S. Functional assays for replication protein A (RPA). Methods Enzymol. 409, 11–38 (2006).

    Article  CAS  PubMed  Google Scholar 

  137. Zhou, Y., Meng, X., Zhang, S., Lee, E. Y. & Lee, M. Y. Characterization of human DNA polymerase delta and its subassemblies reconstituted by expression in the MultiBac system. PLoS ONE 7, e39156 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  138. Mason, A. C., Roy, R., Simmons, D. T. & Wold, M. S. Functions of alternative replication protein A in initiation and elongation. Biochemistry 49, 5919–5928 (2010).

    Article  CAS  PubMed  Google Scholar 

  139. Tome, S. et al. Tissue-specific mismatch repair protein expression: MSH3 is higher than MSH6 in multiple mouse tissues. DNA Repair 12, 46–52 (2013).

    Article  CAS  PubMed  Google Scholar 

  140. Jeon, I. et al. Human-to-mouse prion-like propagation of mutant huntingtin protein. Acta Neuropathol. 132, 577–592 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  141. Cibulskis, K. et al. Sensitive detection of somatic point mutations in impure and heterogeneous cancer samples. Nat. Biotechnol. 31, 213–219 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  142. Morgulis, A., Gertz, E. M., Schaffer, A. A. & Agarwala, R. A fast and symmetric DUST implementation to mask low-complexity DNA sequences. J. Comput. Biol. 13, 1028–1040 (2006).

    Article  CAS  PubMed  Google Scholar 

  143. Li, H. Toward better understanding of artifacts in variant calling from high-coverage samples. Bioinformatics 30, 2843–2851 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  144. Alexandrov, L. B. et al. Signatures of mutational processes in human cancer. Nature 500, 415–421 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  145. Alexandrov, L. B., Nik-Zainal, S., Wedge, D. C., Campbell, P. J. & Stratton, M. R. Deciphering signatures of mutational processes operative in human cancer. Cell Rep. 3, 246–259 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  146. Nik-Zainal, S. et al. Landscape of somatic mutations in 560 breast cancer whole-genome sequences. Nature 534, 47–54 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  147. Li, H. & Durbin, R. Fast and accurate long-read alignment with Burrows–Wheeler transform. Bioinformatics 26, 589–595 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

Download references

Acknowledgements

This work was partially supported by the Canadian Institutes of Health Research (FRN388879, J.-Y.M. and FRN148910, C.E.P.), Natural Sciences and Engineering Research Council (RGPIN-2016-08355, C.E.P.), Muscular Dystrophy Canada (C.E.P.), Tribute Communities (C.E.P.), The Petroff Family Fund (C.E.P.), The Kazman Family Fund (C.E.P.), The Marigold Foundation (C.E.P.), The National Center of Neurology and Psychiatry (29-4, M.N.), a JSPS KAKENHI Grant-in-Aid for Young Scientists (Start-up A, 24890110 and 25713034, M.N.), Scientific Research (B, 16H05321, M.N.) and Specially Promoted Research (26000007, K.N.), the Cancer Research UK Catalyst Award (C569/A24991, R.G.), the US National Institutes of Health (2 R01 ES014737, M.Y.W.T.L.) and a US National Institutes of Health (NIH) grant (HG010169 to E.E.E.). E.E.E. is an investigator of the Howard Hughes Medical Institute, A. Shlien holds the Canada Research Chair in Childhood Cancer Genomics, J.-Y.M. holds the Fonds de Recherche de Santé Québec Chair in Genome Stability, and C.E.P. holds the Canada Research Chair in Disease-Associated Genome Instability. We acknowledge the technical support of R. Manabe, K. Hayashi, P. Wang, L. Yu, M. Mirceta, N. Thakkar, I. Panigrahi, D. Ripsman, H. Adhikary, S. Bérubé, Y. Coulombe, A. Couturier, M. Scofield Sorensen and K. Hoekzema. We acknowledge TCAG (SickKids) for expedited sequencing.

Author information

Authors and Affiliations

Authors

Contributions

M.N. performed repeat length analysis, and cell and mouse treatments. G.B.P. performed replication, repair, footprinting and R-loop processing. S.L. performed NA binding, repeat instability analysis, protein interactions, polδ extension assay and instability index assessment. H.H., H.T., M.P.T. and H.M. performed mouse treatments. T.O., J.Li, A. Sakata, A.M. and K.N. synthesized, purified and characterized NA. J.Luo and T.P. performed cell treatments. T.G.-D. performed assessment of NA on mHTT aggregates, HTT translation and TUNEL assay. M.-C.C., N.J., J.-Y.M., M.S.W., X.W. and M.Y.W.T.L. performed protein purification. S.L., G.B.P. and K.C. performed in vitro repair/binding assays. J.H., K.M.M. and E.E.E. performed single-molecule sequencing and bioinformatic analysis. R.G. and M.S.-K. performed MSI assay. S.D., M.L., L.-M.E. and A. Shlien performed whole-genome sequencing and mutation signature analysis. C.E.P., M.N., K.N., G.B.P. and S.L. conceived experiments, analyzed data and wrote the manuscript. M.N., G.B.P. and S.L. contributed equally to the study. All authors discussed the results and commented on the manuscript.

Corresponding author

Correspondence to Christopher E. Pearson.

Ethics declarations

Competing interests

The authors declare no competing interests.

Additional information

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

Extended data

Extended Data Fig. 1 NA does not affect replication efficiency or replication fork progression.

Three circular plasmids containing the SV40 origin of replication, and an expanded (CAG)79•(CTG)79 repeat tract (pDM79EF and pDM79HF) or no repeats (pKN16), were replicated in vitro by human (HeLa) cell extracts without or with NA (7.5 µM or 15 µM) treatment. The location of SV40-ori determines the replication direction and which strand will be used as the leading or the lagging strand template. pDM79HF uses the CAG strand as the lagging strand template, while pDM79EF uses the CTG strand as the lagging strand template (schematic on the top of the gel panel). Replication products were purified and linearized with BamHI. An equal portion of the reaction material was also digested with BamHI and DpnI as DpnI digests un-replicated and partially-replicated material, as shown in the schematic (top figure). The digestion products were electrophoresed on a 1% agarose gel to resolve completely replicated and un-replicated material (bottom figure). Equal amount of unreplicated plasmid DNA was digested with DpnI and stained with Ethidium Bromide to show the complete digestion of unreplicated plasmid DNA (Bottom panel). Panel I, ethidium bromide stained, Panel II, autorad: marker (lane 1); DpnI undigested plasmid DNA (lane 2); DpnI digested unreplicated plasmid DNA (lane 3-4); replicated plasmid DNA, DpnI resistant (lane 5). No difference in DpnI resistant material is observed between replication in the presence or absence of NA, in all the three templates tested (panel III, IV, V). Blots have been cropped and the corresponding full blots are available in the Source Data files.

Source data

Extended Data Fig. 2 NA does not affect non-mutant genetically stable repeats.

a, b, Representative data showing small-pool PCR (spPCR) for the non-expanded CAG/CTG repeat length of CASK and Mdf15 in HD primary fibroblast cells (a) and spPCR for the non-expanded CAG/CTG repeat length of CASK, Mdf15, ATXN8 and the non-expanded HTT allele genes in HT1080-(CAG)850 cells (b). Even under the sensitive mutation detection capacity of spPCR, length variation was not observed in either NA treated- and untreated-cells. Notably, some reactions did not show any product as is typical of the low genomic DNA template dilutions used in spPCR. c, The repeat-tract lengths of the CASK, ATXN8, and Mfd15 loci in HT1080-(CAG)850 cells (initial clone and cells after 30 days incubation with or without NA). Length variation was not observed at any of these repeats of normal length loci in HT1080-(CAG)850 cells (after 30 days incubation with or without NA). Three independent experiments were performed. d, spPCR for the non-expanded CAG tracts in TBP alleles in HD patient fibroblasts treated with or without NA for 40 days. e, spPCR for the non-expanded CAG tracts in TBP alleles in HD R6/2 mouse striatum with four injections of NA or saline. f, Microsatellite instability assay. Assay scores >1.3 indicate increased MSI relative to a control sample set from peripheral blood leukocytes. Both NA positive and NA negative HD cells with (CAG)43 or (CAG)180 scored <1.3, indicating no effect of treatment on MSI. Eight known CMMRD-negative controls and 3 known CMMRD-positive controls were included in the assay.

Extended Data Fig. 3 NA is not a general mutagen.

Towards assessing whether NA-treatment acted as a general mutagen to sequences other than CAG slip-outs, we harnessed the high read accuracy and depth of single molecule, real time, circular consensus sequencing (SMRT-CCS). Single-molecule sequencing was done on the HPRT1 gene – widely used as a surrogate indicator of the global effect of induced genetic variation. For each replicate, we calculated the relative mutation rate between NA- and saline-treated cells as the mutation rate for NA-treated cells minus the rate for saline-treated cells and identified excess mutation rates based on an absolute relative rate >0.5%. a, Schematic of HPRT1 sequencing for mutation detection. Briefly, cells were grown under identical conditions differing only by the addition of NA (50 μM) or saline, DNAs were isolated, HPRT1 exons 2 and 3 PCR amplified and sequenced. b, Quality control for our analysis. c,d, Comparison of sequence variations between NA-treated and saline treated is presented. We chose to compare the single-molecule sequence reads of individual X chromosome-linked HPRT1 alleles (exons 2 and 3) from our male HD patient-derived cells (c), and our male R6/2 mice (d), that had been NA- or saline-treated. Each read represents a single cell (Supplementary Note). Graphs show the distribution of sequence variants by relative mutation rate between three experimental replicates of NA-treated and saline-treated cells sequenced with PacBio single-molecule long reads.

Extended Data Fig. 4 NA does not affect HTT transcription or translation.

a, NA does not affect transcription across expanded repeats in HTT in HD patient cells, determined by quantitative real-time reverse transcriptase (qRT)-PCR and normalized to U6 RNA. Data are indicated as the mean ± s.d. of independent triplicates. b, Western blot showing that NA does not affect HTT translation in HD patient cells with (CAG)43. Western blots were repeated 4 times with similar results. Blots have been cropped and the corresponding full blots are available in the Source Data files. c, Extraction of NA from DNA by solvents.

Source data

Extended Data Fig. 5 NA induces contractions during R-loops processing.

a, Schematic of R-loop formation, processing, and analysis. Pre-formed double-R-loops were processed by terminally differentiated (retinoic acid) human neuron-like cell extracts (SH-SY5Y) in the absence or presence of NA (50 μM), as described and DNA repeat lengths scored as expansions, contractions, or stable, by the STRIP assay (Methods). b, Representative example of STRIP analysis. Transcription products were isolated, processed and transformed in E. coli cells, previously shown to stably maintain the (CAG)79•(CTG)79 lengths (Methods). Plasmids isolated from individual bacterial colonies were digested with restriction enzymes to release the repeat containing fragment, resolved on 4% polyacrylamide gels and scored for instability. c, Graphical analysis of STRIP results. Two-sided χ2 test was performed to compare 191 untreated colonies vs. 100 NA-treated colonies.

Extended Data Fig. 6 Dosing regimen.

A single drug administration involved six separate stereotactic injections (three injections of drug in saline or saline into three different striatal regions of either the left or right striatum, respectively). At the onset mice were 6-weeks old.

Extended Data Fig. 7 Instability Index calculation.

Instability index determination was as described77,78, using a relative peak height threshold, with modifications. To quantify the levels of instability from GeneMapper traces peak height was used to determine a relative threshold of 20% based upon the main peak in the shorter mode of the control striatum (see points 1 & 2 in the figure). We used a conservative threshold factor (20%) as this detects peaks with good signal intensity and is more resistant to amplification variation than lower thresholds. Lower thresholds (10%, 5%) can provide more sensitive quantification. Peaks falling below his threshold were excluded from analysis. Peak heights were scored (see point 3) and normalized to the total of all peak heights in a given scan (see point 4). Since we are comparing the effect of NA versus saline upon instability in the striatum, the CAG length distribution in tail is not a factor in this comparison, but is for determining absolute instability, as in previous studies62,66. So as to facilitate comparison between NA and saline-treated striatum, these were normalized by multiplying the values by the change in CAG length of each peak relative to the highest peak in saline-treated striatum (see point 5), as opposed to the highest peak in the tail, as previously done77,78. These normalized values (see point 6) were summed to generate the instability index (see point 7). Striatum analysis for mouse vi is shown as an example R6/2, 6-weeks treated with four injections spanning 4 weeks of saline (red) or NA (blue). Peaks of the main allele in the saline-treated striatum, NA-treated striatum and tail of the same mouse, are indicated by triangle-brackets at the top (see point 1).

Extended Data Fig. 8 A total of ten HD mice revealed consistent NA-induced contractions of expanded CAG repeats.

Instability Indices in striatum of ten mice (iv-xiii) treated four times with saline in the right striatum and NA in the left striatum. Indices in NA-treated striatum were significantly different from the control saline-treated striatum (Mann-Whitney, P= 0.00035). Instability Indices for mouse v and xi are positive for both NA and saline as there are less data points to the left of the highest peak compared to the points to the right. Still, after NA treatment there is a reduction in the index.

Extended Data Fig. 9 NA does not induce cell death in the CNS and cell proliferation, and does not affect transcription across the Htt locus.

a, Histological study, mouse striatum with saline, NA in saline, or no injection, followed by H&E staining. Three independent experiments were performed. b, NeuN staining showing that NA does not induce cell death. Quantification of NeuN positive cells below. Data are indicated as mean ± s.d. of triplicates. c, Doublecortin staining showing that NA does not induce cell proliferation. Three independent experiments were performed. d, The effect of NA on TUNEL signal as assessed via fluorescent microscopy and immunohistochemistry. Representative 40x magnification confocal images of striatal medium spiny neurons (MSNs) of R6/2 mice treated with saline (right striata) and 50 μM NA (left striata) stained for TUNEL (red, staining apoptotic cells), and DARPP32 (green, staining MSNs). Panel locations (i-vi) correspond to the locations outlined in Fig. 7 (middle panels). e, NA does not affect transcription across expanded repeats in HTT in HD patient cells and mouse striatum, determined by quantitative real-time reverse transcriptase (qRT)-PCR and normalized to U6 RNA, expressed as the ratio of NA-treated vs. PBS-treated R6/2 striatum. Data are indicated as mean ± s.d. of independent triplicates.

Supplementary information

Supplementary Information

Supplementary Note, Figs. 1–6 and Table 2

Reporting Summary

Supplementary Table 1

Summary of the instability data from analysis of HD patient fibroblasts GM09197 with (CAG)180 and GM02191 with (CAG)43, and the cell model HT1080, and complete spPCR datasets for each repetition of each experiment and the associated graphs.

Source data

Source Data Fig. 1

Statistical source data;

Source Data Fig. 1

unprocessed gel.

Source Data Fig. 2

Statistical source data;

Source Data Fig. 2

unprocessed gel.

Source Data Fig. 3

Statistical source data; unprocessed gel.

Source Data Fig. 4

Statistical source data;

Source Data Fig. 4

unprocessed gel.

Source Data Fig. 5

Statistical source data;

Source Data Fig. 5

unprocessed gel.

Source Data Fig. 6

Statistical source data;

Source Data Fig. 6

unprocessed gel.

Source Data Fig. 7

Statistical source data.

Source Data Fig. 8

Statistical source data.

Source Data Extended Data Fig. 1

Unprocessed gel.

Source Data Extended Data Fig. 4

Unprocessed gel.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Nakamori, M., Panigrahi, G.B., Lanni, S. et al. A slipped-CAG DNA-binding small molecule induces trinucleotide-repeat contractions in vivo. Nat Genet 52, 146–159 (2020). https://doi.org/10.1038/s41588-019-0575-8

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41588-019-0575-8

This article is cited by

Search

Quick links

Nature Briefing: Translational Research

Sign up for the Nature Briefing: Translational Research newsletter — top stories in biotechnology, drug discovery and pharma.

Get what matters in translational research, free to your inbox weekly. Sign up for Nature Briefing: Translational Research