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:

Nuclear condensates of YAP fusion proteins alter transcription to drive ependymoma tumourigenesis

Abstract

Nuclear localization of HIPPO-YAP fusion proteins has been implicated in supratentorial ependymoma development. Here, unexpectedly, we find that liquid–liquid phase separation, rather than nuclear localization, of recurrent patient-derived YAP fusions, YAP-MAMLD1 and C11ORF95-YAP, underlies ependymoma tumourigenesis from neural progenitor cells. Mutagenesis and chimaera assays demonstrate that an intrinsically disordered region promotes oligomerization of the YAP fusions into nuclear, puncta-like, membrane-less condensates. Oligomerization and nuclear condensates induced by YAP fusion with a coiled-coil domain of transcriptional activator GCN4 also promote ependymoma formation. YAP-MAMLD1 concentrates transcription factors and co-activators, including BRD4, MED1 and TEAD, in condensates while excluding transcriptional repressive PRC2, and induces long-range enhancer–promoter interactions that promote transcription and oncogenic programmes. Blocking condensate-mediated transcriptional co-activator activity inhibits tumourigenesis, indicating a critical role of liquid phase separation for YAP fusion oncogenic activity in ependymoma. YAP fusions containing the intrinsically disordered region features are common in human tumours, suggesting that nuclear condensates could be targeted to treat YAP-fusion-induced cancers.

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: YAP patient-derived fusions, but not nuclear-localized YAP, are sufficient to drive EPN-like tumour formation.
Fig. 2: YAP fusion proteins undergo liquid-like phase separation.
Fig. 3: IDR is required for phase separation and tumourigenesis.
Fig. 4: Oligomerization of YAP induces phase separation.
Fig. 5: IDR mediates exclusion of PRC2 from nuclear puncta and inhibition of PRC2 activity.
Fig. 6: Phase separation mediates oncogenic programmes in YAP-M-expressing cells.
Fig. 7: YAP fusion protein condensates function as scaffolds for transcriptional machinery.
Fig. 8: YAP fusion-induced nuclear condensates alter transcription to drive EPN tumourigenesis.

Similar content being viewed by others

Data availability

Sequencing data that support the findings of this study have been deposited in the NCBI Gene Expression Omnibus (GEO) under accession codes GSE207293 including RNA-seq and ChIP–seq data with the accession number GSE181867, and methyl-seq with accession number GSE206926 and HiChIP with accession number GSE207801. Molecular Signatures Database v5.1 was obtained from the following website: https://www.gsea-msigdb.org/gsea/msigdb. Reference genome mm10 was obtained from the following website: https://hgdownload.soe.ucsc.edu/downloads.html. Source data are provided with this paper. All other data supporting the findings of this study are available from the corresponding author on reasonable request

Code availability

No custom code was used in this study. Open-source algorithms were used as detailed in analysis methods including FastQC v0.11.6 (https://www.bioinformatics.babraham.ac.uk/projects/fastqc/), STAR v2.7.8a (https://github.com/STAR-Fusion/STAR-Fusion/wiki/STAR-Fusion-release-and-CTAT-Genome-Lib-Compatibility-Matrix), Bowtie2 v2.4.0 (https://bowtie-bio.sourceforge.net/bowtie2/index.shtml), DEseq2 1.36.0 (https://bioconductor.org/packages/release/bioc/html/DESeq2.html), MACS v1.4.2 (http://liulab.dfci.harvard.edu/MACS), HOMER (http://homer.salk.edu/homer/index.html), GSEA 4.2.3 (http://www.broadinstitute.org/gsea/index.jsp), Cytoscape 3.9.1 (https://cytoscape.org/), Toppcluster (https://toppcluster.cchmc.org/), bwameth v0.20 (https://github.com/brentp/bwa-meth), MarkDuplicates v1.122.0 (http://broadinstitute.github.io/picard), MethylDackel v0.5.2 (https://github.com/dpryan79/MethylDackel.git), Metilene v0.2-8 (http://www.bioinf.uni-leipzig.de/Software/metilene) and MAPS pipeline for HiChIP–seq (https://github.com/ijuric/MAPS).

References

  1. Wang, Q. et al. Supratentorial pediatric cortical ependymomas: a comprehensive retrospective study. Neurosurg. Rev. 44, 1543–1551 (2021).

    Article  PubMed  Google Scholar 

  2. Hubner, J. M., Kool, M., Pfister, S. M. & Pajtler, K. W. Epidemiology, molecular classification and WHO grading of ependymoma. J. Neurosurg. Sci. 62, 46–50 (2018).

    PubMed  Google Scholar 

  3. Zacharoulis, S. et al. Treatment and outcome of children with relapsed ependymoma: a multi-institutional retrospective analysis. Childs Nerv. Syst. 26, 905–911 (2010).

    Article  PubMed  Google Scholar 

  4. Taylor, M. D. et al. Radial glia cells are candidate stem cells of ependymoma. Cancer Cell 8, 323–335 (2005).

    Article  CAS  PubMed  Google Scholar 

  5. Pajtler, K. W. et al. Molecular classification of ependymal tumors across all CNS compartments, histopathological grades, and age groups. Cancer Cell 27, 728–743 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  6. Johnson, R. A. et al. Cross-species genomics matches driver mutations and cell compartments to model ependymoma. Nature 466, 632–636 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  7. Gojo, J. et al. Single-cell RNA-seq reveals cellular hierarchies and impaired developmental trajectories in pediatric ependymoma. Cancer Cell 38, 44–59.e49 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  8. Zanconato, F., Battilana, G., Cordenonsi, M. & Piccolo, S. YAP/TAZ as therapeutic targets in cancer. Curr. Opin. Pharmacol. 29, 26–33 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  9. Harvey, K. F., Zhang, X. & Thomas, D. M. The Hippo pathway and human cancer. Nat. Rev. Cancer 13, 246–257 (2013).

    Article  CAS  PubMed  Google Scholar 

  10. Hong, W. & Guan, K. L. The YAP and TAZ transcription co-activators: key downstream effectors of the mammalian Hippo pathway. Semin. Cell Dev. Biol. 23, 785–793 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  11. Varelas, X. The Hippo pathway effectors TAZ and YAP in development, homeostasis and disease. Development 141, 1614–1626 (2014).

    Article  CAS  PubMed  Google Scholar 

  12. Pajtler, K. W. et al. YAP1 subgroup supratentorial ependymoma requires TEAD and nuclear factor I-mediated transcriptional programmes for tumorigenesis. Nat. Commun. 10, 3914 (2019).

    Article  PubMed  PubMed Central  Google Scholar 

  13. Parker, M. et al. C11orf95-RELA fusions drive oncogenic NF-κB signalling in ependymoma. Nature 506, 451–455 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  14. Szulzewsky, F. et al. Comparison of tumor-associated YAP1 fusions identifies a recurrent set of functions critical for oncogenesis. Genes Dev. 34, 1051–1064 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  15. Meng, Z., Moroishi, T. & Guan, K. L. Mechanisms of Hippo pathway regulation. Genes Dev. 30, 1–17 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  16. Wu, L. M. N. et al. Programming of Schwann cells by Lats1/2-TAZ/YAP signaling drives malignant peripheral nerve sheath tumorigenesis. Cancer Cell 33, 292–308 (2018).

  17. Zhao, S. et al. PiggyBac transposon vectors: the tools of the human gene encoding. Transl. Lung Cancer Res. 5, 120–125 (2016).

    CAS  PubMed  PubMed Central  Google Scholar 

  18. Dong, J. et al. Elucidation of a universal size-control mechanism in Drosophila and mammals. Cell 130, 1120–1133 (2007).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  19. Zhao, B. et al. Inactivation of YAP oncoprotein by the Hippo pathway is involved in cell contact inhibition and tissue growth control. Genes Dev. 21, 2747–2761 (2007).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  20. Chen, Q. Z. et al. Homeostatic control of Hippo signaling activity revealed by an endogenous activating mutation in YAP. Genes Dev. 29, 1285–1297 (2015).

    Article  PubMed  PubMed Central  Google Scholar 

  21. Andreiuolo, F. et al. Childhood supratentorial ependymomas with YAP1-MAMLD1 fusion: an entity with characteristic clinical, radiological, cytogenetic and histopathological features. Brain Pathol. 29, 205–216 (2019).

    Article  CAS  PubMed  Google Scholar 

  22. Takadera, M. et al. Phenotypic characterization with somatic genome editing and gene transfer reveals the diverse oncogenicity of ependymoma fusion genes. Acta Neuropathol. Commun. 8, 203 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  23. Eder, N. et al. YAP1/TAZ drives ependymoma-like tumour formation in mice. Nat. Commun. 11, 2380 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  24. Ligon, K. L. et al. The oligodendroglial lineage marker OLIG2 is universally expressed in diffuse gliomas. J. Neuropathol. Exp. Neurol. 63, 499–509 (2004).

    Article  CAS  PubMed  Google Scholar 

  25. Lu, F. et al. Olig2-dependent reciprocal shift in PDGF and EGF receptor signaling regulates tumor phenotype and mitotic growth in malignant glioma. Cancer Cell 29, 669–683 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  26. Venugopal, C., McFarlane, N. M., Nolte, S., Manoranjan, B. & Singh, S. K. Processing of primary brain tumor tissue for stem cell assays and flow sorting. J. Vis. Exp. 67, 4111 (2012).

  27. Jiang, S., Fagman, J. B., Chen, C., Alberti, S. & Liu, B. Protein phase separation and its role in tumorigenesis. eLife 9, e60264 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  28. Alberti, S., Gladfelter, A. & Mittag, T. Considerations and challenges in studying liquid–liquid phase separation and biomolecular condensates. Cell 176, 419–434 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  29. Bolognesi, A., Sliwa-Gonzalez, A., Prasad, R. & Barral, Y. Fluorescence recovery after photo-bleaching (FRAP) and fluorescence loss in photo-bleaching (FLIP) experiments to study protein dynamics during budding yeast cell division. Methods Mol. Biol. 1369, 25–44 (2016).

    Article  CAS  PubMed  Google Scholar 

  30. Chong, S. et al. Imaging dynamic and selective low-complexity domain interactions that control gene transcription. Science 361, eaar2555 (2018).

    Article  PubMed  PubMed Central  Google Scholar 

  31. Sabari, B. R. et al. Coactivator condensation at super-enhancers links phase separation and gene control. Science 361, eaar3958 (2018).

    Article  PubMed  PubMed Central  Google Scholar 

  32. Zhao, B. et al. TEAD mediates YAP-dependent gene induction and growth control. Genes Dev. 22, 1962–1971 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  33. Liu-Chittenden, Y. et al. Genetic and pharmacological disruption of the TEAD-YAP complex suppresses the oncogenic activity of YAP. Genes Dev. 26, 1300–1305 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  34. Na Ayutthaya, P. P., Lundberg, D., Weigel, D. & Li, L. Blue native polyacrylamide gel electrophoresis (BN-PAGE) for the analysis of protein oligomers in plants. Curr. Protoc. Plant Biol. 5, e20107 (2020).

    Article  CAS  PubMed  Google Scholar 

  35. Ciani, B. et al. Molecular basis of coiled-coil oligomerization-state specificity. Proc. Natl Acad. Sci. USA 107, 19850–19855 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  36. Banani, S. F., Lee, H. O., Hyman, A. A. & Rosen, M. K. Biomolecular condensates: organizers of cellular biochemistry. Nat. Rev. Mol. Cell Biol. 18, 285–298 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  37. Forsova, O. S. & Zakharov, V. V. High-order oligomers of intrinsically disordered brain proteins BASP1 and GAP-43 preserve the structural disorder. FEBS J. 283, 1550–1569 (2016).

    Article  CAS  PubMed  Google Scholar 

  38. Cao, R. et al. Role of histone H3 lysine 27 methylation in Polycomb-group silencing. Science 298, 1039–1043 (2002).

    Article  CAS  PubMed  Google Scholar 

  39. Laugesen, A., Hojfeldt, J. W. & Helin, K. Molecular mechanisms directing PRC2 recruitment and H3K27 methylation. Mol. Cell 74, 8–18 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  40. Yu, J. R., Lee, C. H., Oksuz, O., Stafford, J. M. & Reinberg, D. PRC2 is high maintenance. Genes Dev. 33, 903–935 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

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

    Article  PubMed  PubMed Central  Google Scholar 

  42. Hanahan, D. & Weinberg, R. A. Hallmarks of cancer: the next generation. Cell 144, 646–674 (2011).

    Article  CAS  PubMed  Google Scholar 

  43. Creyghton, M. P. et al. Histone H3K27ac separates active from poised enhancers and predicts developmental state. Proc. Natl Acad. Sci. USA 107, 21931–21936 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  44. Hnisz, D. et al. Convergence of developmental and oncogenic signaling pathways at transcriptional super-enhancers. Mol. Cell 58, 362–370 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  45. Tang, F., Yang, Z., Tan, Y. & Li, Y. Super-enhancer function and its application in cancer targeted therapy. NPJ Precis. Oncol. 4, 2 (2020).

    Article  PubMed  PubMed Central  Google Scholar 

  46. Wang, X., Cairns, M. J. & Yan, J. Super-enhancers in transcriptional regulation and genome organization. Nucleic Acids Res. 47, 11481–11496 (2019).

    CAS  PubMed  PubMed Central  Google Scholar 

  47. Mumbach, M. R. et al. HiChIP: efficient and sensitive analysis of protein-directed genome architecture. Nat. Methods 13, 919–922 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  48. McKeown, M. R. & Bradner, J. E. Therapeutic strategies to inhibit MYC. Cold Spring Harb. Perspect. Med. 4, a014266 (2014).

    Article  PubMed  PubMed Central  Google Scholar 

  49. Pearson, A. D. et al. Bromodomain and extra-terminal inhibitors—a consensus prioritisation after the Paediatric Strategy Forum for medicinal product development of epigenetic modifiers in children—ACCELERATE. Eur. J. Cancer 146, 115–124 (2021).

    Article  CAS  PubMed  Google Scholar 

  50. Lu, Y. et al. Phase separation of TAZ compartmentalizes the transcription machinery to promote gene expression. Nat. Cell Biol. 22, 453–464 (2020).

    Article  CAS  PubMed  Google Scholar 

  51. Cai, D. et al. Phase separation of YAP reorganizes genome topology for long-term YAP target gene expression. Nat. Cell Biol. 21, 1578–1589 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  52. Hnisz, D., Shrinivas, K., Young, R. A., Chakraborty, A. K. & Sharp, P. A. A phase separation model for transcriptional control. Cell 169, 13–23 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  53. Li, H. et al. SUZ12 promotes human epithelial ovarian cancer by suppressing apoptosis via silencing HRK. Mol. Cancer Res. 10, 1462–1472 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  54. Varambally, S. et al. The polycomb group protein EZH2 is involved in progression of prostate cancer. Nature 419, 624–629 (2002).

    Article  CAS  PubMed  Google Scholar 

  55. 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).

    Article  CAS  PubMed  Google Scholar 

  56. 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).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  57. Bayliss, J. et al. Lowered H3K27me3 and DNA hypomethylation define poorly prognostic pediatric posterior fossa ependymomas. Sci. Transl. Med. 8, 366ra161 (2016).

    Article  PubMed  PubMed Central  Google Scholar 

  58. Li, Y. et al. Genome-wide analyses reveal a role of Polycomb in promoting hypomethylation of DNA methylation valleys. Genome Biol. 19, 18 (2018).

    Article  PubMed  PubMed Central  Google Scholar 

  59. Takeshima, H., Yamashita, S., Shimazu, T., Niwa, T. & Ushijima, T. The presence of RNA polymerase II, active or stalled, predicts epigenetic fate of promoter CpG islands. Genome Res. 19, 1974–1982 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  60. Antonescu, C. R. et al. Novel YAP1-TFE3 fusion defines a distinct subset of epithelioid hemangioendothelioma. Genes Chromosomes Cancer 52, 775–784 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  61. Veeraraghavan, J., Ma, J., Hu, Y. & Wang, X. S. Recurrent and pathological gene fusions in breast cancer: current advances in genomic discovery and clinical implications. Breast Cancer Res. Treat. 158, 219–232 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  62. Szulzewsky, F., Holland, E. C. & Vasioukhin, V. YAP1 and its fusion proteins in cancer initiation, progression and therapeutic resistance. Dev. Biol. 475, 205–221 (2021).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  63. Sekine, S. et al. Recurrent YAP1-MAML2 and YAP1-NUTM1 fusions in poroma and porocarcinoma. J. Clin. Invest. 129, 3827–3832 (2019).

    Article  PubMed  PubMed Central  Google Scholar 

  64. Heallen, T. et al. Hippo pathway inhibits Wnt signaling to restrain cardiomyocyte proliferation and heart size. Science 332, 458–461 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  65. Liao, Y. et al. OLIG2 maintenance is not essential for diffuse intrinsic pontine glioma cell line growth but regulates tumor phenotypes. Neuro Oncol. 23, 1183–1196 (2021).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  66. Lastowska, M. et al. Transcriptional profiling of paediatric ependymomas identifies prognostically significant groups. J. Pathol. Clin. Res. 7, 565–576 (2021).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  67. Wansink, D. G. et al. Fluorescent labeling of nascent RNA reveals transcription by RNA polymerase II in domains scattered throughout the nucleus. J. Cell Biol. 122, 283–293 (1993).

    Article  CAS  PubMed  Google Scholar 

  68. Gao, Y. et al. Single Cas9 nickase induced generation of NRAMP1 knockin cattle with reduced off-target effects. Genome Biol. 18, 13 (2017).

    Article  PubMed  PubMed Central  Google Scholar 

  69. Zhang, L. et al. Single-cell transcriptomics in medulloblastoma reveals tumor-initiating progenitors and oncogenic cascades during tumorigenesis and relapse. Cancer Cell 36, 302–318 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  70. Skene, P. J. & Henikoff, S. An efficient targeted nuclease strategy for high-resolution mapping of DNA binding sites. eLife 16, 21856 (2017).

    Article  Google Scholar 

  71. Juhling, F. et al. metilene: fast and sensitive calling of differentially methylated regions from bisulfite sequencing data. Genome Res. 26, 256–262 (2016).

    Article  PubMed  PubMed Central  Google Scholar 

  72. Juric, I. et al. MAPS: model-based analysis of long-range chromatin interactions from PLAC-seq and HiChIP experiments. PLoS Comput. Biol. 15, e1006982 (2019).

    Article  PubMed  PubMed Central  Google Scholar 

Download references

Acknowledgements

We thank R. J. Gilbertson for C11ORF95-YAP1 construct and R. Tjian for LacO-LacI and RPB1 constructs, R. L Johnson for Lats1/2 floxed mice, X. Zhang for MED1 antibody, X. Zhou for cartoon drawing and Y. Wang for technical assistance and plasmids of CRISPR/Cas9-mediated knock-in. X. Hu and X. Wu performed experiments at the Cincinnati Children’s Hospital, and current affiliation is Fudan University. This study was in part funded by grants from the CancerFree KIDS, Pray-Hope-Believe Foundation and TeamConnor Childhood Cancer Foundation, and CureStartsNow Foundation to Q.R.L.

Author information

Authors and Affiliations

Authors

Contributions

X.H. and Q.R.L. designed the project and analysed data. X.H. and X.W. performed cell, molecular and animal experiments. X.H. and K.B. performed IUE animal experiments. C.Z. performed HiChIP–seq and constructed YAP-NLS. D.X. performed native gel study and CUT&RUN–seq experiments. L.Z. performed ChIP–seq. S.O. performed allografts and drug treatment. X.L. and M.S. performed immunohistochemistry. Z.L. performed methyl-seq and short hairpin RNA knockdown. J.T. and M.L. provided EPN samples. F.S. and E.C.H. provided YAP-TFE3, YAP-SS18 and YAP-FAM118B. L.L. and M.H. analysed HiChIP. M.X. provided resources. X.H. and Q.R.L. wrote and revised the manuscript. Q.R.L. supervised the study.

Corresponding author

Correspondence to Q. Richard Lu.

Ethics declarations

Competing interests

The authors declare no competing interests.

Peer review

Peer review information

Nature Cell Biology thanks Howard Fine, Peter St George-Hyslop 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 The activation of HIPPO-YAP/TAZ in human ependymoma.

a, Enrichment of HIPPO signaling pathway in ependymoma (EPN) relative to high grade glioma (HGG), low grade glioma (LGG), diffuse intrinsic pontine glioma (DIPG), and medulloblastoma (MB). Kolmogorov-Smirnov test. p and q (FDR-corrected p) values were indicated here. b-c, YAP and TAZ (b) YAP target genes CTGF and CYR61 (c) expression in human brain tumors. Black lines indicate mean. EPN (n = 98) stands for ependymoma, HGG (n = 139) for high-grade glioma, LGG (n = 261) for low-grade glioma, DIPG (n = 32) for diffuse intrinsic pontine glioma, and MB (n = 133) for medulloblastoma. Mann-Whitney test. d, Mutation rates of the core components of the HIPPO signaling pathway in human brain tumor. Chi-squared analysis (df=4, n = 2302), p = 1.4e-4. All ‘n’ here referred to number of individual cases.

Extended Data Fig. 2 Activation of HIPPO-YAP/TAZ leads to ependymoma in mice.

a, Diagram of Lats1/2 knockout by transduction of Cre in Lats1fl/f;Lats2fl/fl mice via in utero electroporation (IUE). b, Immunostaining of YAP in the brain of Lats1fl/f;Lats2fl/fl E17.5 embryos electroporated with pCAG-Cre or control constructs. Arrows indicate nuclear puncta formation. Scale bars, 10 μm. c, Hematoxylin and eosin (H&E) staining showing control and Lat1/2 double conditional knockout (cKO) tumors from 4-month-old mice. Scale bars, 1 mm (left) and 20 μm (right). d, Quantification of tumorigenic phenotypes of indicated mice. i, e,g, Immunostaining of ependymoma-related markers (e) and neural stemness-related markers (g) in Lats1/2 cKO tumors. Scale bars, 20 μm. f, h, Quantification of cells positive for ependymoma markers (f) and neural stemness-related markers (h) in Lats1/2 cKO tumors and control tissues. Data are presented as mean ± s.e.m (unpaired two-tailed Student’s t-test). n = 3 independent experiments. i, Sections of cortical tissues from mice transduced with YAP-M or YAP-C and mouse glioma tissue immunostained for OLIG2. Scale bars, 10 μm.

Extended Data Fig. 3 Sphere, allograft formation and methylation profiling of YAP-M fusion-driven mouse tumors.

a, Representative images of neurospheres formed by NSCs isolated from E17.5 embryos of wild-type (NSC) and tumor cells that express YAP fusion proteins YAP-M and YAP-C. Scale bars, 500 μm. b, Percentage of wells containing NSCs, YAP-M-expressing cells, and YAP-C-expressing cells that formed spheres in continuous passages. Data are presented as mean ±s.e.m (unpaired two-tailed Student’s t-test). n = 9 independent experiments. c, Sphere diameters from the fourth passage of NSCs and YAP-M- and YAP-C-expressing cells. Data are presented as mean ± s.e.m (unpaired two-tailed Student’s t-test). n = 30 independent spheres. d, Representative bioluminescence images of the NSG mice engrafted with 1.8×105 and 6×104 YAP-M-expressing tumor cells at 50 day after injection. e, Allograft tumor formation at different doses of YAP-M-expressing tumor cells. f, Heat-map representation of an unsupervised clustering of DNA methylation profiles of human ST-EPN-YAP tumors (n = 13), mouse tumors induced by YAP-MAMLD1 (n = 3) and mouse neural stem cells (NSC) (n = 3) across promoter regions of around top 600 variable methylated gene orthologs among human EPN tumors. The signature genes specific for ST-EPN-YAP are shown in the right side. Each row represents a gene; each column represents a sample.

Source data

Extended Data Fig. 4 YAP-M and YAP-C fusion proteins form liquid-like nuclear condensates.

a, Sections of cortical tumor tissues from mice transduced with YAP-M or YAP-C immunostained for YAP. Scale bar, 10 μm. b, 293 T cells transfected with empty vector or vectors expressing YAP, YAP-S112A, YAP-NLS, YAP-S112A-NLS, YAP-M, and YAP-C fusions immunostained for YAP (red). DAPI (blue) was used to stain nuclei. Scale bar, 5 μm. c, C2C12 cells transfected with YAP-M construct. Scale bar, 20 μm (left) and 5 μm (Right). d, Quantification of nuclear fluorescence intensity of the cells transfected with YAP-derivatives. Data shown are means; n = 30 cells for each group, pooled from 2 independent experiments. one-tailed unpaired t-test. e,f, YAP western blot analysis (e) and quantifications (f) showing the nuclear levels of YAP-NLS, YAP-M or YAP-C. Histone H1, loading control. n = 2 independent experiments with similar results. g, Live-cell imaging of C2C12 cells expressing YAP-C-GFP. The arrows indicate representative YAP-C nuclear puncta that fused over time. This assay was performed twice (two independent transfections) with similar results. Scale bar, 5 μm. h, Time-lapse images of YAP-M-GFP or YAP-C-GFP after 5% 1,6-hexanediol treatment.

Source data

Extended Data Fig. 5 Endogenous YAP-MAMLD1 proteins undergo phase separation.

a, Tissue sections of a human ST-EPN tumor carrying YAP-MAMLD1 fusion and a low-grade glioma (LGG) were immunostained for YAP (green). DAPI (blue), nuclei. Scale bar, 10 μm. b, Scheme for site-specific knock-in (KI) of YAP-M fusion at the YAP1 locus via CRISPR/cas9n. c, HeLa cells with YAP-M knock-in were immunostained for YAP (red) and HA-tagged YAP-M (green). DAPI (blue), nuclei. Scale bar, 5 μm. d, Western blot analysis of nuclear and total protein levels of YAP and YAP-M in HeLa cells without or with knocked-in YAP-M. GAPDH and histone H1 as loading controls. e, Quantification of nuclear and total protein levels of YAP and YAP-M in HeLa cells without (Ctrl) or with YAP-M knock-in. n = 3 independent experiments. Data are presented as mean ± s.e.m (unpaired two-tailed Student’s t-test).

Source data

Extended Data Fig. 6 Disruption of YAP-TEAD interaction does not abolish phase-separation of YAP-M proteins.

a, YAPS94A-MAMLD1, which lost the ability of YAP to interact with TEAD, undergoes phase separation in C2C12 cells. Scale bar, 5 μm. b, YAP-dependent luciferase activity in NIH 3T3 cells expressing 8×GTIIC and YAP-MAMLD1 or YAPS94A-MAMLD1 constructs. Data are means ± s.e.m.; p values were determined using unpaired two-tailed Student’s t-tests; n = 3 biologically independent samples. c, C2C12 cells transfected with YAP-M, treated with 5 μM of verteporfin for 24hrs and stained for YAP (red). DAPI (blue) was used to stain nuclei. Scale bar, 5 μm.

Extended Data Fig. 7 YAP fusions contain IDRs and form oligomers in the nucleus to promote tumorigenesis.

a, Plot of disordered regions in YAP-C fusion identified using PONDR. VSL2 scores are shown on the y axis and positions are shown on the x axis. The blue bar designates the IDR investigated. b, Amino acid compositions of YAP-M (upper) and YAP-C (lower). c, Immunostaining of YAP in the C2C12 cells transfected with YAP-IDR construct. Arrows indicate condensate-like structures. Scale bar, 5 μm. d, C2C12 cells transfected with CC-YAP or CC-YAP-NLS stained for YAP (red). DAPI (blue) was used to stain nuclei. Scale bar, 5 μm. e, Representative bioluminescence images of the mice electroporated with CC-YAP or CC-YAP-NLS during tumor progression. f, H&E staining of brain tumor tissues of the mice that express CC-YAP-NLS. Arrow indicates tumor lesion. Scale bar, 1 mm.

Extended Data Fig. 8 Expression of YAP-M promotes oncogenic programs in NSCs.

a, Volcano plot of transcriptome profiles of YAP-M-expressing NSCs versus YAP-expressing NSCs. Red and green dots represent genes significantly upregulated and downregulated upon YAP-M expression (P < 0.01). Differential expression was determined by FDR-corrected p-value from Wald test. b, GSEA analysis of genes differentially expressed in YAP-M-expressing NSCs (n = 4) versus SVZ of wild-type mice (n = 3). n = biologically independent replicates. c, Pathway enrichment analysis of genes differentially expressed in YAP-M-expressing NSCs versus YAP-expressing NSCs (red) or SVZ of wild-type mice (orange). Distinct pathways and biological processes between the molecular subgroups of ependymal tumors are shown (FDR corrected p < 0.01). Nodes represent enriched gene sets, which are grouped by their similarity. Node size is proportional to the total number of genes within each gene set. The illustrated network map was simplified by manual curation to remove general and uninformative sub-networks. Fisher’s exact test was applied to determine the significance level.

Source data

Extended Data Fig. 9 YAP-M fusion induces long-range super-enhancer interactions to promote oncogenic signaling.

a, Circos map of combined omic profiles of YAP-M tumors. The innermost layer represents oncogenic genes near super-enhancers (SEs). The second layer includes SEs (n = 627). The third layer includes all enhancers (n = 16,381). The fourth layer represents the top 10,000 binding loci of YAP. The fifth layer includes differentially expressed genes in tumors from mice transfected with the YAP-M expression construct relative to NSCs that express YAP. The sixth layer represents the mouse genome (mm10). b, Distribution of H3K27ac binding genomic regions in the indicated cells. c, YAP-M fusion induces long-range enhancer/promoter interactions at the gene loci associated oncogenic signaling. Genomic tracks depicting long-range enhancer/promoter interactions at the indicated loci of oncogenesis-promoting genes in the control and YAP-M transduced NSCs. Pink arcs, DNA contacts; histograms, YAP1 and H3K27ac ChIP-seq peaks. d,e, Co-staining of HA-tagged YAP-M or YAP-MΔIDR (green) and H3K27ac (red) and immunoblot analysis in C2C12 cells transfected with YAP-M or YAP-MΔIDR constructs. Scale bars, 10 μm.

Source data

Extended Data Fig. 10 The effect of pharmacological and genetic inhibition of Brd4 on nuclear puncta formation and tumor growth, and IDR-containing YAP fusions in different cancers.

a, C2C12 cells transfected with YAP-M, treated with 0.1 μM of JQ1 or vehicle and stained for YAP (red). Scale bars, 10 μm. b, Western blot showing the effect of shRNAs to repress BRD4 or MED1 expression. c, Growth of murine NPCs expressing YAP-M fusion was inhibited by knock-down of BRD4 or MED1. 3 independent transfections for each group. Data are presented as mean ± s.e.m. Unpaired student’s two-sided t-test was applied to compare cell numbers at the terminal time point. d, Representative bioluminescence images of the mice electroporated with YAP-M construct after treated with vehicle or BMS-986158 (a chemical inhibitor of Brd4) during tumor progression. e, Survival analysis of mice transfected with the YAP-M expression construct treated with BMS-986158 or vehicle. N = 6 animals/group. Log rank test. f, IDR regions in YAP fusion proteins are predicted in multiple cancer types. g, YAP immunostaining in the HEK293T cells transfected with vectors expressing YAP fusions (YAP1-TFE3, YAP1-SS18 and YAP1-FAM118B) shows nuclear puncta-like structures. Scale bars: 10 μm.

Source data

Supplementary information

Reporting Summary

Supplementary Video 1

Dynamic fusion of nuclear puncta of YAP-MAMLD1-GFP.

Source data

Source Data Fig. 4

Unprocessed western blots.

Source Data Fig. 5

Statistical source data.

Source Data Fig. 5

Unprocessed western blots.

Source Data Fig. 6

Statistical source data.

Source Data Fig. 7

Statistical source data.

Source Data Fig. 7

Unprocessed western blots.

Source Data Extended Data Fig. 3

Statistical source data.

Source Data Extended Data Fig. 4

Unprocessed western blots.

Source Data Extended Data Fig. 5

XH_unmodified_gels_ED_Fig4e.

Source Data Extended Data Fig. 8

Statistical source data.

Source Data Extended Data Fig. 9

Statistical source data.

Source Data Extended Data Fig. 9

Unprocessed western blots.

Source Data Extended Data Fig. 10

Unprocessed western blots.

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

Hu, X., Wu, X., Berry, K. et al. Nuclear condensates of YAP fusion proteins alter transcription to drive ependymoma tumourigenesis. Nat Cell Biol 25, 323–336 (2023). https://doi.org/10.1038/s41556-022-01069-6

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41556-022-01069-6

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