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:

Sonic hedgehog accelerates DNA replication to cause replication stress promoting cancer initiation in medulloblastoma

An Author Correction to this article was published on 19 April 2021

This article has been updated

Abstract

The mechanisms generating cancer-initiating mutations are not well understood. Sonic hedgehog (SHH) pathway activation is frequent in medulloblastoma (MB), with PTCH1 mutations being a common initiating event. Here we investigated the role of the developmental mitogen SHH in initiating carcinogenesis in the cells of origin: granule cell progenitors (GCPs). We delineate a molecular mechanism for tumor initiation in MB. Exposure of GCPs to Shh causes a distinct form of DNA replication stress, increasing both origin firing and fork velocity. Shh promotes DNA helicase loading and activation, with increased Cdc7-dependent origin firing. The S-phase duration is reduced and hyper-recombination occurs, causing copy number neutral loss of heterozygosity—a frequent event at the PTCH1/ptch1 locus. Moreover, Cdc7 inhibition to attenuate origin firing reduces recombination and preneoplastic tumor formation in mice. Therefore, tissue-specific replication stress induced by Shh promotes loss of heterozygosity, which in tumor-prone Ptch1+/− GCPs results in loss of this tumor suppressor—an early cancer-initiating event.

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: Copy number variants and CN-LOH are the source of PTCH1 wild-type allele inactivation.
Fig. 2: Shh causes DNA replication stress.
Fig. 3: Shh alters DNA replication dynamics.
Fig. 4: Shh promotes pre-replication complex assembly.
Fig. 5: Shh induces helicase activation and origin firing.
Fig. 6: Shh-dependent replication initiation domains revealed by EdU-seq.
Fig. 7: Origin firing is required for Shh-dependent replication stress and recombination.
Fig. 8: In vivo Cdc7 inhibition reduces origin firing and hyper-recombination, preventing MB initiation.

Similar content being viewed by others

Data availability

Deep-sequencing data supporting the findings of this study have been deposited in the Gene Expression Omnibus under accession codes GSE147409 (EdU-seq) and GSE147410 (RNA-seq). Human LOH data in SHH-MB were derived from the dataset EGAD00001003127 obtained with authorization from the International Cancer Genome Consortium (https://icgc.org/). Mouse CNV data were downloaded from GSE19381. All other data supporting the findings of this study are available from the corresponding author on reasonable request. Source data are provided with this paper.

Change history

References

  1. Azevedo, F. A. et al. Equal numbers of neuronal and nonneuronal cells make the human brain an isometrically scaled-up primate brain. J. Comp. Neurol. 513, 532–541 (2009).

    PubMed  Google Scholar 

  2. Wechsler-Reya, R. J. & Scott, M. P. Control of neuronal precursor proliferation in the cerebellum by Sonic hedgehog. Neuron 22, 103–114 (1999).

    CAS  PubMed  Google Scholar 

  3. Oliver, T. G. et al. Loss of patched and disruption of granule cell development in a pre-neoplastic stage of medulloblastoma. Development 132, 2425–2439 (2005).

    CAS  PubMed  Google Scholar 

  4. Yang, Z. J. et al. Medulloblastoma can be initiated by deletion of patched in lineage-restricted progenitors or stem cells. Cancer Cell 14, 135–145 (2008).

    CAS  PubMed  PubMed Central  Google Scholar 

  5. Schuller, U. et al. Acquisition of granule neuron precursor identity is a critical determinant of progenitor cell competence to form Shh-induced medulloblastoma. Cancer Cell 14, 123–134 (2008).

    CAS  PubMed  PubMed Central  Google Scholar 

  6. Kool, M. et al. Molecular subgroups of medulloblastoma: an international meta-analysis of transcriptome, genetic aberrations, and clinical data of WNT, SHH, group 3, and group 4 medulloblastomas. Acta Neuropathol. 123, 473–484 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  7. Northcott, P. A. et al. Medulloblastomics: the end of the beginning. Nat. Rev. Cancer 12, 818–834 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  8. Northcott, P. A. et al. The whole-genome landscape of medulloblastoma subtypes. Nature 547, 311–317 (2017).

    CAS  PubMed  PubMed Central  Google Scholar 

  9. Tamayo-Orrego, L. et al. Evasion of cell senescence leads to medulloblastoma progression. Cell Rep. 14, 2925–2937 (2016).

    CAS  PubMed  Google Scholar 

  10. Bartkova, J. et al. DNA damage response as a candidate anti-cancer barrier in early human tumorigenesis. Nature 434, 864–870 (2005).

    CAS  PubMed  Google Scholar 

  11. Gorgoulis, V. G. et al. Activation of the DNA damage checkpoint and genomic instability in human precancerous lesions. Nature 434, 907–913 (2005).

    CAS  PubMed  Google Scholar 

  12. Negrini, S., Gorgoulis, V. G. & Halazonetis, T. D. Genomic instability—an evolving hallmark of cancer. Nat. Rev. Mol. Cell Biol. 11, 220–228 (2010).

    CAS  PubMed  Google Scholar 

  13. Tomasetti, C., Li, L. & Vogelstein, B. Stem cell divisions, somatic mutations, cancer etiology, and cancer prevention. Science 355, 1330–1334 (2017).

    CAS  PubMed  PubMed Central  Google Scholar 

  14. Macheret, M. & Halazonetis, T. D. DNA replication stress as a hallmark of cancer. Annu. Rev. Pathol. 10, 425–448 (2015).

    CAS  PubMed  Google Scholar 

  15. Hills, S. A. & Diffley, J. F. DNA replication and oncogene-induced replicative stress. Curr. Biol. 24, R435–R444 (2014).

    CAS  PubMed  Google Scholar 

  16. Lasko, D., Cavenee, W. & Nordenskjold, M. Loss of constitutional heterozygosity in human cancer. Annu. Rev. Genet. 25, 281–314 (1991).

    CAS  PubMed  Google Scholar 

  17. Pazzaglia, S. et al. Two-hit model for progression of medulloblastoma preneoplasia in Patched heterozygous mice. Oncogene 25, 5575–5580 (2006).

    CAS  PubMed  Google Scholar 

  18. Ishida, Y. et al. Genomic and gene expression signatures of radiation in medulloblastomas after low-dose irradiation in Ptch1 heterozygous mice. Carcinogenesis 31, 1694–1701 (2010).

    CAS  PubMed  Google Scholar 

  19. Goodrich, L. V., Milenkovic, L., Higgins, K. M. & Scott, M. P. Altered neural cell fates and medulloblastoma in mouse patched mutants. Science 277, 1109–1113 (1997).

    CAS  PubMed  Google Scholar 

  20. Zeman, M. K. & Cimprich, K. A. Causes and consequences of replication stress. Nat. Cell Biol. 16, 2–9 (2014).

    CAS  PubMed  PubMed Central  Google Scholar 

  21. Zheng, D. Q., Zhang, K., Wu, X. C., Mieczkowski, P. A. & Petes, T. D. Global analysis of genomic instability caused by DNA replication stress in Saccharomyces cerevisiae. Proc. Natl Acad. Sci. USA 113, E8114–E8121 (2016).

    CAS  PubMed  PubMed Central  Google Scholar 

  22. Arlt, M. F. et al. Replication stress induces genome-wide copy number changes in human cells that resemble polymorphic and pathogenic variants. Am. J. Hum. Genet. 84, 339–350 (2009).

    CAS  PubMed  PubMed Central  Google Scholar 

  23. Di Micco, R. et al. Oncogene-induced senescence is a DNA damage response triggered by DNA hyper-replication. Nature 444, 638–642 (2006).

    CAS  PubMed  Google Scholar 

  24. Mille, F. et al. The Shh receptor Boc promotes progression of early medulloblastoma to advanced tumors. Dev. Cell 31, 34–47 (2014).

    CAS  PubMed  Google Scholar 

  25. Lee, Y. et al. ATR maintains select progenitors during nervous system development. EMBO J. 31, 1177–1189 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  26. Lee, Y. et al. Neurogenesis requires TopBP1 to prevent catastrophic replicative DNA damage in early progenitors. Nat. Neurosci. 15, 819–826 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  27. Lang, P. Y. et al. ATR maintains chromosomal integrity during postnatal cerebellar neurogenesis and is required for medulloblastoma formation. Development 143, 4038–4052 (2016).

    CAS  PubMed  PubMed Central  Google Scholar 

  28. Zhong, Y. et al. The level of origin firing inversely affects the rate of replication fork progression. J. Cell Biol. 201, 373–383 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  29. Conti, C. et al. Replication fork velocities at adjacent replication origins are coordinately modified during DNA replication in human cells. Mol. Biol. Cell 18, 3059–3067 (2007).

    CAS  PubMed  PubMed Central  Google Scholar 

  30. Courbet, S. et al. Replication fork movement sets chromatin loop size and origin choice in mammalian cells. Nature 455, 557–560 (2008).

    CAS  PubMed  Google Scholar 

  31. O’Donnell, M., Langston, L. & Stillman, B.Principles and concepts of DNA replication in bacteria, archaea, and eukarya. Cold Spring Harb. Perspect. Biol. 5, a010108 (2013).

    PubMed  PubMed Central  Google Scholar 

  32. Bell, S. P. & Kaguni, J. M.Helicase loading at chromosomal origins of replication. Cold Spring Harb. Perspect. Biol. 5, a010124 (2013).

    PubMed  PubMed Central  Google Scholar 

  33. Montagnoli, A. et al. Identification of Mcm2 phosphorylation sites by S-phase-regulating kinases. J. Biol. Chem. 281, 10281–10290 (2006).

    CAS  PubMed  Google Scholar 

  34. Ilves, I., Petojevic, T., Pesavento, J. J. & Botchan, M. R. Activation of the MCM2-7 helicase by association with Cdc45 and GINS proteins. Mol. Cell 37, 247–258 (2010).

    CAS  PubMed  PubMed Central  Google Scholar 

  35. Pruitt, S. C., Bailey, K. J. & Freeland, A. Reduced Mcm2 expression results in severe stem/progenitor cell deficiency and cancer. Stem Cells 25, 3121–3132 (2007).

    CAS  PubMed  Google Scholar 

  36. Macheret, M. & Halazonetis, T. D. Intragenic origins due to short G1 phases underlie oncogene-induced DNA replication stress. Nature 555, 112–116 (2018).

    CAS  PubMed  PubMed Central  Google Scholar 

  37. Tubbs, A. et al. Dual roles of poly(dA:dT) tracts in replication initiation and fork collapse. Cell 174, 1127–1142 e19 (2018).

    CAS  PubMed  PubMed Central  Google Scholar 

  38. Hiratani, I. et al. Global reorganization of replication domains during embryonic stem cell differentiation. PLoS Biol. 6, e245 (2008).

    PubMed  PubMed Central  Google Scholar 

  39. Macheret, M. & Halazonetis, T. D. Monitoring early S-phase origin firing and replication fork movement by sequencing nascent DNA from synchronized cells. Nat. Protoc. 14, 51–67 (2019).

    CAS  PubMed  Google Scholar 

  40. Sukup-Jackson, M. R. et al. Rosa26-GFP direct repeat (RaDR-GFP) mice reveal tissue- and age-dependence of homologous recombination in mammals in vivo. PLoS Genet. 10, e1004299 (2014).

    PubMed  PubMed Central  Google Scholar 

  41. Cavalli, F. M. G. et al. Intertumoral heterogeneity within medulloblastoma subgroups. Cancer Cell 31, 737–754 e6 (2017).

    CAS  PubMed  PubMed Central  Google Scholar 

  42. Tamayo-Orrego, L., Swikert, S. M. & Charron, F.Evasion of cell senescence in SHH medulloblastoma. Cell Cycle 15, 2102–2107 (2016).

    CAS  PubMed  PubMed Central  Google Scholar 

  43. Archer, T. C. et al. Proteomics, post-translational modifications, and integrative analyses reveal molecular heterogeneity within medulloblastoma subgroups. Cancer Cell 34, 396–410.e8 (2018).

    CAS  PubMed  PubMed Central  Google Scholar 

  44. Iwai, K. et al. Molecular mechanism and potential target indication of TAK-931, a novel CDC7-selective inhibitor. Sci. Adv. 5, eaav3660 (2019).

    CAS  PubMed  PubMed Central  Google Scholar 

  45. Siddiqui, K., On, K. F. & Diffley, J. F.Regulating DNA replication in eukarya. Cold Spring Harb. Perspect. Biol. 5, a012930 (2013).

    PubMed  PubMed Central  Google Scholar 

  46. Bartkova, J. et al. Oncogene-induced senescence is part of the tumorigenesis barrier imposed by DNA damage checkpoints. Nature 444, 633–637 (2006).

    CAS  PubMed  Google Scholar 

  47. Viale, A. et al. Cell-cycle restriction limits DNA damage and maintains self-renewal of leukaemia stem cells. Nature 457, 51–56 (2009).

    CAS  PubMed  Google Scholar 

  48. Walter, D. et al. Exit from dormancy provokes DNA-damage-induced attrition in haematopoietic stem cells. Nature 520, 549–552 (2015).

    PubMed  Google Scholar 

  49. Bester, A. C. et al. Nucleotide deficiency promotes genomic instability in early stages of cancer development. Cell 145, 435–446 (2011).

    CAS  PubMed  PubMed Central  Google Scholar 

  50. Bartek, J., Lukas, C. & Lukas, J. Checking on DNA damage in S phase. Nat. Rev. Mol. Cell Biol. 5, 792–804 (2004).

    CAS  PubMed  Google Scholar 

  51. Maya-Mendoza, A. et al. High speed of fork progression induces DNA replication stress and genomic instability. Nature 559, 279–284 (2018).

    CAS  PubMed  Google Scholar 

  52. Anglana, M., Apiou, F., Bensimon, A. & Debatisse, M. Dynamics of DNA replication in mammalian somatic cells: nucleotide pool modulates origin choice and interorigin spacing. Cell 114, 385–394 (2003).

    CAS  PubMed  Google Scholar 

  53. Mantiero, D., Mackenzie, A., Donaldson, A. & Zegerman, P. Limiting replication initiation factors execute the temporal programme of origin firing in budding yeast. EMBO J. 30, 4805–4814 (2011).

    CAS  PubMed  PubMed Central  Google Scholar 

  54. Barlow, J. H. et al. Identification of early replicating fragile sites that contribute to genome instability. Cell 152, 620–632 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  55. Poli, J. et al. dNTP pools determine fork progression and origin usage under replication stress. EMBO J. 31, 883–894 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  56. Gilbert, D. M. Replication origin plasticity, Taylor-made: inhibition vs recruitment of origins under conditions of replication stress. Chromosoma 116, 341–347 (2007).

    PubMed  Google Scholar 

  57. Forsburg, S. L. Eukaryotic MCM proteins: beyond replication initiation. Microbiol. Mol. Biol. Rev. 68, 109–131 (2004).

    CAS  PubMed  PubMed Central  Google Scholar 

  58. Takahashi, T. S., Basu, A., Bermudez, V., Hurwitz, J. & Walter, J. C. Cdc7–Drf1 kinase links chromosome cohesion to the initiation of DNA replication in Xenopus egg extracts. Genes Dev. 22, 1894–1905 (2008).

    CAS  PubMed  PubMed Central  Google Scholar 

  59. Jiang, J. & Hui, C. C. Hedgehog signaling in development and cancer. Dev. Cell 15, 801–812 (2008).

    CAS  PubMed  PubMed Central  Google Scholar 

  60. Briscoe, J. & Therond, P. P. The mechanisms of Hedgehog signalling and its roles in development and disease. Nat. Rev. Mol. Cell Biol. 14, 416–429 (2013).

    PubMed  Google Scholar 

  61. Izzi, L. et al. Boc and Gas1 each form distinct Shh receptor complexes with Ptch1 and are required for Shh-mediated cell proliferation. Dev. Cell 20, 788–801 (2011).

    CAS  PubMed  PubMed Central  Google Scholar 

  62. Lukas, C., Falck, J., Bartkova, J., Bartek, J. & Lukas, J. Distinct spatiotemporal dynamics of mammalian checkpoint regulators induced by DNA damage. Nat. Cell Biol. 5, 255–260 (2003).

    CAS  PubMed  Google Scholar 

  63. Gallo, D., Wang, G., Yip, C. M. & Brown, G. W. Analysis of replicating yeast chromosomes by DNA combing. Cold Spring Harb. Protoc. 2016, pdb.prot085118 (2016).

    PubMed  Google Scholar 

  64. Mendez, J. & Stillman, B. Chromatin association of human origin recognition complex, cdc6, and minichromosome maintenance proteins during the cell cycle: assembly of prereplication complexes in late mitosis. Mol. Cell. Biol. 20, 8602–8612 (2000).

    CAS  PubMed  PubMed Central  Google Scholar 

  65. Bayani, J. & Squire, J. A. Preparation of cytogenetic specimens from tissue samples. Curr. Protoc. Cell Biol. 22, 22.2 (2004).

    Google Scholar 

  66. Hoch, N. C. et al. XRCC1 mutation is associated with PARP1 hyperactivation and cerebellar ataxia. Nature 541, 87–91 (2017).

    CAS  PubMed  Google Scholar 

  67. Martynoga, B., Morrison, H., Price, D. J. & Mason, J. O. Foxg1 is required for specification of ventral telencephalon and region-specific regulation of dorsal telencephalic precursor proliferation and apoptosis. Dev. Biol. 283, 113–127 (2005).

    CAS  PubMed  Google Scholar 

  68. Lohse, M. et al. RobiNA: a user-friendly, integrated software solution for RNA-seq-based transcriptomics. Nucleic Acids Res. 40, W622–W627 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  69. Dobin, A. et al. STAR: ultrafast universal RNA-seq aligner. Bioinformatics 29, 15–21 (2013).

    CAS  PubMed  Google Scholar 

  70. Anders, S., Pyl, P. T. & Huber, W. HTSeq—a Python framework to work with high-throughput sequencing data. Bioinformatics 31, 166–169 (2015).

    CAS  PubMed  Google Scholar 

  71. Dreszer, T. R. et al. The UCSC Genome Browser database: extensions and updates 2011. Nucleic Acids Res. 40, D918–D923 (2012).

    CAS  PubMed  Google Scholar 

  72. Trapnell, C. et al. Transcript assembly and quantification by RNA-seq reveals unannotated transcripts and isoform switching during cell differentiation. Nat. Biotechnol. 28, 511–515 (2010).

    CAS  PubMed  PubMed Central  Google Scholar 

  73. Love, M. I., Huber, W. & Anders, S. Moderated estimation of fold change and dispersion for RNA-seq data with DESeq2. Genome Biol. 15, 550 (2014).

    PubMed  PubMed Central  Google Scholar 

  74. Langmead, B. & Salzberg, S. L. Fast gapped-read alignment with Bowtie 2. Nat. Methods 9, 357–359 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  75. Li, H. et al. The Sequence Alignment/Map format and SAMtools. Bioinformatics 25, 2078–2079 (2009).

    PubMed  PubMed Central  Google Scholar 

  76. Zhang, Y. et al. Model-based analysis of ChIP-Seq (MACS). Genome Biol. 9, R137 (2008).

    PubMed  PubMed Central  Google Scholar 

  77. Quinlan, A. R. & Hall, I. M. BEDTools: a flexible suite of utilities for comparing genomic features. Bioinformatics 26, 841–842 (2010).

    CAS  PubMed  PubMed Central  Google Scholar 

  78. Shen, R. & Seshan, V. E. FACETS: allele-specific copy number and clonal heterogeneity analysis tool for high-throughput DNA sequencing. Nucleic Acids Res. 44, e131 (2016).

  79. Dominguez-Sola, D. et al. Non-transcriptional control of DNA replication by c-Myc. Nature 448, 445–451 (2007).

    CAS  PubMed  Google Scholar 

  80. Chan, K. L. et al. Replication stress induces sister-chromatid bridging at fragile site loci in mitosis. Nat. Cell Biol. 11, 753–760 (2009).

    CAS  PubMed  Google Scholar 

  81. Lukas, C. et al. 53BP1 nuclear bodies form around DNA lesions generated by mitotic transmission of chromosomes under replication stress. Nat. Cell Biol. 13, 243–253 (2011).

    CAS  PubMed  Google Scholar 

Download references

Acknowledgements

We thank A. Helness, S. Terouz, E. Massicotte, F. Depault, J. Cardin and J. Barthe for expert assistance, A. Dumont and M. Rondeau for RNA-seq library preparation and O. Neyret for EdU-seq library preparation. We thank O. Jumanca and the Montreal Clinical Research Institute animal facility staff for animal handling. We thank E. Drobetsky for critical reading of the manuscript. We thank the International Cancer Genome Consortium for providing access to human MB data. L.T.-O. was supported by the Caldas (Colciencias) and Djavad Mowafaghian Foundation fellowships and is now supported by an EMBO Long-Term Fellowship (ALTF-739–2019). D.G. was supported by a post-graduate scholarship from the Natural Sciences and Engineering Research Council of Canada. This work was supported by the Canadian Institutes of Health Research (FDN334023 to F.C.), Fonds de Recherche du QuébecSanté (to F.C.), Canada Foundation for Innovation (33768 to F.C.), Canadian Cancer Society Research Institute (Impact grant 702310 to G.W.B. and Innovation grant 705366 to T.H.), Medical Research Council (MC_UU_00007/5 to A.P.J.) and European Union’s Horizon 2020 Research and Innovation Programme ERC Advanced Grant (number 788093 to A.P.J.). F.C. holds the Canada Research Chair in Developmental Neurobiology.

Author information

Authors and Affiliations

Authors

Contributions

L.T.-O. and F.C. conceived of the study. L.T.-O., D.G., F.R., S.M., S.S. and B.H. performed the experiments. A.B. performed the in silico analyses. T.H. provided reagents and advice. D.G. and L.T.-O. analyzed the data. A.P.J. provided critical advice on the manuscript and supported parts of the work, as well as providing support to L.T.-O., during revision. L.T.-O., G.W.B., A.P.J. and F.C. wrote the manuscript.

Corresponding author

Correspondence to Frédéric Charron.

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 Shh, but not other GCP mitogens, causes replication stress.

a, Most γ-H2AX positive GCPs are actively replicating. 90% of BrdU-positive cells are positive for γ-H2AX, but only 14% of BrdU-negative cells are γ-H2AX positive. Two-sided Fisher’s exact test, n= 530 cells from 3 experiments (428 cells were BrdU+ and 102 were BrdU-). b, Flow cytometry scatter plot showing γ-H2AX (y axis) and DNA content (x axis) of dissociated GCPs; gate shows S-phase cells, n=3844 cells analyzed. Data from 2 independent experiments. c, BrdU and 53bp1 immunostaining of GCPs actively proliferating in response to Shh in the external granule-cell layer of the cerebellum, n=2 experiments; scale bars: 20µm and 5µm (inset). d, Images of GCPs pulsed with CldU and IdU and stained for γ-H2AX, representative of quantifications shown in eg. e, Percentage of γ-H2AX-positive and γ-H2AX-negative cells in S-phase, G2 and G0/1; two-sided Fisher’s exact test. f, Number of γ-H2AX foci/cell in S-phase, G2 and G0/1; one-way anova and Tukey post-test, median and inter-quartile range. g, Relative γ-H2AX levels in S-phase, G2 and G0/1 GCPs, mean±95% CI; n=264 S-phase cells, n= 55 G2 cells, and n=87 G0/G1 cells from 3 experiments (eg). h, Percentage of 53bp1-positive GCPs in S- and G2-phases of the cell cycle (left) and representative images (right); cells were pulsed sequentially with 25µM CldU for 1.5hr and 125µM IdU for 20min and fixed; IdU-positive cells are in S-phase, while CldU-positive and IdU-negative cells are in G2; two-sided Fisher’s exact test; n=241 S-phase cells and 32 G2 cells from 3 experiments. Scale bar, 5µm (d,h). i, Percentage of Ki67 (left) and p-histone H3 (pH3) positive cells in GCPs treated with vehicle (Ctl), bFgf, Egf, Igf1 or Shh; mean±sem; the number of images used was: n= 9 for Ctl, n= 7 for bFgf, n= 8 for Egf, n= 9 for Igf1, and n=10 for Shh from 2 experiments . j, Relative γ-H2AX levels in GCPs treated with vehicle (Ctl), bFgf, Egf or Shh; n= 223 cells for Ctl, n= 201 cells for bFgf, n= 200 cells for Egf, and n=163 cells for Shh, from 2 experiments. k, Relative γ-H2AX levels in BrdU+ cells (S-phase) obtained by immunofluorescence of GCPs treated with vehicle, Igf1 and Shh (mean±95% CI,); n= 34 cells in Ctl-, 60 cells in Igf1-, and 115 cells in Shh-treated samples; images related to Fig. 2e; scale bar, 25µm. Data representative of 3 independent experiments. l, Example of a Rpa70-positive cytokinesis bridge in GCPs treated with 500 nM hydroxyurea; experiment performed two times; cytokinesis failure or ultrafine anaphase bridges (UFBs) were not detected in Shh-treated GCPs.

Source data

Extended Data Fig. 2 Replicative γ-H2AX levels are elevated in the wild-type EGL compared to the intestinal epithelium, and RNA sequencing of Ctl-, Shh- and Igf1-treated GCPs.

a,b, Representative IHC (a) and IF (b) images of the intestinal epithelium and EGL stained for γ-H2AX (a) or for γ-H2AX and Ki67 (b). c, Percentage of proliferating (Ki67+) cells positive for γ-H2AX in the intestinal epithelium and EGL; two-sided Fisher’s exact test. d, Number of γ-H2AX foci/cell in proliferating (Ki67+) intestinal epithelium and EGL; mean±95% CI, two-sided t-test; n=589 intestinal cells, and n= 1290 EGL cells from 3 experiments (bd). e, Principal component analysis (PCA) of differentially-expressed genes between Ctl-, Shh- and Igf1-treated GCPs; n=3 experiments. f, Examples of Igf1 and Shh-regulated genes; mean±sem (n=3 experiments); one-way anova and Tukey post-test. *, p≤0.05; **, p≤0.01; ***, p≤0.001. Source data for the RNAseq results shown in f can be found in Supplementary Table 1.

Source data

Extended Data Fig. 3 Nucleosides are sufficient to induce origin firing and γ-H2AX in GCPs.

a, GSEA (Gene-set enrichment analysis) plot of pyrimidine metabolism in Shh vs. Ctl GCPs, n=3 experiments. b, mRNA levels of representative nucleotide metabolism genes from a; mean±sem, one-way anova and Tukey post-test, n=3 experiments, *, p≤0.05; **, p≤0.01; ***, p≤0.001. c, Mass spectrometry analysis of nucleotide levels in Ctl-, Shh- and Igf1-treated GCPs, n=4 experiments; mean±sem, one-way anova and Tukey post-test, *, p≤0.05; **, p≤0.01; ***, p≤0.001. d, Flow cytometry scatter plot of Ctl- and nucleoside (AGCT)-treated GCPs; n= 63329 Ctl cells and n= 61648 AGCT cells analyzed from 2 experiments. e, BrdU fluorescence in S-phase (BrdU+) GCPs in control and nucleoside-treated GCPs; n= 36 Ctl- and n= 39 AGCT-treated cells from a representative experiment, two-tailed t-test. f, Western blot for γ-H2AX of GCPs treated with Ctl, nucleosides and Shh; bar graph shows quantification of γ-H2AX levels relative to actin (n=4 experiments; mean±sem). g, S-phase time (hrs.) of GCPs treated with nucleosides and/or Shh (n=2 experiments). h, DNA fork rate (kbp/min) of control and nucleoside-treated GCPs, two-sided Mann-Whitney test, median; n=190 forks in Ctl- and n=283 forks in nucleoside-treated GCPs. i, Average DNA fork density of Ctl- and nucleoside-treated GCPs; 300Mb of combed DNA were scored, n=2 experiments (g, i). Source data for the RNAseq results shown in b can be found in Supplementary Table 1.

Source data

Extended Data Fig. 4 Shh regulates the pre-replication complex.

a, Relative fork density in GCPs in response to Ctl, Igf1, Shh, Shh+Cdc7i and nucleosides; mean±95% CI, two-tailed t-test; the number of experiments per condition is: n= 6 for Ctl, n= 2 for nucleosides, n= 2 for Igf1, n= 4 for Shh, and n= 2 for Shh+Cdc7i. b–e, Relative mRNA levels of origin recognition complex (Orc) subunits (b), licensing and preinitiation complex factors (c), components of the CMG (Cdc45-Mcm-Gins) helicase (d) and helicase-activating kinases (e) in Ctl-, Shh- and Igf1-stimulated GCPs; mean±sem, n=3 experiments, one-way anova and Tukey post-test, *, p≤0.05; **, p≤0.01; ***, p≤0.001. f, Flow cytometry scatter plot displaying chromatin-bound Mcm2 (y-axis) and DNA content (Dapi, x-axis) in Ctl- and Shh-treated GCPs; n=2 experiments. g, Histogram of chromatin-bound Mcm2 in the G0/G1 population (gate on f), n=2 experiments. h, Merge images of immunofluorescence shown in Fig. 4c, n=4 experiments. Source data for the RNAseq results shown in b-e can be found in Supplementary Table 1.

Source data

Extended Data Fig. 5 Shh promotes helicase activation.

a, Chromatin-bound p-Mcm2 ser40/41 in asynchronous and S-phase (BrdU+) Ctl-, Cdc7i-, Shh- and Shh+Cdc7i-treated GCPs; mean±95%CI; n= 275 cells for Ctl-, n=130 cells for Cdc7i-, n= 263 cells for Shh-, and n= 277 cells for Shh+Cdc7i-treated samples. For BrdU+ cells: n= 44 in Ctl, n= 20 in Cdc7i, n=71 in Shh, and n=65 in Shh+Cdc7i samples. Data from 2 independent experiments. b, Chromatin-bound Gins2 in asynchronous and S-phase (BrdU+) Ctl-, Cdc7i-, Shh- and Shh+Cdc7i-treated GCPs; mean±95%CI, n= 353 cells for Ctl-, n=291 cells for Cdc7i-, n= 322 cells for Shh-, and n= 191 cells for Shh+Cdc7i-treated samples; for BrdU+ cells: n= 112 in Ctl, n= 75 in Cdc7i, n=160 in Shh, and n=90 in Shh+Cdc7i samples; data from 2 independent experiments. Scale bar, 5µm (a, b). c, DNA synthesis (BrdU fluorescence/cell) in S-phase cells in Ctl, Shh or Shh+Cdc7i-treated GCPs; mean±95% CI; one-way anova, n= 31 cells in Ctl, n= 70 in Shh and n= 28 in Shh+Cdc7i samples. Data from 2 independent experiments. d, S-phase time in Shh and Shh+Cdc7i treated GCPs; median, two-tailed t-test, n= 23 images in Shh and n=21 images in Shh+Cdc7i; data from 2 independent experiments.

Source data

Extended Data Fig. 6 Shh induces replication initiation.

a, Percentage of BrdU+ cells in Ctl-, Cdc7i-, Shh- and Shh+Cdc7i-treated GCPs; mean±95% CI, one-way anova and Tukey post-test, n=20 images in Ctl- and n=10 in Cdc7-, Shh-, and Shh+Cdc7i-treated samples from two experiments. b, Fork speed (kbp/min) in Ctl-, Cdc7i-, Shh- and Shh+Cdc7i-treated GCPs; median, Kruskal-Wallis test; n= 285 forks in Ctl, n=120 forks in Cdc7i, n= 256 forks in Shh, and n= 224 forks in Shh+Cdc7i. Data from 2 independent experiments. c, Inter-origin distance (IOD) in Ctl-, Cdc7i-2-, Shh- and Shh+Cdc7i-2-treated GCPs; Cdc7i-2 is 200nM Tak-931 added during the last 4hr. of the experiment; Kruskal-Wallis test; median and inter-quartile range; median IOD indicated below graph; number of IOD measured is: n= 49 in Ctl-, n=108 in Cdc7i-2-, n=173 in Shh-, and n=146 in Shh+Cdc7i-2-treated samples; experiment performed one time. d, Top, composite plot of EdU peak signal (Fpkm); bottom, composite plot for nucleotide frequency (%) centered around EdU peak. e, Annotation of genomic regions containing EdU-seq peaks in Ctl- and Shh-treated GCPs according to sequence composition. f, Annotation of regions containing peaks according to sequence composition for repetitive and non-repetitive sequences; n=942 peaks for Ctl samples, and n= 4321 peaks for Shh samples from 2 independent experiments (df). g, Distance (kbp) between replication initiation zones in Ctl-, Shh and Shh+Cdc7i-treated GCPs; Median and inter-quartile range (IQR); one-way anova; average also indicated (+) in g; n= 2627 distances in Ctl, 4466 distances in Shh, and 3137 distances in Ctl+Cdc7i samples; data obtained from one experiment.

Source data

Extended Data Fig. 7 Shh promotes homologous recombination.

a, GSEA plot for double strand break (DSB) repair enriched genes in Shh-treated compared to control GCPs, n=3 animals/group. b, GSEA plot for the Fanconi anemia pathway in Shh-treated vs. control GCPs, n=3 animals/group. c, mRNA levels of non-homologous end-joining genes in Ctl-, Shh- and Igf1-stimulated GCPs; mean±sem, one-way anova and Tukey post-test, n=3 animals/group; *, p≤0.05; **, p≤0.01; ***, p≤0.001. d, mRNA levels of homologous recombination genes in Ctl-, Shh- and Igf1-stimulated GCPs; mean±sem, n=3 animals/group, one-way anova; *, p≤0.05; **, p≤0.01; ***, p≤0.001. e, Rad51 total nuclear levels assessed by immunofluorescence in BrdU- or BrdU+ GCPs treated with control or Shh; mean±95% CI; n=207 Ctl and n=210 Shh (asynchronous cells); n= 61 in Ctl and n=110 Shh (BrdU+ cells). f, Chromatin-bound Rad51 levels assessed by immunofluorescence in BrdU- or BrdU+ GCPs treated with control or Shh; n= 92 Ctl and n=97 Shh (BrdU-negative cells), n=56 Ctl and n=51 Shh (BrdU+ cells); mean±95% CI, one-way anova, Tukey post-test. Data from 2 independent experiments (e, f). Representative images on the right; scale bar, 5µm. g, Number of 53bp1 foci/cell in S-phase (BrdU+) Ctl- and Shh-treated GCPs; mean±95%CI, two-sided t-test, n= 44 Ctl and n= 35 Shh S-phase cells from 2 experiments. h, SCEs/metaphase in GCPs treated with Ctl, Cdc7i, bFgF, Egf and Shh; mean±95%CI, one-way anova, Tukey post-test, n= 20 in Ctl and in Cdc7i, n= 45 in bFgf, n= 50 in Egf, and n= 29 metaphases in Shh, from 2 experiments. i, Percent RaDR-GFP-positive events indicative of recombination events in Ctl- and Shh-treated GCPs; n=4 experiments, mean, paired t-test. Source data for the RNAseq results shown in c-d can be found in Supplementary Table 1.

Source data

Extended Data Fig. 8 High CDC7 levels characterize SHH-MB-α and are an indicator of poor prognosis in MB.

a, Overall survival analysis of 612 human medulloblastoma patients from the 4 different MB groups extracted from the Cavalli cohort and analyzed using R2; Kaplan-Meier test. b, Overall survival analysis of 172 human SHH-MB patients from the Cavalli cohort and analyzed using R2; Kaplan-Meier test. c, Dot plot of CDC7 expression in the four subtypes of SHH-MB. d, CDC7 expression in the four subtypes of SHH-MB; median and range, one-way anova test. e, GSEA plots generated from genes upregulated in SHH-MB-α versus SHH-MB-βγδ; n=612 human MBs. f, Top genes enriched in the DNA replication and DNA recombination gene sets presented in e.

Extended Data Fig. 9 Proteomics data uncovers a subtype (SHHa) of human MB characterized by DNA replication and MCM2 activation.

a, Heatmap displaying normalized enrichment scores (NES) of DNA replication and recombination gene ontologies for two subtypes of SHH-MB (SHHa and SHHb) based on proteomics data obtained from Ref. 43. b, NES of DNA recombination comparing SHHa and SHHb at the mRNA and protein level; n=9 SHHa tumors and n=5 SHHb tumors; median and range. c, d, mRNA (c) and protein (d) levels of MCM2 in SHHa and SHHb; n=10 SHHa tumors and n=5 SHHb tumors; median and range. e, Relative peptide abundance of different MCM2 phosphorylation events indicative of helicase activation (Ser40, Ser139 and Ser26/27) in SHHa and SHHb; n=10 SHHa tumors and n=5 SHHb tumors; median and range.

Extended Data Fig. 10 In vivo Cdc7 inhibition does not affect growth or cerebellum development but reduces helicase activation.

a, b, Comparison of mouse body weight at P4 and P7 in Ctl- and Cdc7i-treated pups; mean±sem, two-sided t-test. n= 16 Ctl- and n= 19 Cdc7i-treated P4 mice (a); n= 14 Ctl- and n= 18 Cdc7i-treated P7 mice (b). c, d, IGL area (c) and perimeter (d) at midline in Ctl- and Cdc7i-treated P16 mice (n=6 mice); mean±sem, two-sided t-test. e, EGL thickness in Ctl- and Cdc7i-treated P7 mice; mean±sem, two-sided t-test, n=5 mice/group. f, g Number of phospho-histone-H3 (pH3)-positive cells per 103 µm2 in Ctl- and Cdc7i-treated P7 mice (f) and representative images (g); mean±sem, two-sided t-test; n=5 animals/group. h, Percentage of BrdU+ GCPs in the EGL of the cerebellum of Ctl- and Cdc7i-treated mice; mean±sem, two-sided t-test; n=5 animals/group; scale bar, 100µm. i, Ki67 immunocytochemistry in Ctl- and Cdc7i-treated P7 mice. j, k, Quantitation of p-Mcm2 levels (j) and representative images (k) in the EGL of Ctl- and Cdc7i-treated mice; mean±sem (n=5 mice), two-sided t-test.

Source data

Supplementary information

Reporting Summary

43018_2020_94_MOESM2_ESM.xlsx

Supplementary Table 1 Additional source data for the RNA-seq analyses shown in Extended Data Figs. 2–4 and 7. FPKM and normalized expression in control-, Igf1- and Shh-treated GCPs. The table presents data for genes involved in Igf1 and Shh signaling (Extended Data Fig. 2f), nucleotide metabolism genes (Extended Data Fig. 3b), origin recognition protein complex subunits, licensing factors, CMG helicase and helicase-activating kinases (Extended Data Fig. 4b–e) and non-homologous end-joining and homologous recombination genes (Extended Data Fig. 7c,d). Data are from three independent experiments (mean ± s.e.m.), analyzed by one-way ANOVA and Tukey’s post-hoc test.

Source data

Source Data Fig. 1

Statistical source data.

Source Data Fig. 2

Statistical source data.

Source Data Fig. 3

Statistical source data.

Source Data Fig. 4

Statistical source data.

Source Data Fig. 4

Unprocessed blots.

Source Data Fig. 5

Statistical source data.

Source Data Fig. 5

Unprocessed blots.

Source Data Fig. 6

Statistical source data.

Source Data Fig. 7

Statistical source data.

Source Data Fig. 8

Statistical source data.

Source Data Extended Data Fig. 1

Statistical source data.

Source Data Extended Data Fig. 2

Statistical source data.

Source Data Extended Data Fig. 3

Statistical source data.

Source Data Extended Data Fig. 4

Statistical source data.

Source Data Extended Data Fig. 4

Unprocessed blots.

Source Data Extended Data Fig. 5

Statistical source data.

Source Data Extended Data Fig. 6

Statistical source data.

Source Data Extended Data Fig. 7

Statistical source data.

Source Data Extended Data Fig. 10

Statistical source data.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Tamayo-Orrego, L., Gallo, D., Racicot, F. et al. Sonic hedgehog accelerates DNA replication to cause replication stress promoting cancer initiation in medulloblastoma. Nat Cancer 1, 840–854 (2020). https://doi.org/10.1038/s43018-020-0094-7

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s43018-020-0094-7

This article is cited by

Search

Quick links

Nature Briefing: Cancer

Sign up for the Nature Briefing: Cancer newsletter — what matters in cancer research, free to your inbox weekly.

Get what matters in cancer research, free to your inbox weekly. Sign up for Nature Briefing: Cancer