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.

  • Letter
  • Published:

SMARCB1-mediated SWI/SNF complex function is essential for enhancer regulation

Abstract

SMARCB1 (also known as SNF5, INI1, and BAF47), a core subunit of the SWI/SNF (BAF) chromatin-remodeling complex1,2, is inactivated in nearly all pediatric rhabdoid tumors3,4,5. These aggressive cancers are among the most genomically stable6,7,8, suggesting an epigenetic mechanism by which SMARCB1 loss drives transformation. Here we show that, despite having indistinguishable mutational landscapes, human rhabdoid tumors exhibit distinct enhancer H3K27ac signatures, which identify remnants of differentiation programs. We show that SMARCB1 is required for the integrity of SWI/SNF complexes and that its loss alters enhancer targeting—markedly impairing SWI/SNF binding to typical enhancers, particularly those required for differentiation, while maintaining SWI/SNF binding at super-enhancers. We show that these retained super-enhancers are essential for rhabdoid tumor survival, including some that are shared by all subtypes, such as SPRY1, and other lineage-specific super-enhancers, such as SOX2 in brain-derived rhabdoid tumors. Taken together, our findings identify a new chromatin-based epigenetic mechanism underlying the tumor-suppressive activity of SMARCB1.

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

Figure 1: Histone modification landscape in primary rhabdoid tumors and cell lines.
Figure 2: SMARCB1 is essential in maintaining SWI/SNF complex integrity.
Figure 3: SMARCB1 re-expression alters SWI/SNF complex targeting at typical enhancers.
Figure 4: Residual SWI/SNF complexes are specifically maintained at super-enhancers in SMARCB1-deficient rhabdoid tumors.
Figure 5: Working model.

Similar content being viewed by others

Accession codes

Primary accessions

Gene Expression Omnibus

References

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

    Article  CAS  Google Scholar 

  2. Wu, J.I., Lessard, J. & Crabtree, G.R. Understanding the words of chromatin regulation. Cell 136, 200–206 (2009).

    Article  CAS  Google Scholar 

  3. 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 

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

    Article  CAS  Google Scholar 

  5. 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 

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

    Article  CAS  Google Scholar 

  7. Hasselblatt, M. et al. High-resolution genomic analysis suggests the absence of recurrent genomic alterations other than SMARCB1 aberrations in atypical teratoid/rhabdoid tumors. Genes Chromosom. Cancer 52, 185–190 (2013).

    Article  CAS  Google Scholar 

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

    Article  CAS  Google Scholar 

  9. Zhou, V.W., Goren, A. & Bernstein, B.E. Charting histone modifications and the functional organization of mammalian genomes. Nat. Rev. Genet. 12, 7–18 (2011).

    Article  Google Scholar 

  10. Zhu, J. et al. Genome-wide chromatin state transitions associated with developmental and environmental cues. Cell 152, 642–654 (2013).

    Article  CAS  Google Scholar 

  11. Roadmap Epigenomics Consortium. et al. Integrative analysis of 111 reference human epigenomes. Nature 518, 317–330 (2015).

  12. Chun, H.-J.E. 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 

  13. Johann, P.D. et al. Atypical teratoid/rhabdoid tumors are comprised of three epigenetic subgroups with distinct enhancer landscapes. Cancer Cell 29, 379–393 (2016).

    Article  CAS  Google Scholar 

  14. Wang, S. et al. Target analysis by integration of transcriptome and ChIP–seq data with BETA. Nat. Protoc. 8, 2502–2515 (2013).

    Article  CAS  Google Scholar 

  15. Stappenbeck, T.S. & Miyoshi, H. The role of stromal stem cells in tissue regeneration and wound repair. Science 324, 1666–1669 (2009).

    Article  CAS  Google Scholar 

  16. 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 

  17. Hnisz, D. et al. Super-enhancers in the control of cell identity and disease. Cell 155, 934–947 (2013).

    Article  CAS  Google Scholar 

  18. Bergsland, M. et al. Sequentially acting Sox transcription factors in neural lineage development. Genes Dev. 25, 2453–2464 (2011).

    Article  CAS  Google Scholar 

  19. Yu, H. et al. Frizzled 1 and frizzled 2 genes function in palate, ventricular septum and neural tube closure: general implications for tissue fusion processes. Development 137, 3707–3717 (2010).

    Article  CAS  Google Scholar 

  20. Abrahams, A., Parker, M.I. & Prince, S. The T-box transcription factor Tbx2: its role in development and possible implication in cancer. IUBMB Life 62, 92–102 (2010).

    CAS  PubMed  Google Scholar 

  21. 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 

  22. Isakoff, M.S. et al. Inactivation of the Snf5 tumor suppressor stimulates cell cycle progression and cooperates with p53 loss in oncogenic transformation. Proc. Natl. Acad. Sci. USA 102, 17745–17750 (2005).

    Article  CAS  Google Scholar 

  23. Tolstorukov, M.Y. et al. Swi/Snf chromatin remodeling/tumor suppressor complex establishes nucleosome occupancy at target promoters. Proc. Natl. Acad. Sci. USA 110, 10165–10170 (2013).

    Article  CAS  Google Scholar 

  24. Yu, Y. et al. Olig2 targets chromatin remodelers to enhancers to initiate oligodendrocyte differentiation. Cell 152, 248–261 (2013).

    Article  CAS  Google Scholar 

  25. Bossen, C. et al. The chromatin remodeler Brg1 activates enhancer repertoires to establish B cell identity and modulate cell growth. Nat. Immunol. 16, 775–784 (2015).

    Article  CAS  Google Scholar 

  26. Alexander, J.M. et al. Brg1 modulates enhancer activation in mesoderm lineage commitment. Development 142, 1418–1430 (2015).

    Article  CAS  Google Scholar 

  27. Doan, D.N. et al. Loss of the INI1 tumor suppressor does not impair the expression of multiple BRG1-dependent genes or the assembly of SWI/SNF enzymes. Oncogene 23, 3462–3473 (2004).

    Article  CAS  Google Scholar 

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

    Article  Google Scholar 

  29. Kaur, H. et al. The chromatin-modifying protein HMGA2 promotes atypical teratoid/rhabdoid cell tumorigenicity. J. Neuropathol. Exp. Neurol. 74, 177–185 (2015).

    Article  CAS  Google Scholar 

  30. Wang, X. et al. TCR-dependent transformation of mature memory phenotype T cells in mice. J. Clin. Invest. 121, 3834–3845 (2011).

    Article  CAS  Google Scholar 

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

    Article  CAS  Google Scholar 

  32. 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 

  33. 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  Google Scholar 

  34. Kharchenko, P.V., Tolstorukov, M.Y. & Park, P.J. Design and analysis of ChIP–seq experiments for DNA-binding proteins. Nat. Biotechnol. 26, 1351–1359 (2008).

    Article  CAS  Google Scholar 

  35. de Hoon, M.J.L., Imoto, S., Nolan, J. & Miyano, S. Open source clustering software. Bioinformatics 20, 1453–1454 (2004).

    Article  CAS  Google Scholar 

  36. Saldanha, A.J. Java Treeview—extensible visualization of microarray data. Bioinformatics 20, 3246–3248 (2004).

    Article  CAS  Google Scholar 

  37. Kim, D. et al. TopHat2: accurate alignment of transcriptomes in the presence of insertions, deletions and gene fusions. Genome Biol. 14, R36 (2013).

    Article  Google Scholar 

  38. Trapnell, C. et al. Differential analysis of gene regulation at transcript resolution with RNA–seq. Nat. Biotechnol. 31, 46–53 (2013).

    Article  CAS  Google Scholar 

  39. 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 

Download references

Acknowledgements

We thank members of the Roberts and Park laboratories for assistance and discussion. We thank J. Francois (Boston Children's Hospital), J. Roth (Children's Hospital of Philadelphia), M. Lear (St. Jude Children's Research Hospital), and J. Silterra (Broad Institute) for their help in acquiring and performing preliminary clinical analysis of primary tumor samples; R. Rubio (Dana-Farber Cancer Institute) and M. Uziel (Broad Institute) for their assistance in sequencing samples; N. Shoresh (Broad Institute) for assistance in accessing Roadmap Epigenomics data; and R. Tomaino (Harvard Medical School) for assistance in proteomic analysis. X.W. was supported by the Pathway to Independence Award from the US National Institutes of Health (K99CA197640), a postdoctoral fellowship from the Rally Foundation for Childhood Cancer Research and The Truth 365, and a research grant from St. Baldrick's Foundation. R.S.L. was partially supported by an NSF Graduate Research Fellowship. This work was supported by US National Institutes of Health grants R01CA172152 (C.W.M.R.), R01CA113794 (C.W.M.R.), and U54HG006991 (B.E.B.). The Avalanna Fund, the Cure AT/RT Now Foundation, the Garrett B. Smith Foundation, Miles for Mary, and ALSAC/St. Jude (C.W.M.R.) provided additional support.

Author information

Authors and Affiliations

Authors

Contributions

X.W., R.S.L., B.H.A., B.E.B., P.J.P., and C.W.M.R. conceived the experiments and study design. X.W., J.R.H., E.P.T., and E.C.T. performed all cell line experiments. R.S.L. and S.M.G. performed all primary tumor experiments. B.H.A., R.S.L., S.W., J.M., and Y.D. performed computational analyses of the data. X.W., R.S.L., B.H.A., and S.W. performed statistical analyses. T.C.A., S.L.P., and J.A.B. contributed primary tumor samples and clinical data. J.N.W. designed the SMARCB1 re-expression vector. X.W., R.S.L., B.H.A., J.R.H., J.M., Y.D., M.Y.T., B.E.B., P.J.P., and C.W.M.R. contributed to the interpretation of experiments. X.W., R.S.L., B.H.A., and C.W.M.R. wrote the manuscript with input from all co-authors.

Corresponding authors

Correspondence to Bradley E Bernstein, Peter J Park or Charles W M Roberts.

Ethics declarations

Competing interests

The authors declare no competing financial interests.

Integrated supplementary information

Supplementary Figure 1 Histone modification landscape in primary rhabdoid tumors and cell lines.

Rhabdoid tumors from different tissues show commonalities in H3K4me1 or H3K4me3 signal but are clearly distinct in terms of H3K27ac signal at enhancers. (a) Summary of primary rhabdoid tumor samples and cell lines used in this study and types of ChIP–seq or RNA–seq performed in these samples. (bd) The first two principal components of H3K4me3 (b), H3K4me1 (c), and H3K27ac (d) signal across the union of all promoters or enhancers across the samples.

Supplementary Figure 2 SMARCB1 is essential in maintaining SWI/SNF complex integrity.

(a) mRNA levels of the indicated SWI/SNF complex subunits upon SMARCB1 re-expression (G401, BT16) shown by RNA–seq. (b) Protein levels of the indicated SWI/SNF complex subunits upon MG132 treatment in BT16 cells. (ce) Immunoprecipitation (IP) of the SWI/SNF complex by SMARCC1, SMARCA4, or ARID1A from the nuclear extracts of A204 (b), TM87 (d), and ES-2 (e) cells, before or after Dox induction (SMARCB1 re-expression); immunoblotting was performed for SWI/SNF complex SMARCB1, SMARCC1, SMARCA4, ARID1A, ARID1B, SMARCC2, SMARCD1, SMARCE1, ACTL6A, and DPF2. A204 and TM87 are both SMARCB1-mutant rhabdoid tumor cell lines derived from soft tissues, and ES-2 is an ovarian cancer cell line with wild-type SMARCB1. (f) SMARCC1 immunoprecipitation and mass spectrometry results in BT16 cells with or without SMARCB1 re-expression. (g) Glycerol sedimentation (10–30%) assay of the SWI/SNF complex from SMARCB1-mutant BT16 cells without Dox (top) or with Dox (bottom) and immunoblotting for the indicated SWI/SNF complex subunits. (h) mRNA levels of the indicated SWI/SNF complex subunits upon SMARCB1 re-expression (TM87) or SMARCB1 loss (MEFs) shown by RNA–seq.

Supplementary Figure 3 Correlation of the changes in SMARCA4 and SMARCC1 binding upon re-expression of SMARCB1 in G401, BT16, and TTC549 cells.

(ac) Log2-tranformed fold change in SMARCA4 signal versus log2-transformed fold change in SMARCC1 signal at SWI/SNF binding sites upon SMARCB1 re-expression in G401 (a), BT16 (b), and TTC549 (c) cells. The two subunits show consistent changes under the two conditions. Therefore, data for the two subunits were used together to identify SWI/SNF binding sites. The different colors denote sites classified as ‘gained/strengthened’, ‘stable’, or ‘weakened/lost’.

Supplementary Figure 4 SMARCB1 re-expression alters the SWI/SNF complex targeting at typical enhancers in BT16 cells.

(a) Number of SWI/SNF (SMARCC1 and SMARCA4) binding sites in regions of enrichment for different histone marks in BT16 cells with or without SMARCB1 re-expression. (b) Average enrichment of SMARCC1 and SMARCA4 without versus with SMARCB1 re-expression in TSS-distal or TSS-proximal SWI/SNF binding sites. (c) Heat maps depicting SMARCC1, SMARCA4, H3K27ac, H3K4me1, and H3K4me3 signal intensities for TSS-distal SWI/SNF binding sites, grouped by change upon SMARCB1 re-expression. The rows show 9-kb regions, centered on SMARCC1/SMARCA4 peaks, ranked by overall signal intensities of SMARCC1/SMARCA4. The average profiles for each heat map are shown above, where different y-axis ranges are denoted as 1/2× or 1/4×. (d) Representative screenshot of SMARCC1, SMARCA4, H3K4me1, H3K4me3, and H3K27ac signal with or without SMARCB1 re-expression in G401 cells showing increased SMARCC1 and SMARCA4 binding upon SMARCB1 re-expression accompanied by increased and flanking H3K27ac and H3K4me1 marks at enhancers. (e) Correlation of gene expression changes with SMARCC1/SMARCA4 binding or H3K27ac signal at TSS-distal binding sites in BT16 cells upon Dox treatment. (f) GO analysis of genes proximal to enhancers with increased SMARCC1/SMARCA4 signal upon Dox treatment.

Supplementary Figure 5 SMARCB1 re-expression alters SWI/SNF complex targeting at typical enhancers in TTC549 cells.

(a) Number of SWI/SNF (SMARCC1 and SMARCA4) binding sites in regions of enrichment for different histone marks in TTC549 cells with or without SMARCB1 re-expression. (b) Average enrichment of SMARCC1 and SMARCA4 without versus with SMARCB1 re-expression in TSS-distal or TSS-proximal SWI/SNF binding sites. (c) Heat maps depicting SMARCC1, SMARCA4, H3K27ac, H3K4me1, and H3K4me3 signal intensities for TSS-distal SWI/SNF binding sites, grouped by change upon SMARCB1 re-expression. The rows show 9-kb regions, centered on SMARCC1/SMARCA4 peaks, ranked by overall signal intensity of SMARCC1/SMARCA4. The average profiles for each heat map are shown above, where different y-axis ranges are denoted as 1/2× or 1/4×. (d) Representative screenshot of SMARCC1, SMARCA4, H3K4me1, H3K4me3, and H3K27ac signal with or without SMARCB1 re-expression in G401 cells showing increased SMARCC1 and SMARCA4 binding upon SMARCB1 re-expression accompanied by increased and flanking H3K27ac and H3K4me1 marks at enhancers. (e) Correlation of gene expression changes with SMARCC1/SMARCA4 binding or H3K27ac signal at TSS-distal binding sites in TTC549 cells upon Dox treatment. (f) GO analysis of genes proximal to enhancers with increased SMARCC1/SMARCA4 signal upon Dox treatment.

Supplementary Figure 6 RNA–seq analysis of six different rhabdoid tumor cell lines upon SMARCB1 re-expression.

k-means clustering of log2 (fold change) values (Dox/no Dox) for rhabdoid tumor cell lines to ten clusters. Selected enriched GO terms for each cluster (FDR < 0.0001 or ‘development’ terms with FDR < 0.01) are shown.

Supplementary Figure 7 Gene set enrichment analysis of different rhabdoid tumor cell lines upon SMARCB1 re-expression.

GSEA gene sets most enriched in upregulated and downregulated differential expression patterns.

Supplementary Figure 8 Re-expression of SMARCB1 has little effect on the number and structure of super-enhancers in three different rhabdoid tumor cell lines.

(ac) Number of H3K27ac RoEs present under each condition (called in the sample or having greater than 2/3 signal relative to the other condition) in different RoE categories and input-subtracted H3K27ac ChIP signal in stitched H3K27ac regions, excluding signal within H3K4me3 regions, with or without SMARCB1 re-expression in the G401 (a), BT16 (b), and TTC549 (c) cell lines. Red dots indicate super-enhancer calls. Middle plots show ranked values, whereas right plots compare the two conditions on a log scale.

Supplementary Figure 9 Residual SWI/SNF complexes are specifically maintained at super-enhancers in SMARCB1-mutant rhabdoid tumors.

(a,b) Heat maps depicting SMARCC1, SMARCA4, H3K27ac, H3K4me1, and H3K4me3 signal intensities for SMARCC1/SMARCA4 bound within super-enhancers in BT16 (a) and TTC549 (b) cells. (c,d) Scatterplots showing changes in average SMARCC1/SMARCA4 signal for TSS-proximal binding sites or TSS-distal ones split into those outside and inside super-enhancers in BT16 (c) and TTC549 (d) cells. (e,f) Representative screenshots for BT16 (e) and TTC549 (f) cells showing limited changes in SMARCC1, SMARCA4, H3K27ac, or H3K4me1 upon SMARCB1 re-expression inside super-enhancers, in contrast to outside super-enhancers.

Supplementary Figure 10 Residual SWI/SNF complex binds to super-enhancers and regulates super-enhancer gene expression in G401 cells.

(a) Residual SWI/SNF complex binding at a typical enhancer is increased upon SMARCB1 re-expression and is accompanied by increased H3K27ac marks at the gene QKI enhancer. (b) Residual SWI/SNF complex binding at super-enhancers, as well as H3K27ac marks, is only modestly increased upon SMARCB1 re-expression but is largely reduced upon further loss of residual complex subunit SMARCC1 or SMARCD1, as shown at selected loci for two super-enhancer genes (SALL4 and SPRY1). (c,d) The expression of super-enhancer genes is significantly reduced upon loss of residual complex subunit SMARCC1 (c) or SMARCD1 (d).

Supplementary Figure 11 Screenshots of H3K27ac super-enhancers in primary rhabdoid tumors and cell lines.

(a) Super-enhancers are found at the SPRY1 locus in all primary rhabdoid tumors and cell lines. The one sample in which a super-enhancer is not called using our algorithm (RT262) still shows significant H3K27ac signal that overlaps with the peaks observed in the other samples. (b) Super-enhancers are found at the SALL4 locus in all primary rhabdoid tumors and all cell lines. (c) Super-enhancers are found at the HMGA2 locus in six primary rhabdoid tumors and all cell lines. (d) Super-enhancers are found at the SOX2 locus in all primary rhabdoid tumors and all cell lines.

Supplementary information

Supplementary Text and Figures

Supplementary Figures 1–11. (PDF 2475 kb)

Supplementary Table 1

Rhabdoid tumor primary GO terms. (XLSX 53 kb)

Supplementary Table 2

Rhabdoid tumor primary super-enhancers. (XLSX 209 kb)

Supplementary Table 3

Mass spectrometry. (XLSX 162 kb)

Supplementary Table 4

Cell line GO terms. (XLS 140 kb)

Supplementary Data

Full scans of blots. (PDF 30341 kb)

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Wang, X., Lee, R., Alver, B. et al. SMARCB1-mediated SWI/SNF complex function is essential for enhancer regulation. Nat Genet 49, 289–295 (2017). https://doi.org/10.1038/ng.3746

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/ng.3746

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