Skip to main content

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

  • Article
  • Published:

A non-canonical SWI/SNF complex is a synthetic lethal target in cancers driven by BAF complex perturbation

Abstract

Mammalian SWI/SNF chromatin remodelling complexes exist in three distinct, final-form assemblies: canonical BAF (cBAF), PBAF and a newly characterized non-canonical complex (ncBAF). However, their complex-specific targeting on chromatin, functions and roles in disease remain largely undefined. Here, we comprehensively mapped complex assemblies on chromatin and found that ncBAF complexes uniquely localize to CTCF sites and promoters. We identified ncBAF subunits as synthetic lethal targets specific to synovial sarcoma and malignant rhabdoid tumours, which both exhibit cBAF complex (SMARCB1 subunit) perturbation. Chemical and biological depletion of the ncBAF subunit, BRD9, rapidly attenuates synovial sarcoma and malignant rhabdoid tumour cell proliferation. Importantly, in cBAF-perturbed cancers, ncBAF complexes maintain gene expression at retained CTCF-promoter sites and function in a manner distinct from fusion oncoprotein-bound complexes. Together, these findings unmask the unique targeting and functional roles of ncBAF complexes and present new cancer-specific therapeutic targets.

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

Access options

Rent or buy this article

Prices vary by article type

from$1.95

to$39.95

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

Fig. 1: mSWI/SNF family complexes are biochemically and functionally distinct.
Fig. 2: Differential localization of the mSWI/SNF complexes, ncBAF, cBAF and PBAF on chromatin.
Fig. 3: Components of ncBAF are selective synthetic lethal dependencies in synovial sarcoma and MRT cell lines.
Fig. 4: Subunit domains of ncBAF underlie complex-specific synthetic lethalities.
Fig. 5: ncBAF is not required for SS18–SSX1-mediated gene expression and primarily regulates fusion-independent sites.
Fig. 6: ncBAF is required for the maintenance of gene expression and retains co-localization with promoters and CTCF in SMARCB1-deficient cancers.

Similar content being viewed by others

Data availability

The ChIP-seq and RNA-seq datasets generated and/or analysed during the current study have been deposited in the Gene Expression Omnibus repository under accession number GSE113042 (https://www.ncbi.nlm.nih.gov/geo/query/acc.cgi?acc=GSE113042). Other datasets that were previously published and used in this study have been deposited in the Gene Expression Omnibus repository under accession numbers GSE90634 and GSE108025 available at https://www.ncbi.nlm.nih.gov/geo/query/acc.cgi?acc=GSE90634 and https://www.ncbi.nlm.nih.gov/geo/query/acc.cgi?acc=GSE108025, respectively. The fitness data were derived from Project Achilles through the Project Achilles Data Portal (https://portals.broadinstitute.org/achilles/about). The dataset derived from this resource that supports the findings of this study is available at https://portals.broadinstitute.org/achilles/datasets/all. The fitness data were also derived from Project DRIVE. The dataset derived from this resource that supports the findings of this study is available at https://oncologynibr.shinyapps.io/drive/. All proteomics/mass-spectrometry data are deposited at the ProteomeXchange Consortium via the PRIDE partner repository with the dataset identifier PXD011103.

References

  1. Narlikar, GeetaJ., Sundaramoorthy, R. & Owen-Hughes, T. Mechanisms and functions of ATP-dependent chromatin-remodeling enzymes. Cell 154, 490–503 (2013).

    Article  CAS  Google Scholar 

  2. Clapier, C. R. & Cairns, B. R. The biology of chromatin remodeling complexes. Annu. Rev. Biochem. 78, 273–304 (2009).

    Article  CAS  Google Scholar 

  3. Ho, L. et al. An embryonic stem cell chromatin remodeling complex, esBAF, is essential for embryonic stem cell self-renewal and pluripotency. Proc. Natl Acad. Sci. USA 106, 5181–5186 (2009).

    Article  CAS  Google Scholar 

  4. Lessard, J. et al. An essential switch in subunit composition of a chromatin remodeling complex during neural development. Neuron 55, 201–215 (2007).

    Article  CAS  Google Scholar 

  5. Lickert, H. et al. Baf60c is essential for function of BAF chromatin remodelling complexes in heart development. Nature 432, 107–112 (2004).

    Article  CAS  Google Scholar 

  6. Priam, P. et al. SMARCD2 subunit of SWI/SNF chromatin-remodeling complexes mediates granulopoiesis through a CEBPε dependent mechanism. Nat. Genet. 49, 753–764 (2017).

    Article  CAS  Google Scholar 

  7. Witzel, M. et al. Chromatin-remodeling factor SMARCD2 regulates transcriptional networks controlling differentiation of neutrophil granulocytes. Nat. Genet. 49, 742–752 (2017).

    Article  CAS  Google Scholar 

  8. Staahl, B. T. et al. Kinetic analysis of npBAF to nBAF switching reveals exchange of SS18 with CREST and integration with neural developmental pathways. J. Neurosci. 33, 10348–10361 (2013).

    Article  CAS  Google Scholar 

  9. Yoo, A. S., Staahl, B. T., Chen, L. & Crabtree, G. R. MicroRNA-mediated switching of chromatin-remodelling complexes in neural development. Nature 460, 642–646 (2009).

    Article  CAS  Google Scholar 

  10. Yoo, A. S. et al. MicroRNA-mediated conversion of human fibroblasts to neurons. Nature 476, 228–231 (2011).

    Article  CAS  Google Scholar 

  11. Pedersen, T. A., Kowenz-Leutz, E., Leutz, A. & Nerlov, C. Cooperation between C/EBPα TBP/TFIIB and SWI/SNF recruiting domains is required for adipocyte differentiation. Genes Dev. 15, 3208–3216 (2001).

    Article  CAS  Google Scholar 

  12. Pan, J. et al. Interrogation of mammalian protein complex structure, function, and membership using genome-scale fitness screens. Cell Syst. 6, 555–568 (2018).

    Article  CAS  Google Scholar 

  13. Alpsoy, A. & Dykhuizen, E. C. Glioma tumor suppressor candidate region gene 1 (GLTSCR1) and its paralog GLTSCR1-like form SWI/SNF chromatin remodeling subcomplexes. J. Biol. Chem. 293, 3892–3903 (2018).

    Article  CAS  Google Scholar 

  14. Wang, W. et al. Diversity and specialization of mammalian SWI/SNF complexes. Genes Dev. 10, 2117–2130 (1996).

    Article  CAS  Google Scholar 

  15. Kaeser, M. D., Aslanian, A., Dong, M. Q., Yates, J. R. 3rd & Emerson, B. M. BRD7, a novel PBAF-specific SWI/SNF subunit, is required for target gene activation and repression in embryonic stem cells. J. Biol. Chem. 283, 32254–32263 (2008).

    Article  CAS  Google Scholar 

  16. Kadoch, C. et al. Proteomic and bioinformatic analysis of mammalian SWI/SNF complexes identifies extensive roles in human malignancy. Nat. Genet. 45, 592–601 (2013).

    Article  CAS  Google Scholar 

  17. Shain, A. H. & Pollack, J. R. The spectrum of SWI/SNF mutations, ubiquitous in human cancers. PLoS ONE 8, e55119 (2013).

    Article  Google Scholar 

  18. Biegel, J. A. et al. Germ-line and acquired mutations of INI1 in atypical teratoid and rhabdoid tumors. Cancer Res. 59, 74–79 (1999).

    CAS  PubMed  Google Scholar 

  19. Eaton, K. W., Tooke, L. S., Wainwright, L. M., Judkins, A. R. & Biegel, J. A. Spectrum of SMARCB1/INI1 mutations in familial and sporadic rhabdoid tumors. Pediatr. Blood Cancer 56, 7–15 (2011).

    Article  Google Scholar 

  20. Versteege, I. et al. Truncating mutations of hSNF5/INI1 in aggressive paediatric cancer. Nature 394, 203–206 (1998).

    Article  CAS  Google Scholar 

  21. Jones, S. et al. Frequent mutations of chromatin remodeling gene ARID1A in ovarian clear cell carcinoma. Science 330, 228–231 (2010).

    Article  CAS  Google Scholar 

  22. Varela, I. et al. Exome sequencing identifies frequent mutation of the SWI/SNF complex gene PBRM1 in renal carcinoma. Nature 469, 539–542 (2011).

    Article  CAS  Google Scholar 

  23. McBride, M. J. et al. The SS18–SSX fusion oncoprotein hijacks BAF complex targeting and function to drive synovial sarcoma. Cancer Cell 33, 1128–1141 (2018).

    Article  CAS  Google Scholar 

  24. Helming, K. C. et al. ARID1B is a specific vulnerability in ARID1A-mutant cancers. Nat. Med. 20, 251–254 (2014).

    Article  CAS  Google Scholar 

  25. Hoffman, G. R. et al. Functional epigenetics approach identifies BRM/SMARCA2 as a critical synthetic lethal target in BRG1-deficient cancers. Proc. Natl Acad. Sci. USA 111, 3128–3133 (2014).

    Article  CAS  Google Scholar 

  26. Meyers, R. M. et al. Computational correction of copy number effect improves specificity of CRISPR–Cas9 essentiality screens in cancer cells. Nat. Genet. 49, 1779–1784 (2017).

    Article  CAS  Google Scholar 

  27. Tsherniak, A. et al. Defining a cancer dependency map. Cell 170, 564–576 (2017).

    Article  CAS  Google Scholar 

  28. Wang, T. et al. Gene essentiality profiling reveals gene networks and synthetic lethal interactions with oncogenic Ras. Cell 168, 890–903 (2017).

    Article  CAS  Google Scholar 

  29. McDonald, E. R. 3rd et al. Project DRIVE: a compendium of cancer dependencies and synthetic lethal relationships uncovered by large-scale, deep RNAi screening. Cell 170, 577–592 (2017).

    Article  CAS  Google Scholar 

  30. Cowley, G. S. et al. Parallel genome-scale loss of function screens in 216 cancer cell lines for the identification of context-specific genetic dependencies. Sci. Data 1, 140035 (2014).

    Article  CAS  Google Scholar 

  31. Bell, A. C. & Felsenfeld, G. Methylation of a CTCF-dependent boundary controls imprinted expression of the Igf2 gene. Nature 405, 482–485 (2000).

    Article  CAS  Google Scholar 

  32. Bell, A. C., West, A. G. & Felsenfeld, G. The protein CTCF is required for the enhancer blocking activity of vertebrate insulators. Cell 98, 387–396 (1999).

    Article  CAS  Google Scholar 

  33. Hark, A. T. et al. CTCF mediates methylation-sensitive enhancer-blocking activity at the H19/Igf2 locus. Nature 405, 486–489 (2000).

    Article  CAS  Google Scholar 

  34. Kanduri, C. et al. Functional association of CTCF with the insulator upstream of the H19 gene is parent of origin-specific and methylation-sensitive. Curr. Biol. 10, 853–856 (2000).

    Article  CAS  Google Scholar 

  35. Alver, B. H. et al. The SWI/SNF chromatin remodelling complex is required for maintenance of lineage specific enhancers. Nat. Commun. 8, 14648 (2017).

    Article  Google Scholar 

  36. Mathur, R. et al. ARID1A loss impairs enhancer-mediated gene regulation and drives colon cancer in mice. Nat. Genet. 49, 296–302 (2017).

    Article  CAS  Google Scholar 

  37. Wang, X. et al. SMARCB1-mediated SWI/SNF complex function is essential for enhancer regulation. Nat. Genet. 49, 289–295 (2017).

    Article  CAS  Google Scholar 

  38. Nakayama, R. T. et al. SMARCB1 is required for widespread BAF complex-mediated activation of enhancers and bivalent promoters. Nat. Genet. 49, 1613–1623 (2017).

    Article  CAS  Google Scholar 

  39. Kadoch, C. & Crabtree, G. R. Reversible disruption of mSWI/SNF (BAF) complexes by the SS18–SSX oncogenic fusion in synovial sarcoma. Cell 153, 71–85 (2013).

    Article  CAS  Google Scholar 

  40. Clark, J. et al. Identification of novel genes, SYT and SSX, involved in the t(X;18)(p11.2; q11.2) translocation found in human synovial sarcoma. Nat. Genet. 7, 502–508 (1994).

    Article  CAS  Google Scholar 

  41. Hohmann, A. F. et al. Sensitivity and engineered resistance of myeloid leukemia cells to BRD9 inhibition. Nat. Chem. Biol. 12, 672–679 (2016).

    Article  CAS  Google Scholar 

  42. Martin, L. J. et al. Structure-based design of an in vivo active selective BRD9 inhibitor. J. Med. Chem. 59, 4462–4475 (2016).

    Article  CAS  Google Scholar 

  43. Remillard, D. et al. Degradation of the BAF complex factor BRD9 by heterobifunctional ligands. Angew. Chem. Int. Ed. 56, 5738–5743 (2017).

    Article  CAS  Google Scholar 

  44. Wang, X. et al. Oncogenesis caused by loss of the SNF5 tumor suppressor is dependent on activity of BRG1, the ATPase of the SWI/SNF chromatin remodeling complex. Cancer Res. 69, 8094–8101 (2009).

    Article  CAS  Google Scholar 

  45. Chun, H. J. et al. Genome-wide profiles of extra-cranial malignant rhabdoid tumors reveal heterogeneity and dysregulated developmental pathways. Cancer Cell 29, 394–406 (2016).

    Article  CAS  Google Scholar 

  46. Theodoulou, N. H. et al. Discovery of I-BRD9, a selective cell active chemical probe for bromodomain containing protein 9 inhibition. J. Med. Chem. 59, 1425–1439 (2016).

    Article  CAS  Google Scholar 

  47. Coatham, M. et al. Concurrent ARID1A and ARID1B inactivation in endometrial and ovarian dedifferentiated carcinomas. Mod. Pathol. 29, 1586–1593 (2016).

    Article  CAS  Google Scholar 

  48. Tauziede-Espariat, A. et al. Loss of SMARCE1 expression is a specific diagnostic marker of clear cell meningioma: a comprehensive immunophenotypical and molecular analysis. Brain Pathol. 2, 466–474 (2017).

    Article  Google Scholar 

  49. Naka, N. et al. Synovial sarcoma is a stem cell malignancy. Stem Cells 28, 1119–1131 (2010).

    CAS  Google Scholar 

  50. Munoz, D. M. et al. CRISPR screens provide a comprehensive assessment of cancer vulnerabilities but generate false-positive hits for highly amplified genomic regions. Cancer Discov. 6, 900–913 (2016).

    Article  CAS  Google Scholar 

  51. Mashtalir, N. et al. Autodeubiquitination protects the tumor suppressor BAP1 from cytoplasmic sequestration mediated by the atypical ubiquitin ligase UBE2O. Mol. Cell 54, 392–406 (2014).

    Article  CAS  Google Scholar 

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

    Article  CAS  Google Scholar 

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

    Article  Google Scholar 

  54. Zhu, L. J. et al. ChIPpeakAnno: a Bioconductor package to annotate ChIP-seq and ChIP-chip data. BMC Bioinform. 11, 237 (2010).

    Article  Google Scholar 

  55. Dale, R. K., Pedersen, B. S. & Quinlan, A. R. Pybedtools: a flexible Python library for manipulating genomic datasets and annotations. Bioinformatics 27, 3423–3424 (2011).

    Article  CAS  Google Scholar 

  56. Liao, Y., Smyth, G. K. & Shi, W. The Subread aligner: fast, accurate and scalable read mapping by seed-and-vote. Nucleic Acids Res. 41, e108 (2013).

    Article  Google Scholar 

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

    Article  CAS  Google Scholar 

  58. Whyte, W. A. et al. Master transcription factors and mediator establish super-enhancers at key cell identity genes. Cell 153, 307–319 (2013).

    Article  CAS  Google Scholar 

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

    Article  CAS  Google Scholar 

  60. McLean, C. Y. et al. GREAT improves functional interpretation of cis-regulatory regions. Nat. Biotechnol. 28, 495–501 (2010).

    Article  CAS  Google Scholar 

  61. Machanick, P. & Bailey, T. L. MEME-ChIP: motif analysis of large DNA datasets. Bioinformatics 27, 1696–1697 (2011).

    Article  CAS  Google Scholar 

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

    Article  Google Scholar 

  63. Feng, J. et al. GFOLD: a generalized fold change for ranking differentially expressed genes from RNA-seq data. Bioinformatics 28, 2782–2788 (2012).

    Article  CAS  Google Scholar 

  64. Ramirez, F., Dundar, F., Diehl, S., Gruning, B. A. & Manke, T. deepTools: a flexible platform for exploring deep-sequencing data. Nucleic Acids Res. 42, W187–W191 (2014).

    Article  CAS  Google Scholar 

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

    Article  CAS  Google Scholar 

  66. Tripathi, S. et al. Meta- and orthogonal integration of Influenza "OMICs" data defines a role for UBR4 in virus budding. Cell Host Microbe 18, 723–735 (2015).

    Article  CAS  Google Scholar 

Download references

Acknowledgements

We are grateful to members of the Kadoch Laboratory for thoughtful discussions and insights relating to this study. We thank F. Winston, R. Kingston and Y. Shi for advice and critical review of the results. We thank A. Kuo and G.R. Crabtree for the homemade BAF155 (SMARCC1) polyclonal antibody used in these studies. We thank Z. Herbert, M. Berkeley and members of the Molecular Biology Core Facility for library preparation and sequencing. We also thank W. Wang, Z. Zhang and B. Li for conducting the CRISPR screening and cellular proliferation experiments. This work was supported in part by the NIH DP2 New Innovator Award no. 1DP2CA195762–01, American Cancer Society Research Scholar Award no. RSG-14-051-01-DMC and the Pew–Stewart Scholars in Cancer Research Grant awarded to C.K. B.C.M. holds the Albert J. Ryan Fellowship granted by the Division of Medical Sciences (Harvard Medical School). In addition, this work was supported in part by NIH Grant no. 5 T32 GM095450-04 (M.J.M.), a Harvard University Graduate School of Arts and Sciences Fellowship (M.J.M.), a Ford Foundation Fellowship (A.M.V), Howard Hughes Medical Institute Gilliam Fellows Program (A.M.V.) and the National Science Foundation Graduate Research Fellowship Program (J.P.). The content is solely the responsibility of the authors and does not necessarily represent the official views of the National Institute of General Medical Sciences or the National Institutes of Health.

Author information

Authors and Affiliations

Authors

Contributions

B.C.M., A.R.D., S.H.C., and C.K. conceived and designed the study. B.C.M. designed and performed most experiments, A.R.D. performed all bioinformatic analyses and statistical calculations. S.H.C. designed and performed GLTSCR1/1L biochemistry and contributed to ChIP-seq interpretation and analysis, Z.M.M. performed GLTSCR1/1L biochemistry, N.M. and J.P. were involved in the design and execution of experiments pertaining to ncBAF biochemistry, M.J.M, and A.M.V. were involved in the design and execution of synovial sarcoma and MRT experiments, and J.P. contributed to the analysis and interpretation of large-scale dependency data. D.I.R. synthesized dBRD9 and assisted in experimental design using dBRD9, and H.J.Z. and N.F. assisted with conducting GLTSCR1/1L biochemistry experiments. H.M.C., Q.Z. and M.B. directed the CRISPR tiling experiments and L.M.M.S. performed bioinformatic analyses of these datasets. C.A.L contributed important insights and aided in data analysis and interpretation. N.S.G. and J.E.B. supervised the development of the dBRD9 small molecule. B.C.M., A.R.D., S.H.C. and C.K. wrote the manuscript.

Corresponding author

Correspondence to Cigall Kadoch.

Ethics declarations

Competing interests

C.K. is a scientific founder, fiduciary Board of Directors member, Scientific Advisory Board (SAB) member, shareholder and consultant for Foghorn Therapeutics, Inc. (Cambridge, MA). H.M.C., Q.Z, M.B. and L.M.M.S are employees and shareholders of Foghorn Therapeutics, Inc. (Cambridge, MA). N.S.G. is a scientific founder, SAB member and equity holder in Gatekeeper, Syros, Petra, Soltego and C4 Therapeutics. J.E.B is now an executive and shareholder of Novartis AG. He is a founder and former shareholder of Tensha Therapeutics (a bromodomain company, now Roche) and Syros (a chromatin biotech). The other 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.

Integrated supplementary information

Supplementary Figure 1 Mammalian SWI/SNF family complexes exist in three distinct classes.

(a). Heatmap representing correlations of fitness scores between mSWI/SNF complexes genes in genome-scale shRNA-based genetic perturbation screens (Project DRIVE, Novartis). (b). Table of total peptide counts (raw spectral counts) for each mass spectrometry experiment performed on mSWI/SNF complexes purified using HA-tagged baits. (c,d). Immunoprecipitation of endogenous GLTSCR1 (c) and GLTSCR1L (d) followed by immunoblot captures BRD9-specific mSWI/SNF subunits but not canonical BAF- or PBAF- specific subunits. Immunoprecipitations performed in n = 3 biologically independent experiments. See also Supplementary Figure 7j,k. (e). Immunoprecipitation of BRD9 followed by immunoblot for various subunits performed in NCIH-1437, BJ fibroblasts, IMR90, and ES-2 cell lines. Immunoprecipitations performed in n = 2 biologically independent experiments. See also Supplementary Figure 7l.

Supplementary Figure 2 mSWI/SNF complex subtypes differentially localize on chromatin.

(a). Schematic of subunits selected for ChIP-seq in EoL-1 cells: BRD9 and GLTSCR1 (ncBAF-specific), DPF2 (BAF-specific), BRD7 (PBAF-specific) and SMARCA4 and SMARCC1 (pan-mSWI/SNF) subunits. (b). Pearson correlation of read density between ChIP-seq experiments using two different BRD9 antibodies in EoL-1. ChIP-seq was performed in n = 2 independent samples. (c,d). Venn diagrams representing overlap between SMARCA4 and DPF2 (c) or BRD7 (d) ChIP-seq peaks in EoL-1. (e). Venn diagram of peaks for BRD7 (PBAF), BRD9 (ncBAF), and DPF2 (cBAF) in EoL-1. (f). Distance of each peak to the nearest TSS in indicated ChIP-seq experiments. (g). BAF, PBAF, and ncBAF complex ChIP-Seq read density distribution over the TSS and 2.5kb into the gene body in EoL-1. (h). Localization of CTCF and ncBAF, BAF, and PBAF complexes at the SH2B3 locus. CTCF-BRD9 overlap sites are shaded in gray. ChIP-seq was performed in n = 2 independent samples. (i). Distribution of CTCF, H3K27ac, H3K4me1, and H3K4me3 marks across all mSWI/SNF sites genome-wide in EoL-1, clustered into four groups. (j). ChIP-seq read density summary plots of DFP2-, BRD9-, and BRD7- bound mSWI/SNF complexes over active enhancers, active promoters, CTCF sites, and primed sites in EoL-1. (k). Example track depicting differential mSWI/SNF complex binding at the CMC1 locus. ChIP-seq was performed in n = 2 independent samples for mSWI/SNF subunits and n = 1 for histone marks. (l). Heatmap of CTCF, BRD9, H3K4me3, and H3K4me1 ChIP-seq occupancy over all CTCF sites in EoL-1, split into proximal and distal sites, and ranked by BRD9 density.

Supplementary Figure 3 Synovial sarcoma and malignant rhabdoid tumor cell lines are sensitive to ncBAF perturbation.

(a). Schematic for CRISPR-Cas9-based synthetic lethal screening. (b). CERES dependency scores for ncBAF subunits BRD9, GLTSCR1, and SMARCD1 across all soft tissue and bone cancers, ranked by BRD9 score. (c). Waterfall plots of ATARIS (Project DRIVE) scores across n = 387 cancer cell lines for indicated subunits; dashed line = −0.75 score. (d). BAF subunit perturbations in WT, SS, and MRT settings. (e). Heatmap of the z-score of CERES scores across all n = 408 cancer cell lines ranked by median z-score. (f). Immunoblot for ncBAF subunits in HEK-293T cells upon 250nM dBRD9 treatment or BRD9 KO (n = 2). See also Supplementary Figure 7m. (g). Immunoblot and proliferation on SYO-1 cells with either SS18-SSX1 (shSSX) or control (shCtrl) (n = 2 biologically independent experiments for each). Each data point represents mean ± SD from n = 3 biologically independent samples, p-value calculated by two-sided t-test on day 20. See also Supplementary Figure 7n, Supplementary Table 2. (h). Immunoblot and proliferation (n = 1 experiment) performed on SYO-1 cells treated with GLTSCR1 (shGLT1) or a non-targeting guide. Each data point is mean ± SD from n = 3 biologically independent samples, p-value calculated by two-sided t-test on day 7. See also Supplementary Table 2 (i, j). FACS-based cell cycle analysis (i) and Annexin V staining (j) for SYO-1 cells after 8 days of compound treatment (n = 1). (k). (Left) Immunoblot on G401 MRT cells treated with DMSO or dBRD9 (250 nM) (n = 2 biologically independent experiments); (Right) Proliferation experiments performed in G401, each data point represents mean ± SD from n = 3 biologically independent samples, p-value calculated by two-sided t-test on day 7. See also Supplementary Figure 7o, Supplementary Table 2. (l). Proliferation experiment in SMARCB1-intact ESX cells treated with DMSO or dBRD9 (250nM). Each data point represents mean ± SD from n = 3 biologically independent samples. (m, n). Colony formation assay (as in Fig. 3i) performed on HSSYII (m) and Aska (n) cells (n = 2 biologically independent experiments). (o, p, q). Colony formation performed on HCT-116 (n), Calu-6 (o), and RD rhabdomyosarcoma (p) cell lines (n = 2 biologically independent experiments). (r). SS18 and SMARCC1 IP/immunoblot in BRD9 KO HEK-293T cells (n = 2 biologically independent experiments). See also Supplementary Figure 7p.

Supplementary Figure 4 CRISPR tiling screening performed across mSWI/SNF subunit genes in SYO-1 cells reveals subunit- and subunit domain- specific dependencies.

(a). Box plot of dropout scores for guides targeting the SS18 gene or non-targeting control in CRISPR tiling screens in SYO-1. Box represents interquartile range (IQR), bar in center shows data median. Minima and maxima shown extend to from the box +/− 1.5*IQR with data falling outside of that range shown as points. (SS18: n = 166, Control: n = 200). (b). CRISPR tiling screening performed across the SS18 gene in SYO-1 cells. (c,d). CRISPR tiling screening performed across the SMARCB1 (c) and SMARCE1 (d) genes in SYO-1 cells. (e). Box plot of dropout scores for guides to SMARCD family paralogs or non-targeting control in CRISPR tiling screen in SYO-1. Box represents interquartile range (IQR), bar in center shows data median. Minima and maxima shown extend to from the box +/− 1.5*IQR with data falling outside of that range shown as points. (SMARCD1: n = 147, SMARCD2: n = 167, SMARCD3: n = 137, Control: n = 200). (f). CRISPR tiling screening performed across SMARCD1 gene indicating no specific domain dropout. (g). Box plot of dropout score for guides to SMARCC family paralogs or non-targeting control in CRISPR tiling screen in SYO-1. Box represents interquartile range (IQR), bar in center shows data median. Minima and maxima shown extend to from the box +/− 1.5*IQR with data falling outside of that range shown as points. (SMARCC1: n = 314, SMARCC2: n = 430, Control: n = 200). (h). CRISPR tiling screening performed across the SMARCC1 gene indicating no specific domain dropout. (i). Immunoprecipitation of mammalian GLTSCR1 full-length (GLTSCR1-FL) and GLTSCR1 N-terminal deletion (G1-Ndel) followed by immunoblot (n = 2 biologically independent experiments).

Supplementary Figure 5 BRD9 and SS18-SSX regulate distinct gene sets in synovial sarcoma.

(a). Enriched gene sets for gene groups 1, 2, and 3 from Fig. 5b. (b). Schematic depicting experimental conditions in CRL7250 fibroblast cells used in RNA-seq experiments. (c). GSEA performed on RNA-seq experiments from conditions outlined in Supplementary Figure 5b. (d). Example tracks at an SS18-SSX fusion-dependent site (left) and bar graph of gene expression by RNA-seq (right) in SYO-1 at the FLRT2 locus. n = 2 independent samples for each ChIP-seq experiment. Bar represents mean RPKM of n = 2 RNA replicates for each condition with RPKM for each sample plotted as a dot. (e). Example tracks at SS18-SSX fusion-independent sites (left) and bar graphs of gene expression by RNA-seq (right) in SYO-1 at the SLC7A5 and SRM loci. n = 2 independent samples for each ChIP-seq experiment. Bar represents mean RPKM of n = 2 RNA replicates for each condition with RPKM for each sample plotted as a dot. (f). Violin plot of CERES scores for genes that changed with a significance of p-adjusted<1e-3 after 6 days of dBRD9 treatment in MOLM-13 cells. p-adjusted values are Benjamini-Hochberg adjusted Wald p-values, p-value between sets of genes was calculated by two-sided t-test. Violin plot shows kernel density estimation with data quartiles represented as lines, the data median is shown as a dot.

Supplementary Figure 6 BRD9 maintains gene expression at retained, CTCF-marked promoter sites in BAF-perturbed settings of synovial sarcoma and malignant rhabdoid tumor.

(a). Hockey stick plot of TTC1240 H3K27ac signal with MRT-specific super enhancers as defined by Chun et al. marked in red. (b). Track showing BRD9 (DMSO), SMARCA4 (DMSO), SMARCA4 (250nM dBRD9), and H3K27ac (empty vector condition) occupancy at the LIF locus in TTC1240. n = 2 independent samples for each ChIP-seq experiment. (c). Boxplots of H3K27ac and BRD9 ChIP occupancy at the promoters of active genes (n = 1064 sig. changing genes, n = 11503 non_changing genes). n = 2 independent samples for each ChIP-seq experiment, p-value calculated using two-sided t-test. Box represents interquartile range (IQR), bar in center shows data median. Minima and maxima shown extend to from the box +/− 1.5*IQR. (d). GREAT analysis of GO Biological Process for genes near SMARCA4 sites lost upon dBRD9 treatment. (e). ChIP-Seq density heatmap of SMARCA4, BRD9, H3K4me3, H3K4me1, H3K27ac, and CTCF over SMARCA4 proximal (<2 kb to TSS) and distal sites (>2 kb to TSS) in TTC1240 Empty sorted by BRD9 density. (f). ChIP-Seq density heatmap of SS18, BRD9, H3K4me3, SYO-1 CTCF, and EOL-1 CTCF over shScr BRD9 sites in Aska, ranked by difference in SS18 density between shScr and shSSX conditions. (g). BRD9 ChIP-seq density over CTCF sites ordered by BRD9 density in shCtrl condition in SYO-1 cells. (h). BRD9 ChIP-seq density before and after SMARCB1 reintroduction in TTC1240 cells over CTCF sites.

Supplementary Figure 7 All raw and unprocessed immunoblots.

(a). Western blots related to Fig. 1d. (b). Western blots related to Fig. 1e. (c). Western blots related to Fig. 3d. (d). Western blots related to Fig. 3e. (e). Western blots related to Fig. 3h. (f). Western blots related to Fig. 4g. (g). Western blots related to Fig. 4h. (h). Western blots related to Fig. 5a. (i). Western blots related to Fig. 5d. (j). Western blots related to Supplementary Figure 1c. (k). Western blots related to Supplementary Figure 1d. (l). Western blots related to Supplementary Figure 1e. (m). Western blots related to Supplementary Figure 3f. (n). Western blots related to Supplementary Figure 3g. (o). Western blots related to Supplementary Figure 3k. (p). Western blots related to Supplementary Figure 3r. (q). Western blots related to Supplementary Figure 4i.

Supplementary information

Supplementary Information

Supplementary Figures 1–7 and Supplementary Table legends.

Reporting Summary

Supplementary Table 1

Proteomics for BRD9, DPF2, BRD7 purifications in HEK-293T cells (raw peptide counts).

Supplementary Table 2

Statistics source data for cell line proliferation experiments.

Supplementary Table 3

CRISPR–Cas9 tiled sgRNA guide information.

Supplementary Table 4

mSWI/SNF gene CRISPR–Cas9 dropout scores in SYO-1 cells

Supplementary Table 5

Sequencing statistics and quality control metrics for genomic data.

Supplementary Table 6

All primer sets used in this study.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Michel, B.C., D’Avino, A.R., Cassel, S.H. et al. A non-canonical SWI/SNF complex is a synthetic lethal target in cancers driven by BAF complex perturbation. Nat Cell Biol 20, 1410–1420 (2018). https://doi.org/10.1038/s41556-018-0221-1

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41556-018-0221-1

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