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:

Dominant-negative SMARCA4 mutants alter the accessibility landscape of tissue-unrestricted enhancers

Abstract

Mutation of SMARCA4 (BRG1), the ATPase of BAF (mSWI/SNF) and PBAF complexes, contributes to a range of malignancies and neurologic disorders. Unfortunately, the effects of SMARCA4 missense mutations have remained uncertain. Here we show that SMARCA4 cancer missense mutations target conserved ATPase surfaces and disrupt the mechanochemical cycle of remodeling. We find that heterozygous expression of mutants alters the open chromatin landscape at thousands of sites across the genome. Loss of DNA accessibility does not directly overlap with Polycomb accumulation, but is enriched in ‘A compartments’ at active enhancers, which lose H3K27ac but not H3K4me1. Affected positions include hundreds of sites identified as superenhancers in many tissues. Dominant-negative mutation induces pro-oncogenic expression changes, including increased expression of Myc and its target genes. Together, our data suggest that disruption of enhancer accessibility represents a key source of altered function in disorders with SMARCA4 mutations in a wide variety of tissues.

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: Heterozygous SMARCA4 mutations contribute to many cancers.
Fig. 2: SMARCA4 cancer mutations affect functional surfaces and induce distinct dynamic defects.
Fig. 3: Convergent effects of SMARCA4 cancer mutants on the DNA-accessibility landscape.
Fig. 4: Accessibility losses and PRC1 changes do not directly overlap.
Fig. 5: Heterozygous SMARCA4 mutation induces loss of accessibility and H3K27ac at active enhancers and superenhancers from many tissues.
Fig. 6: Chromatin organization influences dominant-negative effects of SMARCA4 mutants on the open chromatin landscape.
Fig. 7: SMARCA4 ATPase mutants can induce pro-oncogenic gene expression changes.

Similar content being viewed by others

References

  1. Hodges, C., Kirkland, J. G. & Crabtree, G. R. The many roles of BAF (mSWI/SNF) and PBAF complexes in cancer. Cold Spring Harb. Perspect. Med. 6, a026930 (2016).

    Google Scholar 

  2. Son, E. Y. & Crabtree, G. R. The role of BAF (mSWI/SNF) complexes in mammalian neural development. Am. J. Med. Genet. C. Semin. Med. Genet. 166, 333–349 (2014).

    Article  CAS  Google Scholar 

  3. 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  PubMed  PubMed Central  Google Scholar 

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

    Article  PubMed  PubMed Central  Google Scholar 

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

    Article  CAS  PubMed  Google Scholar 

  6. 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 

  7. 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  PubMed  PubMed Central  Google Scholar 

  8. Davoli, T. et al. Cumulative haploinsufficiency and triplosensitivity drive aneuploidy patterns and shape the cancer genome. Cell 155, 948–962 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  9. Wilson, B. G. & Roberts, C. W. SWI/SNF nucleosome remodellers and cancer. Nat. Rev. Cancer 11, 481–492 (2011).

    Article  CAS  PubMed  Google Scholar 

  10. Wei, D. et al. SNF5/INI1 deficiency redefines chromatin remodeling complex composition during tumor development. Mol. Cancer Res. 12, 1574–1585 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

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

    Article  CAS  PubMed  Google Scholar 

  12. Oruetxebarria, I. et al. P16INK4a is required for hSNF5 chromatin remodeler-induced cellular senescence in malignant rhabdoid tumor cells. J. Biol. Chem. 279, 3807–3816 (2004).

    Article  CAS  PubMed  Google Scholar 

  13. Kia, S. K., Gorski, M. M., Giannakopoulos, S. & Verrijzer, C. P. SWI/SNF mediates polycomb eviction and epigenetic reprogramming of the INK4b-ARF-INK4a locus. Mol. Cell. Biol. 28, 3457–3464 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  14. Wilson, B. G. et al. Epigenetic antagonism between polycomb and SWI/SNF complexes during oncogenic transformation. Cancer Cell 18, 316–328 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  15. Lee, R. S. et al. A remarkably simple genome underlies highly malignant pediatric rhabdoid cancers. J. Clin. Invest. 122, 2983–2988 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  16. Lawrence, M. S. et al. Mutational heterogeneity in cancer and the search for new cancer-associated genes. Nature 499, 214–218 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  17. de Bruijn, D. R. et al. The synovial-sarcoma-associated SS18-SSX2 fusion protein induces epigenetic gene (de)regulation. Cancer Res. 66, 9474–9482 (2006).

    Article  PubMed  Google Scholar 

  18. Lubieniecka, J. M. et al. Histone deacetylase inhibitors reverse SS18-SSX-mediated polycomb silencing of the tumor suppressor early growth response 1 in synovial sarcoma. Cancer Res. 68, 4303–4310 (2008).

    Article  CAS  PubMed  Google Scholar 

  19. Zhang, Z. K. et al. Cell cycle arrest and repression of cyclin D1 transcription by INI1/hSNF5. Mol. Cell. Biol. 22, 5975–5988 (2002).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  20. 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  PubMed  PubMed Central  Google Scholar 

  21. 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  PubMed  Google Scholar 

  22. 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  PubMed  PubMed Central  Google Scholar 

  23. Miller, E. L. et al. TOP2 synergizes with BAF chromatin remodeling for both resolution and formation of facultative heterochromatin. Nat. Struct. Mol. Biol. 24, 344–352 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  24. Fillmore, C. M. et al. EZH2 inhibition sensitizes BRG1 and EGFR mutant lung tumours to TopoII inhibitors. Nature 520, 239–242 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  25. Imielinski, M. et al. Mapping the hallmarks of lung adenocarcinoma with massively parallel sequencing. Cell 150, 1107–1120 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  26. The Cancer Genome Atlas Research Network. Comprehensive molecular profiling of lung adenocarcinoma. Nature 511, 543–550 (2014).

    Article  Google Scholar 

  27. Witkowski, L. et al. Germline and somatic SMARCA4 mutations characterize small cell carcinoma of the ovary, hypercalcemic type. Nat. Genet. 46, 438–443 (2014).

    Article  CAS  PubMed  Google Scholar 

  28. Jelinic, P. et al. Recurrent SMARCA4 mutations in small cell carcinoma of the ovary. Nat. Genet. 46, 424–426 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  29. Karnezis, A. N. et al. Dual loss of the SWI/SNF complex ATPases SMARCA4/BRG1 and SMARCA2/BRM is highly sensitive and specific for small cell carcinoma of the ovary, hypercalcaemic type. J. Pathol. 238, 389–400 (2016).

    Article  CAS  PubMed  Google Scholar 

  30. Hodis, E. et al. A landscape of driver mutations in melanoma. Cell 150, 251–263 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  31. Le Loarer, F. et al. SMARCA4 inactivation defines a group of undifferentiated thoracic malignancies transcriptionally related to BAF-deficient sarcomas. Nat. Genet. 47, 1200–1205 (2015).

    Article  PubMed  Google Scholar 

  32. Love, C. et al. The genetic landscape of mutations in Burkitt lymphoma. Nat. Genet. 44, 1321–1325 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  33. Lunning, M. A. & Green, M. R. Mutation of chromatin modifiers; an emerging hallmark of germinal center B-cell lymphomas. Blood Cancer J 5, e361 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  34. Stanton, B. Z. et al. Smarca4 ATPase mutations disrupt direct eviction of PRC1 from chromatin. Nat. Genet. 49, 282–288 (2017).

    Article  CAS  PubMed  Google Scholar 

  35. The Cancer Genome Atlas Research Network et al. The cancer genome atlas pan-cancer analysis project. Nat. Genet. 45, 1113–1120 (2013).

    Article  PubMed Central  Google Scholar 

  36. Zehir, A. et al. Mutational landscape of metastatic cancer revealed from prospective clinical sequencing of 10,000 patients. Nat. Med. 23, 703–713 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  37. Liu, X., Li, M., Xia, X., Li, X. & Chen, Z. Mechanism of chromatin remodelling revealed by the Snf2-nucleosome structure. Nature 544, 440–445 (2017).

    Article  CAS  PubMed  Google Scholar 

  38. Xia, X., Liu, X., Li, T., Fang, X. & Chen, Z. Structure of chromatin remodeler Swi2/Snf2 in the resting state. Nat. Struct. Mol. Biol. 23, 722–729 (2016).

    Article  CAS  PubMed  Google Scholar 

  39. Hauk, G., McKnight, J. N., Nodelman, I. M. & Bowman, G. D. The chromodomains of the Chd1 chromatin remodeler regulate DNA access to the ATPase motor. Mol. Cell 39, 711–723 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  40. Gao, J. et al. Integrative analysis of complex cancer genomics and clinical profiles using the cBioPortal. Sci. Signal. 6, pl1 (2013).

    Article  PubMed  PubMed Central  Google Scholar 

  41. Laurent, B. C., Treich, I. & Carlson, M. The yeast SNF2/SWI2 protein has DNA-stimulated ATPase activity required for transcriptional activation. Genes Dev. 7, 583–591 (1993).

    Article  CAS  PubMed  Google Scholar 

  42. Johnson, T. A., Elbi, C., Parekh, B. S., Hager, G. L. & John, S. Chromatin remodeling complexes interact dynamically with a glucocorticoid receptor-regulated promoter. Mol. Biol. Cell 19, 3308–3322 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  43. Gurskaya, N. G. et al. Engineering of a monomeric green-to-red photoactivatable fluorescent protein induced by blue light. Nat. Biotechnol. 24, 461–465 (2006).

    Article  CAS  PubMed  Google Scholar 

  44. Sif, S., Stukenberg, P. T., Kirschner, M. W. & Kingston, R. E. Mitotic inactivation of a human SWI/SNF chromatin remodeling complex. Genes Dev. 12, 2842–2851 (1998).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  45. Ho, L. et al. esBAF facilitates pluripotency by conditioning the genome for LIF/STAT3 signalling and by regulating polycomb function. Nat. Cell Biol. 13, 903–913 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  46. Buenrostro, J. D., Giresi, P. G., Zaba, L. C., Chang, H. Y. & Greenleaf, W. J. Transposition of native chromatin for fast and sensitive epigenomic profiling of open chromatin, DNA-binding proteins and nucleosome position. Nat. Methods 10, 1213–1218 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  47. Whyte, W. A. et al. Enhancer decommissioning by LSD1 during embryonic stem cell differentiation. Nature 482, 221–225 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  48. Francis, N. J., Kingston, R. E. & Woodcock, C. L. Chromatin compaction by a polycomb group protein complex. Science 306, 1574–1577 (2004).

    Article  CAS  PubMed  Google Scholar 

  49. Endoh, M. et al. Histone H2A mono-ubiquitination is a crucial step to mediate PRC1-dependent repression of developmental genes to maintain ES cell identity. PLoS Genet. 8, e1002774 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  50. 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  PubMed  PubMed Central  Google Scholar 

  51. Ha, M. & Hong, S. Gene-regulatory interactions in embryonic stem cells represent cell-type specific gene regulatory programs. Nucleic Acids Res. 45, 10428–10435 (2017).

    Article  PubMed  PubMed Central  Google Scholar 

  52. Khan, A. & Zhang, X. dbSUPER: a database of super-enhancers in mouse and human genome. Nucleic Acids Res. 44 D1, D164–D171 (2016).

    Article  Google Scholar 

  53. Bultman, S. et al. A Brg1 null mutation in the mouse reveals functional differences among mammalian SWI/SNF complexes. Mol. Cell 6, 1287–1295 (2000).

    Article  CAS  PubMed  Google Scholar 

  54. Lieberman-Aiden, E. et al. Comprehensive mapping of long-range interactions reveals folding principles of the human genome. Science 326, 289–293 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  55. Dekker, J., Marti-Renom, M. A. & Mirny, L. A. Exploring the three-dimensional organization of genomes: interpreting chromatin interaction data. Nat. Rev. Genet. 14, 390–403 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  56. Brookes, E. et al. Polycomb associates genome-wide with a specific RNA polymerase II variant, and regulates metabolic genes in ESCs. Cell Stem Cell 10, 157–170 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  57. Araya, C. L. et al. Identification of significantly mutated regions across cancer types highlights a rich landscape of functional molecular alterations. Nat. Genet. 48, 117–125 (2016).

    Article  CAS  PubMed  Google Scholar 

  58. Soufi, A., Donahue, G. & Zaret, K. S. Facilitators and impediments of the pluripotency reprogramming factors’ initial engagement with the genome. Cell 151, 994–1004 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  59. Rahl, P. B. et al. c-Myc regulates transcriptional pause release. Cell 141, 432–445 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  60. Lin, C. Y. et al. Transcriptional amplification in tumor cells with elevated c-Myc. Cell 151, 56–67 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  61. 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  PubMed  PubMed Central  Google Scholar 

  62. Langmead, B., Trapnell, C., Pop, M. & Salzberg, S. L. Ultrafast and memory-efficient alignment of short DNA sequences to the human genome. Genome Biol. 10, R25 (2009).

    Article  PubMed  PubMed Central  Google Scholar 

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

    Article  PubMed  PubMed Central  Google Scholar 

  64. 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).

    Article  PubMed  PubMed Central  Google Scholar 

  65. Pohl, A. & Beato, M. bwtool: a tool for bigWig files. Bioinformatics 30, 1618–1619 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

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

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  67. Jin, W. et al. Genome-wide detection of DNase I hypersensitive sites in single cells and FFPE tissue samples. Nature 528, 142–146 (2015).

    CAS  PubMed  PubMed Central  Google Scholar 

  68. Barski, A. et al. High-resolution profiling of histone methylations in the human genome. Cell 129, 823–837 (2007).

    Article  CAS  PubMed  Google Scholar 

  69. Kidder, B. L. & Zhao, K. Efficient library preparation for next-generation sequencing analysis of genome-wide epigenetic and transcriptional landscapes in embryonic stem cells. Methods Mol. Biol. 1150, 3–20 (2014).

    Article  CAS  PubMed  Google Scholar 

  70. Hahne, F. & Ivanek, R. Visualizing genomic data using Gviz and Bioconductor. Methods Mol. Biol. 1418, 335–351 (2016).

    Article  PubMed  Google Scholar 

  71. 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  PubMed  Google Scholar 

  72. 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).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  73. Tibshirani, R. Regression shrinkage and selection via the lasso. J. R. Stat. Soc. B 58, 267–288 (1996).

    Google Scholar 

  74. Friedman, J., Hastie, T. & Tibshirani, R. Regularization paths for generalized linear models via coordinate descent. J. Stat. Softw. 33, 1–22 (2010).

    Article  PubMed  PubMed Central  Google Scholar 

  75. Krieger, E. & Vriend, G. YASARA View - molecular graphics for all devices - from smartphones to workstations. Bioinformatics 30, 2981–2982 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  76. Hooft, R. W., Vriend, G., Sander, C. & Abola, E. E. Errors in protein structures. Nature 381, 272 (1996).

    Article  CAS  PubMed  Google Scholar 

  77. Hooft, R. W., Sander, C., Scharf, M. & Vriend, G. The PDBFINDER database: a summary of PDB, DSSP and HSSP information with added value. Comput. Appl. Biosci. 12, 525–529 (1996).

    CAS  PubMed  Google Scholar 

  78. Qiu, J. & Elber, R. SSALN: an alignment algorithm using structure-dependent substitution matrices and gap penalties learned from structurally aligned protein pairs. Proteins 62, 881–891 (2006).

    Article  CAS  PubMed  Google Scholar 

  79. Dixon, J. R. et al. Topological domains in mammalian genomes identified by analysis of chromatin interactions. Nature 485, 376–380 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

Download references

Acknowledgements

We thank C. Weber, E. Chory, K. Cui, G. Hu, and E. Son for helpful discussions and dedicate this manuscript to the lasting memory of Joseph P. Calarco. We gratefully acknowledge the assistance of the DNA Sequencing and Genomics Core facility of NHLBI, the Stanford Cell Sciences Imaging Facility (S10OD01227601), and the Stanford BioX3 cluster (S10RR02664701). Ring1a −/− ;Ring1b fl/fl mESCs were a generous gift from H. Koseki (RIKEN, Japan). This work was supported by the Simons Foundation Autism Research Initiative (G.R.C.), NIH grants R37NS046789 (G.R.C.), R01CA163915 (G.R.C.), T32CA009151 (J.G.K.), R00CA187565 (H.C.H.), the Division of Intramural Research of the NHLBI/NIH (K.Z.)., the Czech Science Foundation grant 16-06357S (V.V.), the Cancer Prevention & Research Institute of Texas grant RR170036 (H.C.H.), and the Howard Hughes Medical Institute (G.R.C.).

Author information

Authors and Affiliations

Authors

Contributions

H.C.H. and B.Z.S. conceived of and performed experiments and wrote the paper. K.C., H.C.H., and V.V. developed the homology model. H.C.H, C.-Y.C., and E.L.M. performed analyses. B.Z.S., H.C.H., C.-Y.C., J.G.K., and W.L.K. prepared materials. K.Z. and G.R.C. designed experiments and wrote the paper.

Corresponding authors

Correspondence to Keji Zhao or Gerald R. Crabtree.

Ethics declarations

Competing interests

The authors declare no competing financial 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 The SMARCA4 ATP cleft and DNA groove are frequently mutated in cancer.

(a) Surface rendering of the SMARCA4 homology model viewed from the DNA-binding surface. Details of homology model are described in the Methods section. Surface residues are colored based on mutation frequency of deleterious missense mutations at the time of analysis in cBioPortal, for residue positions 758–1335. DNA is shown in yellow. (b) Rotation of the view in (a) 90° around the x-axis to show the ATP-binding cleft. ATP is shown in blue. (c) Rotation of the view in (a) 180° around the y-axis. (d) Rotation of the view in (c) –90° around the x-axis reveals few other surface mutation hotspots.

Supplementary Figure 2 Construction of a Smarca4-Dendra2 fluorescent knockin mouse.

(a) A 15.5-kbp homology template was used to target the last exon of SMARCA4 to create a SMARCA4-Dendra2 fusion. Thymidine kinase from herpes simplex virus (HSV-tk) was used for negative selection with gancyclovir, and positive selection from neo was achieved by selecting for G418-resistant colonies. (b) Southern blot of resulting heterozygous knockin mESC colony using the probe shown in panel (a). (c) Western blotting of the heterozygous knockin mESC colony shows a new band with higher molecular weight that also stains positive using anti-Dendra2 antibodies (Evrogen). (d) Fluorescent knockin SMARCA4 constructs assemble with endogenous BAF and PBAF subunits to form natural complexes. (e) Confocal image of resulting fluorescent mESC colony. (f) Homozygous mice obtained through breeding are phenotypically normal. Strain has been deposited for cryopreservation at Jackson Lab (stock #27901). PCR primers for genotyping are provided in the Methods section.

Supplementary Figure 3 SMARCA4-Dendra2 is highly stable and undergoes dynamic changes in interaction with chromatin throughout the cell cycle.

(a) Following photoconversion with 405-nm light using a custom-made LED lamp, SMARCA4-Dendra2 irreversibly photoswitches from green to red. (b) Photoswitching is readily detectable using flow cytometry. (c) The persistence of red fluorescence provides a measure of the stability of the protein fusion. Red fluorescence persists with a half-life of 11.8 ± 1.2 h, comparable to the canonical 12-hour half-life of mESCs. This result indicates that fluorescent fusions of SMARCA4 are stable. (d) Live-cell imaging of mitosis. Mitotic cells are readily distinguishable from interphasic cells due to the altered distribution of fluorescence intensity. (e) Fluorescence recovery after photobleaching (FRAP) of an individual mESC colony. (f) FRAP recovery times of SMARCA4-Dendra2 fusions are faster in mitosis, when the complex is excluded from condensed mitotic chromatin (g) Statistical comparison of FRAP recovery times of SMARCA4-Dendra2 fluorescent fusions. Mitotic cells show significantly increased FRAP recovery times (p<0.002, KS test). (h) Custom LED array for bulk photoconversion of cells using 405-nm LEDs (DigiKey 492-1349-ND).

Supplementary Figure 4 ATPase-dead SMARCA4 shows altered interaction with chromatin during interphase.

(a) Fluorescence deconvolution image of fixed mESCs expressing lentivirally transduced SMARCA4-GFP. During mitosis, SMARCA4 is excluded from chromatin. (b) Live-cell imaging of SMARCA4-GFP and DNA (imaged using DRAQ5) shows comparable results. (c) ATPase-dead K785R mutant (K798R in some citations), a cancer mutation found in the ATP-binding cleft) shows a dramatic dynamic defect, consistent with failure to release immobile chromatin. (d) Wild-type SMARCA4-GFP and SMARCA4-Dendra2 have indistinguishable dynamics in FRAP assays, while K785R mutants show a significant increase in the FRAP recovery times during interphase (p<2.2e-16). (e) During mitosis, mutant and wild-type complexes show similar dynamics, consistent with a model whereby ATP-dependent dynamics predominate during interphase when the complex engages chromatin (Figure 2g).

Supplementary Figure 5 Genome-wide alterations of ATAC-seq following acute Smarca4 knockout and relationship with Ring1b ChIP-seq.

(a) Heat maps and plot showing sites that lose accessibility following acute deletion of SMARCA4. (b) Heat maps and plot showing sites that do not change accessibility following acute deletion of SMARCA4. (c) Heat maps and plot showing sites that gain accessibility following acute deletion of SMARCA4. The criteria for classification is provided in the Methods section. (d) Genome-wide signal of ATAC-seq and Ring1b ChIP-seq, in relation to other chromatin features from the Mouse ENCODE Project. ATAC-seq is highly correlated with permissive marks, while Ring1b shows a high degree of correlation with H3K27me3, the mark of Polycomb Repressive Complex 2. (e) Enrichment of enhancers, transcription start sites (TSSs), transcription termination sites (TTSs), and intergenic regions in ATAC-seq datasets by class. ATAC-seq peaks are enriched at enhancers and TSSs. Small deviations between the knockout and WT/G784E expression reflect variability in genome-wide peak calling. (f) Enrichment of enhancers, transcription start sites (TSSs), transcription termination sites (TTSs), and intergenic regions in Ring1b ChIP-seq datasets by class. Ring1b peaks are enriched at TSSs.

Supplementary Figure 6 Analysis of altered chromatin signals and consistency of results across independent cell-culture replicates.

All fold changes below refer to changes that arise in G784E/WT mESCs compared to WT/WT mESCs. (a) MA plot of fold change of ATAC read density at individual sites across the genome. (b) Reproducibility of ATAC changes across independent cell-culture replicates. (c) MA plot of fold change of H3K4me1 read density at individual sites across the genome. (d) Reproducibility of H3K4me1 changes across independent cell-culture replicates. (e) MA plot of fold change of H3K27ac read density at individual sites across the genome. (f) Reproducibility of H3K27ac changes across independent cell-culture replicates. (g) MA plot of fold change of RNAP2 read density at individual sites across the genome. (h) Reproducibility of RNAP2 changes across independent cell-culture replicates.

Supplementary Figure 7 Examples of enhancer changes across the genome.

(a) Example of enhancer accessibility loss coinciding with H3K27ac loss but not H3K4me1 loss on chromosome 8. (b) Example of enhancer accessibility loss coinciding with H3K27ac loss but not H3K4me1 loss on chromosome 10. (c) Example of enhancer accessibility loss coinciding with H3K27ac loss but not H3K4me1 loss on chromosome 2. These changes are representative of the dominant-negative effects of SMARCA4 mutations on the enhancer landscape.

Supplementary Figure 8 Summary of ChIP-seq data at sites with increased and decreased accessibility measured by ATAC-seq.

(a) Summary of individual H3K4me1 ChIP-seq tracks at ATAC-seq sites identified as increased or decreased. (b) Summary of individual RNAP2 ChIP-seq tracks at ATAC-seq sites identified as increased or decreased. (c) Summary of individual H3K27ac ChIP-seq tracks at ATAC-seq sites identified as increased or decreased. (d) Mean RNA polymerase 2 (RNAP2) ChIP-seq read density at TSSs and enhancers in wild-type and heterozygous mutant cells.

Supplementary Figure 9 SMARCA2 (BRM) is repressed in wild-type and mutant SMARCA4 mESCs.

Expression of SMARCA2 compared to other genes in mutant and wild-type mESCs. Little to no SMARCA2 is expressed in mESCs, either in wild-type or mutant cells. Counting of transcripts reveals that SMARCA2 is expressed at less than 1% compared to SMARCA4, hence confounding effects due to SMARCA2 expression can be excluded. Plotted values are mean normalized expression counts (a.u.), error bars are 95% confidence intervals from independent cell-culture replicates (n=2).

Supplementary information

Source data

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Hodges, H.C., Stanton, B.Z., Cermakova, K. et al. Dominant-negative SMARCA4 mutants alter the accessibility landscape of tissue-unrestricted enhancers. Nat Struct Mol Biol 25, 61–72 (2018). https://doi.org/10.1038/s41594-017-0007-3

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41594-017-0007-3

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