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:

K27M in canonical and noncanonical H3 variants occurs in distinct oligodendroglial cell lineages in brain midline gliomas

Abstract

Canonical (H3.1/H3.2) and noncanonical (H3.3) histone 3 K27M-mutant gliomas have unique spatiotemporal distributions, partner alterations and molecular profiles. The contribution of the cell of origin to these differences has been challenging to uncouple from the oncogenic reprogramming induced by the mutation. Here, we perform an integrated analysis of 116 tumors, including single-cell transcriptome and chromatin accessibility, 3D chromatin architecture and epigenomic profiles, and show that K27M-mutant gliomas faithfully maintain chromatin configuration at developmental genes consistent with anatomically distinct oligodendrocyte precursor cells (OPCs). H3.3K27M thalamic gliomas map to prosomere 2-derived lineages. In turn, H3.1K27M ACVR1-mutant pontine gliomas uniformly mirror early ventral NKX6-1+/SHH-dependent brainstem OPCs, whereas H3.3K27M gliomas frequently resemble dorsal PAX3+/BMP-dependent progenitors. Our data suggest a context-specific vulnerability in H3.1K27M-mutant SHH-dependent ventral OPCs, which rely on acquisition of ACVR1 mutations to drive aberrant BMP signaling required for oncogenesis. The unifying action of K27M mutations is to restrict H3K27me3 at PRC2 landing sites, whereas other epigenetic changes are mainly contingent on the cell of origin chromatin state and cycling rate.

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: Unique cell type hierarchies in H3.1 and H3.3K27M HGGs.
Fig. 2: Chromatin architecture of HOX clusters implicates distinct progenitor domain origins.
Fig. 3: H3.3K27M thalamic gliomas arise from the thalamus proper.
Fig. 4: H3.1K27M ACVR1-mutant gliomas mirror a SHH-specified NKX6-1+ progenitor.
Fig. 5: NKX6-1 is activated in H3.1K27M HGG.
Fig. 6: ACVR1 mutations confer oncogenic BMP signaling in H3.1K27M HGG.
Fig. 7: H3K27M and EZHIP converge to restrict H3K27me3 to PRC2 nucleation sites.
Fig. 8: Uncoupling the effect of histone variants from cell of origin chromatin state and cycling rate.

Similar content being viewed by others

Data availability

ChIP-seq sequencing data for human cell lines and scRNA-seq sequencing data for normal E10, E13, E16 and E18 murine samples have been deposited in the Gene Expression Omnibus (GEO) under accession number GSE188625, whereas E12, E15, P0, P3 and P6 samples have been previously deposited to GEO under GSE133531. Bulk RNA-seq, ChIP-seq, Hi-C, scRNA-seq, scATAC-seq and scMultiome sequencing data for human tumors have been deposited in the European Genome-phenome Archive (EGA) under accession number EGAS00001005773. Processed data for bulk RNA-seq (counts and differential expression analyses), ChIP-seq (genome-wide H3K27ac/me2/me3 levels) and scRNA-seq/scATAC-seq/scMultiome (counts matrices, cell annotations, and chromatin accessibility bigWig files) have been deposited to GEO under the accession number GSE210568 and Zenodo at https://doi.org/10.5281/zenodo.6773261 (ref. 105). Accession numbers for previously published data used in this study are provided in Supplementary Tables 1, 2, 6 and 11. Source data are provided with this paper.

Code availability

Code to reproduce the main results included in the paper is available at https://github.com/fungenomics/HGG-oncohistones and archived on Zenodo at https://doi.org/10.5281/zenodo.6647837 (ref. 106).

References

  1. Krug, B., Harutyunyan, A. S., Deshmukh, S. & Jabado, N. Polycomb repressive complex 2 in the driver’s seat of childhood and young adult brain tumours. Trends Cell Biol. 31, 814–828 (2021).

    CAS  Google Scholar 

  2. Khuong-Quang, D.-A. et al. K27M mutation in histone H3.3 defines clinically and biologically distinct subgroups of pediatric diffuse intrinsic pontine gliomas. Acta Neuropathol. 124, 439–447 (2012).

    CAS  Google Scholar 

  3. Schwartzentruber, J. et al. Driver mutations in histone H3.3 and chromatin remodelling genes in paediatric glioblastoma. Nature 482, 226–231 (2012).

    CAS  Google Scholar 

  4. Pajtler, K. W. et al. Molecular heterogeneity and CXorf67 alterations in posterior fossa group A (PFA) ependymomas. Acta Neuropathol. 136, 211–226 (2018).

    CAS  Google Scholar 

  5. Wu, G. et al. Somatic histone H3 alterations in pediatric diffuse intrinsic pontine gliomas and non-brainstem glioblastomas. Nat. Genet. 44, 251–253 (2012).

    CAS  Google Scholar 

  6. Lewis, P. W. et al. Inhibition of PRC2 activity by a gain-of-function H3 mutation found in pediatric glioblastoma. Science 340, 857–861 (2013).

    CAS  Google Scholar 

  7. Bender, S. et al. Reduced H3K27me3 and DNA hypomethylation are major drivers of gene expression in K27M mutant pediatric high-grade gliomas. Cancer Cell 24, 660–672 (2013).

    CAS  Google Scholar 

  8. Venneti, S. et al. Evaluation of histone 3 lysine 27 trimethylation (H3K27me3) and enhancer of Zest 2 (EZH2) in pediatric glial and glioneuronal tumors shows decreased H3K27me3 in H3F3A K27M mutant glioblastomas. Brain Pathol. 23, 558–564 (2013).

    CAS  Google Scholar 

  9. Jain, S. U. et al. PFA ependymoma-associated protein EZHIP inhibits PRC2 activity through a H3 K27M-like mechanism. Nat. Commun. 10, 2146 (2019).

    Google Scholar 

  10. Ragazzini, R. et al. EZHIP constrains Polycomb Repressive Complex 2 activity in germ cells. Nat. Commun. 10, 3858 (2019).

    Google Scholar 

  11. Hübner, J.-M. et al. EZHIP/CXorf67 mimics K27M mutated oncohistones and functions as an intrinsic inhibitor of PRC2 function in aggressive posterior fossa ependymoma. Neuro. Oncol. 21, 878–889 (2019).

    Google Scholar 

  12. Jain, S. U. et al. H3 K27M and EZHIP impede H3K27-methylation spreading by inhibiting allosterically stimulated PRC2.Mol. Cell 80, 726–735 (2020).

    CAS  Google Scholar 

  13. Mackay, A. et al. Integrated molecular meta-analysis of 1,000 pediatric high-grade and diffuse intrinsic pontine glioma. Cancer Cell 32, 520–537.e5 (2017).

    CAS  Google Scholar 

  14. Fontebasso, A. M. et al. Recurrent somatic mutations in ACVR1 in pediatric midline high-grade astrocytoma. Nat. Genet. 46, 462–466 (2014).

    CAS  Google Scholar 

  15. Taylor, K. R. et al. Recurrent activating ACVR1 mutations in diffuse intrinsic pontine glioma. Nat. Genet. 46, 457–461 (2014).

    CAS  Google Scholar 

  16. Wu, G. et al. The genomic landscape of diffuse intrinsic pontine glioma and pediatric non-brainstem high-grade glioma. Nat. Genet. 46, 444–450 (2014).

    CAS  Google Scholar 

  17. Sievers, P. et al. A subset of pediatric-type thalamic gliomas share a distinct DNA methylation profile, H3K27me3 loss and frequent alteration of EGFR. Neuro. Oncol. 23, 34–43 (2021).

    CAS  Google Scholar 

  18. Buczkowicz, P. et al. Genomic analysis of diffuse intrinsic pontine gliomas identifies three molecular subgroups and recurrent activating ACVR1 mutations. Nat. Genet. 46, 451–456 (2014).

    CAS  Google Scholar 

  19. Fortin, J. et al. Mutant ACVR1 arrests glial cell differentiation to drive tumorigenesis in pediatric gliomas.Cancer Cell 37, 308–323 (2020).

    CAS  Google Scholar 

  20. Hoeman, C. M. et al. ACVR1 R206H cooperates with H3.1K27M in promoting diffuse intrinsic pontine glioma pathogenesis. Nat. Commun. 10, 1023 (2019).

    Google Scholar 

  21. Vladoiu, M. C. et al. Childhood cerebellar tumours mirror conserved fetal transcriptional programs. Nature, https://doi.org/10.1038/s41586-019-1158-7 (2019).

  22. Filbin, M. G. et al. Developmental and oncogenic programs in H3K27M gliomas dissected by single-cell RNA-seq. Science 360, 331–335 (2018).

    CAS  Google Scholar 

  23. Monje, M. et al. Hedgehog-responsive candidate cell of origin for diffuse intrinsic pontine glioma. Proc. Natl Acad. Sci. U. S. A. 108, 4453–4458 (2011).

    CAS  Google Scholar 

  24. Kessaris, N. et al. Competing waves of oligodendrocytes in the forebrain and postnatal elimination of an embryonic lineage. Nat. Neurosci. 9, 173–179 (2006).

    CAS  Google Scholar 

  25. Fogarty, M., Richardson, W. D. & Kessaris, N. A subset of oligodendrocytes generated from radial glia in the dorsal spinal cord. Development 132, 1951–1959 (2005).

    CAS  Google Scholar 

  26. Cai, J. et al. Generation of oligodendrocyte precursor cells from mouse dorsal spinal cord independent of Nkx6 regulation and Shh signaling. Neuron 45, 41–53 (2005).

    CAS  Google Scholar 

  27. Lin, G. L. & Monje, M. Understanding the deadly silence of posterior fossa A ependymoma. Mol. Cell 78, 999–1001 (2020).

    CAS  Google Scholar 

  28. Piunti, A. et al. Therapeutic targeting of polycomb and BET bromodomain proteins in diffuse intrinsic pontine gliomas. Nat. Med. 23, 493–500 (2017).

    CAS  Google Scholar 

  29. Sarthy, J. F. et al. Histone deposition pathways determine the chromatin landscapes of H3.1 and H3.3 K27M oncohistones. Elife 9, e61090 (2020).

    CAS  Google Scholar 

  30. Nagaraja, S. et al. Histone variant and cell context determine H3K27M reprogramming of the enhancer landscape and oncogenic state. Mol. Cell 76, 965–980.e12 (2019).

    CAS  Google Scholar 

  31. Stafford, J. M. et al. Multiple modes of PRC2 inhibition elicit global chromatin alterations in H3K27M pediatric glioma. Sci. Adv. 4, eaau5935 (2018).

    CAS  Google Scholar 

  32. Mack, S. C. et al. Therapeutic targeting of ependymoma as informed by oncogenic enhancer profiling. Nature 553, 101–105 (2018).

    CAS  Google Scholar 

  33. Krug, B. et al. Pervasive H3K27 acetylation leads to ERV expression and a ttherapeutic vulnerability in H3K27M gliomas. Cancer Cell 35, 782–797.e8 (2019).

    CAS  Google Scholar 

  34. Jessa, S. et al. Stalled developmental programs at the root of pediatric brain tumors. Nat. Genet., https://doi.org/10.1038/s41588-019-0531-7 (2019).

  35. Harutyunyan, A. S. et al. H3K27M in gliomas causes a one-step decrease in H3K27 methylation and reduced spreading within the constraints of H3K36 methylation. Cell Rep. 33, 108390 (2020).

    CAS  Google Scholar 

  36. Chen, C. C. L. et al. Histone H3.3G34-mutant interneuron progenitors co-opt PDGFRA for gliomagenesis. Cell, https://doi.org/10.1016/j.cell.2020.11.012 (2020).

  37. Dubois, F. P. B. et al. Structural variants shape driver combinations and outcomes in pediatric high-grade glioma.Nat. Cancer 3, 994–1011 (2022).

    CAS  Google Scholar 

  38. Harutyunyan, A. S. et al. H3K27M induces defective chromatin spread of PRC2-mediated repressive H3K27me2/me3 and is essential for glioma tumorigenesis. Nat. Commun. 10, 1262 (2019).

    Google Scholar 

  39. Kotliar, D. et al. Identifying gene expression programs of cell-type identity and cellular activity with single-cell RNA-Seq. Elife 8, e43803 (2019).

    Google Scholar 

  40. Bressan, R. B. et al. Regional identity of human neural stem cells determines oncogenic responses to histone H3.3 mutants. Cell Stem Cell, https://doi.org/10.1016/j.stem.2021.01.016 (2021).

  41. Funato, K. et al. Dissecting the impact of regional identity and the oncogenic role of human-specific NOTCH2NL in an hESC model of H3.3G34R-mutant glioma.Cell Stem Cell 28, 894–905 (2021).

    CAS  Google Scholar 

  42. Eze, U. C., Bhaduri, A., Haeussler, M., Nowakowski, T. J. & Kriegstein, A. R. Single-cell atlas of early human brain development highlights heterogeneity of human neuroepithelial cells and early radial glia. Nat. Neurosci., https://doi.org/10.1038/s41593-020-00794-1 (2021).

  43. Bhaduri, A. et al. An atlas of cortical arealization identifies dynamic molecular signatures. Nature 598, 200–204 (2021).

    CAS  Google Scholar 

  44. Jacquet, B. V. et al. FoxJ1-dependent gene expression is required for differentiation of radial glia into ependymal cells and a subset of astrocytes in the postnatal brain. Development 136, 4021–4031 (2009).

    CAS  Google Scholar 

  45. Lee, T. I. et al. Control of developmental regulators by Polycomb in human embryonic stem cells. Cell 125, 301–313 (2006).

    CAS  Google Scholar 

  46. Bernstein, B. E. et al. A bivalent chromatin structure marks key developmental genes in embryonic stem cells. Cell 125, 315–326 (2006).

    CAS  Google Scholar 

  47. Philippidou, P. & Dasen, J. S. Hox genes: choreographers in neural development, architects of circuit organization. Neuron 80, 12–34 (2013).

    CAS  Google Scholar 

  48. Noordermeer, D. et al. The dynamic architecture of Hox gene clusters. Science 334, 222–225 (2011).

    CAS  Google Scholar 

  49. Narendra, V. et al. CTCF establishes discrete functional chromatin domains at the Hox clusters during differentiation. Science 347, 1017–1021 (2015).

    CAS  Google Scholar 

  50. Scholpp, S. & Lumsden, A. Building a bridal chamber: development of the thalamus. Trends Neurosci. 33, 373–380 (2010).

    CAS  Google Scholar 

  51. Saint-André, V. et al. Models of human core transcriptional regulatory circuitries. Genome Res. 26, 385–396 (2016).

    Google Scholar 

  52. Uhlén, M. et al. Proteomics. Tissue-based map of the human proteome. Science 347, 1260419 (2015).

    Google Scholar 

  53. Thompson, C. L. et al. A high-resolution spatiotemporal atlas of gene expression of the developing mouse brain. Neuron 83, 309–323 (2014).

    CAS  Google Scholar 

  54. Lein, E. S. et al. Genome-wide atlas of gene expression in the adult mouse brain. Nature 445, 168–176 (2007).

    CAS  Google Scholar 

  55. Visel, A. et al. A high-resolution enhancer atlas of the developing telencephalon. Cell 152, 895–908 (2013).

    CAS  Google Scholar 

  56. Oosterveen, T. et al. Mechanistic differences in the transcriptional interpretation of local and long-range Shh morphogen signaling. Dev. Cell 23, 1006–1019 (2012).

    CAS  Google Scholar 

  57. Vallstedt, A., Klos, J. M. & Ericson, J. Multiple dorsoventral origins of oligodendrocyte generation in the spinal cord and hindbrain. Neuron 45, 55–67 (2005).

    CAS  Google Scholar 

  58. Masahira, N. et al. Olig2-positive progenitors in the embryonic spinal cord give rise not only to motoneurons and oligodendrocytes, but also to a subset of astrocytes and ependymal cells. Dev. Biol. 293, 358–369 (2006).

    CAS  Google Scholar 

  59. Aibar, S. et al. SCENIC: single-cell regulatory network inference and clustering. Nat. Methods 14, 1083–1086 (2017).

    CAS  Google Scholar 

  60. Valer, J. A., Sánchez-de-Diego, C., Pimenta-Lopes, C., Rosa, J. L. & Ventura, F. ACVR1 function in health and disease. Cells 8, 1366 (2019).

    CAS  Google Scholar 

  61. Ferrari, K. J. et al. Polycomb-dependent H3K27me1 and H3K27me2 regulate active transcription and enhancer fidelity. Mol. Cell 53, 49–62 (2014).

    CAS  Google Scholar 

  62. Zhu, C. et al. Joint profiling of histone modifications and transcriptome in single cells from mouse brain. Nat. Methods, https://doi.org/10.1038/s41592-021-01060-3 (2021).

  63. Oksuz, O. et al. Capturing the onset of PRC2-mediated repressive domain formation. Mol. Cell 70, 1149–1162.e5 (2018).

    CAS  Google Scholar 

  64. Lindquist, R. A. et al. Identification of proliferative progenitors associated with prominent postnatal growth of the pons. Nat. Commun. 7, 11628 (2016).

    CAS  Google Scholar 

  65. Nishiyama, A., Shimizu, T., Sherafat, A. & Richardson, W. D. Life-long oligodendrocyte development and plasticity. Semin. Cell Dev. Biol. 116, 25–37 (2021).

    CAS  Google Scholar 

  66. Gonçalves, C. S., Le Boiteux, E., Arnaud, P. & Costa, B. M. HOX gene cluster (de)regulation in brain: from neurodevelopment to malignant glial tumours. Cell. Mol. Life Sci. 77, 3797–3821 (2020).

    Google Scholar 

  67. Shah, N. & Sukumar, S. The Hox genes and their roles in oncogenesis. Nat. Rev. Cancer 10, 361–371 (2010).

    CAS  Google Scholar 

  68. Le Boiteux, E. et al. Widespread overexpression from the four DNA hypermethylated HOX clusters in aggressive (IDHwt) glioma is associated with H3K27me3 depletion and alternative promoter usage. Mol. Oncol. 15, 1995–2010 (2021).

    Google Scholar 

  69. Bond, A. M., Bhalala, O. G. & Kessler, J. A. The dynamic role of bone morphogenetic proteins in neural stem cell fate and maturation. Dev. Neurobiol. 72, 1068–1084 (2012).

    CAS  Google Scholar 

  70. Stuart, T. et al. Comprehensive integration of single-cell data. Cell 177, 1888–1902 (2019).

    CAS  Google Scholar 

  71. McInnes, L. & Healy, J. UMAP: Uniform manifold approximation and projection for dimension reduction. Preprint at arXiv, https://arxiv.org/abs/1802.03426 (2018).

  72. Waltman, L. & van Eck, N. J. A smart local moving algorithm for large-scale modularity-based community detection.The Eur. Phys. J. B 86, 71 (2013).

    Google Scholar 

  73. Nagy, C. et al. Single-nucleus transcriptomics of the prefrontal cortex in major depressive disorder implicates oligodendrocyte precursor cells and excitatory neurons. Nat. Neurosci. 23, 771–781 (2020).

    CAS  Google Scholar 

  74. Tirosh, I. et al. Dissecting the multicellular ecosystem of metastatic melanoma by single-cell RNA-seq. Science 352, 189–196 (2016).

    CAS  Google Scholar 

  75. Stuart, T., Srivastava, A., Madad, S., Lareau, C. A. & Satija, R. Single-cell chromatin state analysis with Signac. Nat. Methods 18, 1333–1341 (2021).

    CAS  Google Scholar 

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

    Google Scholar 

  77. Ramírez, F. et al. deepTools2: a next generation web server for deep-sequencing data analysis. Nucleic Acids Res. 44, W160–W165 (2016).

    Google Scholar 

  78. Kramer, N. E. et al. Plotgardener: cultivating precise multi-panel figures in R. bioRxiv https://doi.org/10.1101/2021.09.08.459338 (2021).

    Article  Google Scholar 

  79. Thorvaldsdóttir, H., Robinson, J. T. & Mesirov, J. P. Integrative Genomics Viewer (IGV): high-performance genomics data visualization and exploration. Brief. Bioinform. 14, 178–192 (2013).

    Google Scholar 

  80. Korsunsky, I. et al. Fast, sensitive and accurate integration of single-cell data with Harmony. Nat. Methods 16, 1289–1296 (2019).

    CAS  Google Scholar 

  81. Li, C. et al. SciBet as a portable and fast single cell type identifier. Nat. Commun. 11, 1818 (2020).

    CAS  Google Scholar 

  82. Hearst, M. A., Dumais, S. T., Osuna, E., Platt, J. & Scholkopf, B. Support vector machines. IEEE Intell. Syst. 13, 18–28 (1998).

    Google Scholar 

  83. Pedregosa, F. et al. Scikit-learn: Machine Learning in Python. J. Mach. Learn. Res. 12, 2825–2830 (2011).

    Google Scholar 

  84. Subramanian, A. et al. Gene set enrichment analysis: a knowledge-based approach for interpreting genome-wide expression profiles. Proc. Natl Acad. Sci. U. S. A. 102, 15545–15550 (2005).

    CAS  Google Scholar 

  85. Liberzon, A. et al. The molecular signatures database (MSigDB) hallmark gene set collection. Cell Syst. 1, 417–425 (2015).

    CAS  Google Scholar 

  86. Kinker, G. S. et al. Pan-cancer single-cell RNA-seq identifies recurring programs of cellular heterogeneity. Nat. Genet. 52, 1208–1218 (2020).

    CAS  Google Scholar 

  87. Bolger, A. M., Lohse, M. & Usadel, B. Trimmomatic: a flexible trimmer for Illumina sequence data. Bioinformatics 30, 2114–2120 (2014).

    CAS  Google Scholar 

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

    CAS  Google Scholar 

  89. Liao, Y., Smyth, G. K. & Shi, W. featureCounts: an efficient general purpose program for assigning sequence reads to genomic features. Bioinformatics 30, 923–930 (2014).

    CAS  Google Scholar 

  90. 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).

    Google Scholar 

  91. Bourgey, M. et al. GenPipes: an open-source framework for distributed and scalable genomic analyses. Gigascience 8, giz037 (2019).

    Google Scholar 

  92. Li, H. & Durbin, R. Fast and accurate short read alignment with Burrows-Wheeler transform. Bioinformatics 25, 1754–1760 (2009).

    CAS  Google Scholar 

  93. Heinz, S. et al. Simple combinations of lineage-determining transcription factors prime cis-regulatory elements required for macrophage and B cell identities. Mol. Cell 38, 576–589 (2010).

    CAS  Google Scholar 

  94. Younesy, H., Möller, T., Lorincz, M. C., Karimi, M. M. & Jones, S. J. M. VisRseq: R-based visual framework for analysis of sequencing data. BMC Bioinf. 16, S2 (2015).

    Google Scholar 

  95. Orlando, D. A. et al. Quantitative ChIP-seq normalization reveals global modulation of the epigenome. Cell Rep. 9, 1163–1170 (2014).

    CAS  Google Scholar 

  96. Carroll, T. S., Liang, Z., Salama, R., Stark, R. & de Santiago, I. Impact of artifact removal on ChIP quality metrics in ChIP-seq and ChIP-exo data. Front. Genet. 5, 75 (2014).

    Google Scholar 

  97. Weinberg, D. N. et al. The histone mark H3K36me2 recruits DNMT3A and shapes the intergenic DNA methylation landscape. Nature https://doi.org/10.1038/s41586-019-1534-3 (2019).

    Article  Google Scholar 

  98. Killick, R. & Eckley, I. A. changepoint: an R package for changepoint analysis.J. Stat. Softw. 58, 1–19 (2014).

    Google Scholar 

  99. Lovén, J. et al. Selective inhibition of tumor oncogenes by disruption of super-enhancers. Cell 153, 320–334 (2013).

    Google Scholar 

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

    CAS  Google Scholar 

  101. Wingett, S. et al. HiCUP: pipeline for mapping and processing Hi-C data. F1000Res. 4, 1310 (2015).

    Google Scholar 

  102. Durand, N. C. et al. Juicer provides a one-click system for analyzing loop-resolution Hi-C experiments. Cell Syst. 3, 95–98 (2016).

    CAS  Google Scholar 

  103. Karch, K. R., Sidoli, S. & Garcia, B. A. Identification and quantification of histone PTMs using high-resolution mass spectrometry. Methods Enzymol. 574, 3–29 (2016).

    CAS  Google Scholar 

  104. Ran, F. A. et al. Double nicking by RNA-guided CRISPR Cas9 for enhanced genome editing specificity. Cell 154, 1380–1389 (2013).

    CAS  Google Scholar 

  105. Jessa, S., Hébert, S. & Kleinman, C. L. HGG-oncohistones processed data (Zenodo, 2022). https://doi.org/10.5281/zenodo.6773261

  106. Jessa, S. et al. HGG-oncohistones analysis code (Zenodo, 2022). https://doi.org/10.5281/zenodo.6647837

Download references

Acknowledgements

We thank the patients and their families for their invaluable contributions to this research, without whom it would be impossible. This work was supported by funding from A Large-Scale Applied Research Project grant from Genome Quebec, Genome Canada, the Government of Canada, and the Ministère de l'Économie, de la Science et de l’Innovation du Québec, with the support of the Ontario Institute for Cancer Research through funding provided by the Government of Ontario to N.J., M.D.T., C.L.K. Fondation Charles Bruneau to N.J., US National Institutes of Health (NIH) (grant P01-CA196539 to N.J. and grants R01CA148699 and R01CA159859 to M.D.T.); the Canadian Institutes for Health Research (CIHR) (grants MOP-286756 and FDN-154307 to N.J. and grant PJT-156086 to C.L.K.); the Canadian Cancer Society (CCSRI) (grant 705182) and the Fonds de Recherche du Québec en Santé (FRQS) salary award to C.L.K.; NSERC (RGPIN-2016-04911) to C.L.K.; CFI Leaders Opportunity Fund 33902 to C.L.K., Genome Canada Science Technology Innovation Centre, Compute Canada Resource Allocation Project (WST-164-AB). Data analyses were enabled by compute and storage resources provided by Compute Canada and Calcul Québec. N.J. is a member of the Penny Cole Laboratory and the recipient of a Chercheur Boursier, Chaire de Recherche Award from the FRQS. This work was performed within the context of the International CHildhood Astrocytoma INtegrated Genomic and Epigenomic (ICHANGE) consortium with funding from Genome Canada and Genome Quebec. S.J. is supported by a fellowship from CIHR. We also acknowledge support from the We Love You Connie, Poppies for Irini and Kat D Strong Foundations (N.J.). The funders had no role in study design, data collection and analysis, decision to publish or preparation of the manuscript. We acknowledge the contributions of D. Marchione and J. Wojcik in MS work.

Author information

Authors and Affiliations

Authors

Contributions

S.J., A.M., A.S.H. and C.L.K. designed and coordinated computational analyses. A.M., A.S.H., N.J. and C.L.K conceptualized experimental work. S.J., A.M., A.S.H., M.H., H.L., S.V., N.K., Z.B., S.H., S.W., M.C., M.B., S.C.M. and C.L.K contributed to computational and statistical data analyses. M.V. performed timed mating and tissue isolation of developing mouse embryos. A.M., A.S.H., M.H., D.F., B.K., A.F.A. and A. Bajic performed experimental work in patient-derived cell lines. D.F. performed primary tissue isolation, sample preparation and optimization of single-cell protocols. B.A.G. coordinated MS. A. Bhaduri provided normal human fetal brain data. M.P., A.G.W., B.E., J.A., R.W.R.D., J.-P.F., S.P., V.L., L.G., K.L.L., P.B., M.D.T., C.L.K. and N.J. contributed to acquisition of resources and primary tumor samples. SJ., A.M., A.S.H., M.H., C.L.K. and N.J. wrote the manuscript with input from L.G. C.L.K. and N.J. jointly supervised the project.

Corresponding authors

Correspondence to Nada Jabado or Claudia L. Kleinman.

Ethics declarations

Competing interests

S.P. is a member of the advisory board for Bayer, Novartis and AstraZeneca and has received speaker fees from Bayer and Esai outside of the submitted work. All other authors declare no competing interests.

Peer review

Peer review information

Nature Genetics thanks Xiao-nan Li and the other, anonymous, reviewer(s) 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.

Extended data

Extended Data Fig. 1 Overview of expanded scRNA-seq mouse developmental reference.

a. Schematic of developing mouse brain, sagittal view, indicating regions and time points included in the single-cell reference atlas. Red: data generated in this study; black: data from Jessa et al., Nature Genetics, 201934. b. Number of cells captured in each time point and brain region after quality control and filtering. c. Overview of single-cell populations from the mouse pons. Dendrogram constructed based on pairwise Spearman correlations between mean expression profiles in each cluster. Cell class and time point are annotated. d. Overview of single-cell populations from the mouse forebrain. Dendrogram constructed based on pairwise Spearman correlations between mean expression profiles in each cluster. Cell class and time point are annotated.

Extended Data Fig. 2 Unique cell type hierarchies in H3.1 and H3.3K27M HGGs.

a. Similarity matrix between all Non-negative Matrix Factorization (NMF) programs assigned to modules. Heatmap represents the pairwise overlap (in number of genes) between programs. b. Annotation of NMF programs. Top: Correlation between each program and QC or biological metrics in each cell. One module (M11) that was explained by technical factors (mitochondrial content and coverage), was consequently removed from further analyses. Bottom: overlap between each program and developmental or MSigDB reference signatures, one line per signature. Only significant overlaps (p-value < 0.001) are shown, and a number of significant overlaps is shown in parentheses. c. Top 15 genes associated with each module. Module-associated genes were selected by identifying the most frequent program-associated genes for all programs contained in the module. d. UMAP for H3.3K27M thalamic HGG (malignant cells only), with cells colored by consensus projected cell type based on the normal mouse brain reference (left), or the normal human fetal thalamus reference (right). Cells are colored as in Fig. 1d. e. Confusion matrix comparing projected cell types for H3.3K27M thalamic HGG based on mouse or human reference. Proportions were computed row-wise and represent the fraction of cells from each mouse label which were assigned to each human label. Bubbles are scaled to the number of cells with each combination of labels.

Extended Data Fig. 3 Some H3.1K27M pontine gliomas arbor a malignant ependymal-like component.

a-b. UMAP plots for two individual H3.1K27M pontine gliomas containing ependymal-like cells. Only malignant cells are shown. Cells are colored by consensus projected cell type. c-d. Heatmaps of copy-number signal computed for each individual sample using InferCNV. Row annotations correspond to cell type projections, indicating whether they are projected to ependymal cells (left, red), and the overall projected cell class, with colors as in (a) and normal cells colored in gray. Cells lacking a consensus projection were excluded. e-f. UMAP plots as in (a-b), with cells colored by expression of FOXJ1 (ependymal transcription factor), DNAH12 (ciliary gene), and single-cell gene set enrichment (ssGSEA) score of candidate FOXJ1 targets in the early postnatal mouse brain. List of FOXJ1 targets was obtained from Jacquet et al., Development, 200944. g. NMF programs from Fig. 1c, displaying only the overlap between program-associated genes with ependymal gene signatures, and filtering out all other developmental signatures. Top column annotation shows the driver alteration of the sample in which each program was identified. A second annotation highlighting H3.1/2K27M tumors is included for clarity. Module 10, significantly overlapping ependymal signatures, is enriched for programs from this tumor entity. h-i. Activity of the ependymal-related module 10 in individual samples. Top: UMAP plots as in (a-b), cells are colored by the NMF activity score of the module 10 program from each sample. Bottom: heatmap of NMF score of module 10 program-associated genes; names for selected informative genes are indicated.

Extended Data Fig. 4 H3.1K27M, ACVR1-mutant pontine gliomas arise from an NKX6-1+ ventral brainstem progenitor.

a. Volcano plot of differentially expressed genes between H3.3K27M pons and H3.3K27M thalamus HGG. HOX genes are indicated in purple. Only genes with mean normalized expression > 100 are included. b. Epigenomic state at NKX6-1 and PAX3 in representative H3.3K27M pons HGG primary tumors and cell lines. For scATAC-seq data, each track represents RPKM-normalized aggregated accessibility for one malignant single-cell population. c. Co-expression of NKX6-1 and PAX3 in bulk RNA-seq data for pons HGG with each K27M histone variant.

Extended Data Fig. 5 Assessment of NKX6-1 in brain tumors and normal tissues.

a Immunohistochemistry staining of NKX6-1 protein in normal pancreas tissue as positive control. Arrowhead in left panel indicates region shown at higher magnification in right panel. b-d. Immunohistochemistry staining of NKX6-1 protein in Histone 3 WT, H3.3G34R, and H3.3K27M high-grade glioma patient tumors. e. Antibody staining of NKX6-1 in human tissues from the Human Protein Atlas. Detection levels for each cell type are indicated below, ‘-’ indicates that NKX6-1 was not detected. Image credit: Human Protein Atlas. Images available from http://v21.proteinatlas.org (links provided in Supplementary Table 19). f. Left: In situ hybridization (ISH) in E13.5 mouse brain from the Allen Brain Atlas (© 2008 Allen Institute for Brain Science. Allen Developing Mouse Brain Atlas. Available from: developingmouse.brain-map.org). Right: quantification of ISH expression levels. g. Left: ISH in P56 mouse brain from the Allen Brain Atlas (© 2004 Allen Institute for Brain Science. Allen Mouse Brain Atlas. Available from mouse.brain-map.org). Right: quantification of ISH expression levels. h. Bulk expression levels of NKX6-1 in adult human tissues from GTEx. Sample sizes for brain tissues are indicated.

Extended Data Fig. 6 Nkx6-1/Pax3 expression is mutually exclusive in the normal brain.

a. Expression of Nkx6-1 and Pax3 in cell types of the normal developing mouse pons reference, showing their expression is largely mutually exclusive. The number of cells where both Nkx6-1 and Pax3 are detected out of the total number of Nkx6-1+ or Pax3+ cells of the cell type is indicated in parentheses. b. Expression of Nkx6-1 target genes with high cell type specificity in ependymal cells. Dendrogram represents cell clusters in the single-cell mouse pons reference, as in Extended Data Figure S1. c. Cell type specificity score for inferred targets of Nkx6-1 and Pax3 in the normal mouse pons. For a given gene, score represents the difference between the highest detection rate of the gene in any single-cell cluster in the normal mouse reference, and the detection rate of the gene in all other cells in the same sample (see Methods).

Extended Data Fig. 7 H3K27M and EZHIP converge to restrict H3K27me3 to PRC2 nucleation sites.

a. Percentage of H3K27me3-marked 10 kb bins overlapping CGIs or SUZ12 peaks in cell lines and tumors. Number of biologically independent samples per group is indicated in parentheses. H3K27me3 was quantified in 10 kb bins genome-wide and the top 1% bins with highest H3K27me3 in each sample were intersected with CGIs/SUZ12 peaks. For SUZ12, the union of peaks called from SUZ12 ChIP-seq in BT245 and DIPGXIII were obtained from Harutyunyan et al., Nature Communications, 201938. Crossbar indicates the median. P-values: left panel (H3.1K27M vs WT GBM, p = 0.024; H3.3K27M vs WT GBM, p = 0.026; PFA-EP vs WT GBM, 0.00099); right panel (H3.1K27M vs WT GBM, p = 0.063; H3.3K27M vs WT GBM, p = 0.023; PFA-EP vs WT GBM, p = 0.0016); n.s., not significant; Welch two-sample t-test. b. Scatterplots of H3K27me2 signal over 100 kb bins genome-wide in pairwise group comparisons. X- and Y- axes represent log2 mean RPKM value per group, normalized by input. Marked bins (mean RPKM > 1 in at least one of the groups in each comparison) are shown in black, while unmarked bins are shown in gray. Joint density and marginal distributions are calculated over marked bins only. Red line indicates the diagonal. RPKM values of H3K27me2 were divided by the respective input sample RPKM and averaged for all samples in the same mutation group using a geometric mean. c. H3K27me3 (top) and H3K27me2 (bottom) ChIP-seq enrichment tracks, in representative K27M-mutant and isogenic CRISPR-KO cell lines. d. Mass spectrometry data of H3K27ac in cell. Number of biologically independent samples per group is indicated in parentheses. Error bars represent mean +/− SD. P-values: H3.1K27M vs WT GBM, 0.0074; H3.3K27M vs WT GBM: 5.3×10-5; n.s., not significant; Welch two-sample t-test. e. Enrichment of H3K27ac over different repeat element families in HGG cell lines and isogenic K27M-KO counterparts.

Extended Data Fig. 8 Cell-of-origin chromatin state contributes to the tumor epigenome.

a. Schematic of analysis. Single-cell epigenomic data for normal mouse OPCs and ependymal cells was obtained from Zhu et al., Nature Biotechnology, 202166, and used to extract cell type-specific epigenomic features. Tumors were clustered based on H3K27ac levels at promoters of these genes. b. Hierarchical clustering of H3.1K27M HGG, H3.1K27M PFA-EP, and EZHIP PFA-EP based on OPC and ependymal-specific epigenomic features. Select features are indicated. c. Top: RNA and single-cell epigenomic data for normal mouse OPCs and ependymal cells66 at ependymal and OPC genes. Bottom: H3K27ac ChIPseq tracks for H3.1K27M HGG, H3.1K27M PFA-EP, and EZHIP PFA-EP at the same genes as in the top panel. Chromosome coordinates are indicated in Supplementary Table 15.

Extended Data Fig. 9 Uncoupling the effect of histone variants from cell-of-origin chromatin state and cycling rate.

a. Validation of CRISPR removal of ACVR1 in H3.1K27M ACVR1-mutant cell lines by MiSeq (multiple deletions on both alleles (complete KO)). b. Validation of CRISPR removal of H3K27M in H3.1K27M cell lines DIPGIV and DIPG36 (1 bp deletion on K27M allele (frameshift)) and DIPG21 (2 bp deletion on K27M allele) by MiSeq and Western Blot. For Western Blot, G477, an H3.1 WT HGG patient-derived cell line, was used as control. CRISPR removal of H3K27M in H3.3K27M cell lines has been reported previously for BT245 and DIPGXIII in Krug et al., Cancer Cell, 2019 33; and for HSJ019 in Harutyunyan et al., Cell Reports, 2020 35. c. Doubling time of H3.3K27M and H3.1K27M HGG cell lines (DIPGXIII, N = 4 biological replicates; HSJ019, N = 3; DIPG36, N = 9; DIPGIV, N = 12). Error bars represent mean +/− SD. d. Doubling time of H3.1K27M cell line DIPGIV in ACVR1 mutant and ACVR1-KO conditions. Error bars represent mean +/− SD. e. Schematic of experimental design. f. Heatmap showing distribution of Rx-normalized ChIPseq signal for H3K27me3 in DIPGXIII at CpG islands (CGIs), flanked by 20kbp on either side. g. Rx-normalized H3K27me3 tracks in each condition at a representative genomic region. Y-axis limit is indicated in brackets and identical for all tracks. h. Left: Rx-normalized H3K27me2 tracks in each condition at the same region as in (g). Y-axis limit is indicated in brackets and identical for all tracks. Right: genome-wide distribution of H3K27me2 domain length in each condition (H3.3K27M, N = 16,630 domains; K27M-KO, N = 3388; H3.1K27M, N = 11,568). i. Heatmap showing distribution of Rx-normalized ChIPseq signal for H3K27me2 in DIPGXIII H3K27me2 domains across the genome in each condition. Domains are scaled to 50 kb and flanked by 50 kb on either side. The maximum of the color scale is set to the 90th percentile value across all data points.

Source data

Supplementary information

Supplementary Information

Supplementary Protocol 1.

Reporting Summary

Supplementary Table 1

Supplementary Tables 1–20 and associated legends.

Source data

Source Data Extended Data Fig. 9

Uncropped scan of western blot for Extended Data Fig. 9.

Rights and permissions

Springer Nature or its licensor (e.g. a society or other partner) holds exclusive rights to this article under a publishing agreement with the author(s) or other rightsholder(s); author self-archiving of the accepted manuscript version of this article is solely governed by the terms of such publishing agreement and applicable law.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Jessa, S., Mohammadnia, A., Harutyunyan, A.S. et al. K27M in canonical and noncanonical H3 variants occurs in distinct oligodendroglial cell lineages in brain midline gliomas. Nat Genet 54, 1865–1880 (2022). https://doi.org/10.1038/s41588-022-01205-w

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41588-022-01205-w

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