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:

ARID1A determines luminal identity and therapeutic response in estrogen-receptor-positive breast cancer

Subjects

Abstract

Mutations in ARID1A, a subunit of the SWI/SNF chromatin remodeling complex, are the most common alterations of the SWI/SNF complex in estrogen-receptor-positive (ER+) breast cancer. We identify that ARID1A inactivating mutations are present at a high frequency in advanced endocrine-resistant ER+ breast cancer. An epigenome CRISPR–CAS9 knockout (KO) screen identifies ARID1A as the top candidate whose loss determines resistance to the ER degrader fulvestrant. ARID1A inactivation in cells and in patients leads to resistance to ER degraders by facilitating a switch from ER-dependent luminal cells to ER-independent basal-like cells. Cellular plasticity is mediated by loss of ARID1A-dependent SWI/SNF complex targeting to genomic sites of the luminal lineage-determining transcription factors including ER, forkhead box protein A1 (FOXA1) and GATA-binding factor 3 (GATA3). ARID1A also regulates genome-wide ER–FOXA1 chromatin interactions and ER-dependent transcription. Altogether, we uncover a critical role for ARID1A in maintaining luminal cell identity and endocrine therapeutic response in ER+ breast cancer.

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: ARID1A loss mediates endocrine therapy resistance.
Fig. 2: ARID1A impacts the accessibility of several transcription factor motifs involved in luminal differentiation.
Fig. 3: ARID1A loss results in enrichment of a basal-like signature in cells and patient samples.
Fig. 4: ARID1A loss causes defects in SWI/SNF targeting to chromatin at luminal lineage-determining transcription factor loci.
Fig. 5: ARID1A regulates ER and FOXA1 chromatin occupancy and ER-dependent transcription.
Fig. 6: Proposed model.

Similar content being viewed by others

Data availability

Sequencing data have been deposited with the Gene Expression Omnibus under accession no. GSE124228. The epigenome CRISPR–CAS9 screen information can be found in Supplementary Table 2. Source data for Figs. 1 and 3 and Extended Data Figs. 26 are available online.

References

  1. Perou, C. M. et al. Molecular portraits of human breast tumours. Nature 406, 747–752 (2000).

    CAS  PubMed  Google Scholar 

  2. Sørlie, T. et al. Gene expression patterns of breast carcinomas distinguish tumor subclasses with clinical implications. Proc. Natl Acad. Sci. USA 98, 10869–10874 (2001).

    PubMed  PubMed Central  Google Scholar 

  3. Koboldt, D. C. et al. Comprehensive molecular portraits of human breast tumours. Nature 490, 61–70 (2012).

    CAS  Google Scholar 

  4. Stephens, P. J. et al. The landscape of cancer genes and mutational processes in breast cancer. Nature 486, 400–404 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  5. Ciriello, G. et al. Comprehensive molecular portraits of invasive lobular breast cancer. Cell 163, 506–519 (2015).

    CAS  PubMed  PubMed Central  Google Scholar 

  6. Banerji, S. et al. Sequence analysis of mutations and translocations across breast cancer subtypes. Nature 486, 405–409 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  7. Shah, S. P. et al. The clonal and mutational evolution spectrum of primary triple-negative breast cancers. Nature 486, 395–399 (2012).

    CAS  PubMed  Google Scholar 

  8. Ellis, M. J. et al. Whole-genome analysis informs breast cancer response to aromatase inhibition. Nature 486, 353–360 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  9. Nik-Zainal, S. et al. Landscape of somatic mutations in 560 breast cancer whole-genome sequences. Nature 534, 47–54 (2016).

    CAS  PubMed  PubMed Central  Google Scholar 

  10. Pereira, B. et al. The somatic mutation profiles of 2,433 breast cancers refines their genomic and transcriptomic landscapes. Nat. Commun. 7, 11479 (2016).

    CAS  PubMed  PubMed Central  Google Scholar 

  11. Green, K. A. & Carroll, J. S. Oestrogen-receptor-mediated transcription and the influence of co-factors and chromatin state. Nat. Rev. Cancer 7, 713–722 (2007).

    CAS  PubMed  Google Scholar 

  12. Schiavon, G. et al. Analysis of ESR1 mutation in circulating tumor DNA demonstrates evolution during therapy for metastatic breast cancer. Sci. Transl. Med. 7, 313ra182 (2015).

    PubMed  PubMed Central  Google Scholar 

  13. Toy, W. et al. ESR1 ligand-binding domain mutations in hormone-resistant breast cancer. Nat. Genet. 45, 1439–1445 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  14. 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).

    CAS  PubMed  PubMed Central  Google Scholar 

  15. Garraway, L. A. & Lander, E. S. Lessons from the cancer genome. Cell 153, 17–37 (2013).

    CAS  PubMed  Google Scholar 

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

    CAS  PubMed  Google Scholar 

  17. 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).

    CAS  PubMed  PubMed Central  Google Scholar 

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

    CAS  PubMed  Google Scholar 

  19. Kelso, T. W. R. et al. Chromatin accessibility underlies synthetic lethality of SWI/SNF subunits in ARID1A-mutant cancers. eLife 6, e30506 (2017).

    PubMed  PubMed Central  Google Scholar 

  20. 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).

    CAS  PubMed  PubMed Central  Google Scholar 

  21. Mashtalir, N. et al. Modular organization and assembly of SWI/SNF family chromatin remodeling complexes. Cell 175, 1272–1288.e20 (2018).

    CAS  PubMed  PubMed Central  Google Scholar 

  22. Razavi, P. et al. The genomic landscape of endocrine-resistant advanced breast cancers. Cancer Cell 34, 427–438.e6 (2018).

    CAS  PubMed  PubMed Central  Google Scholar 

  23. Cheng, D. T. et al. Memorial Sloan Kettering-integrated mutation profiling of actionable cancer targets (MSK-IMPACT): a hybridization capture-based next-generation sequencing clinical assay for solid tumor molecular oncology. J. Mol. Diagn. 17, 251–264 (2015).

    CAS  PubMed  PubMed Central  Google Scholar 

  24. Pan, J. et al. Interrogation of mammalian protein complex structure, function, and membership using genome-scale fitness screens. Cell Syst. 6, 555–568.e7 (2018).

    CAS  PubMed  PubMed Central  Google Scholar 

  25. Sandoval, G. J. et al. Binding of TMPRSS2-ERG to BAF chromatin remodeling complexes mediates prostate oncogenesis. Mol. Cell 71, 554–566.e7 (2018).

    CAS  PubMed  PubMed Central  Google Scholar 

  26. Grant, C. E., Bailey, T. L. & Noble, W. S. FIMO: scanning for occurrences of a given motif. Bioinformatics 27, 1017–1018 (2011).

    CAS  PubMed  PubMed Central  Google Scholar 

  27. Hurtado, A., Holmes, K. A., Ross-Innes, C. S., Schmidt, D. & Carroll, J. S. FOXA1 is a key determinant of estrogen receptor function and endocrine response. Nat. Genet. 43, 27–33 (2011).

    CAS  PubMed  Google Scholar 

  28. Bernardo, G. M. et al. FOXA1 is an essential determinant of ERα expression and mammary ductal morphogenesis. Development 137, 2045–2054 (2010).

    CAS  PubMed  PubMed Central  Google Scholar 

  29. Theodorou, V., Stark, R., Menon, S. & Carroll, J. S. GATA3 acts upstream of FOXA1 in mediating ESR1 binding by shaping enhancer accessibility. Genome Res. 23, 12–22 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  30. Asselin-Labat, M. L. et al. Gata-3 is an essential regulator of mammary-gland morphogenesis and luminal-cell differentiation. Nat. Cell Biol. 9, 201–209 (2007).

    CAS  PubMed  Google Scholar 

  31. Dravis, C. et al. Epigenetic and transcriptomic profiling of mammary gland development and tumor models disclose regulators of cell state plasticity. Cancer Cell 34, 466–482.e6 (2018).

    CAS  PubMed  PubMed Central  Google Scholar 

  32. Mu, P. et al. SOX2 promotes lineage plasticity and antiandrogen resistance in TP53- and RB1-deficient prostate cancer. Science 355, 84–88 (2017).

    CAS  PubMed  PubMed Central  Google Scholar 

  33. Britschgi, A. et al. The Hippo kinases LATS1 and 2 control human breast cell fate via crosstalk with ERα. Nature 541, 541–545 (2017).

    CAS  PubMed  PubMed Central  Google Scholar 

  34. Malta, T. M. et al. Machine learning identifies stemness features associated with oncogenic dedifferentiation. Cell 173, 338–354.e15 (2018).

    CAS  PubMed  PubMed Central  Google Scholar 

  35. Wuidart, A. et al. Early lineage segregation of multipotent embryonic mammary gland progenitors. Nat. Cell Biol. 20, 666–676 (2018).

    CAS  PubMed  PubMed Central  Google Scholar 

  36. Kadoch, C. et al. Dynamics of BAF–Polycomb complex opposition on heterochromatin in normal and oncogenic states. Nat. Genet. 49, 213–222 (2017).

    CAS  PubMed  Google Scholar 

  37. 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).

    CAS  PubMed  PubMed Central  Google Scholar 

  38. Sun, X. et al. Arid1a has context-dependent oncogenic and tumor suppressor functions in liver cancer. Cancer Cell 32, 574–589.e6 (2017).

    CAS  PubMed  PubMed Central  Google Scholar 

  39. Vierbuchen, T. et al. AP-1 transcription factors and the BAF complex mediate signal-dependent enhancer selection. Mol. Cell 68, 1067–1082.e12 (2017).

    CAS  PubMed  PubMed Central  Google Scholar 

  40. Zaret, K. S. & Carroll, J. S. Pioneer transcription factors: establishing competence for gene expression. Genes Dev. 25, 2227–2241 (2011).

    CAS  PubMed  PubMed Central  Google Scholar 

  41. Lupien, M. et al. FoxA1 translates epigenetic signatures into enhancer-driven lineage-specific transcription. Cell 132, 958–970 (2008).

    CAS  PubMed  PubMed Central  Google Scholar 

  42. Toska, E. et al. PI3K pathway regulates ER-dependent transcription in breast cancer through the epigenetic regulator KMT2D. Science 355, 1324–1330 (2017).

    CAS  PubMed  PubMed Central  Google Scholar 

  43. Kuukasjärvi, T., Kononen, J., Helin, H., Holli, K. & Isola, J. Loss of estrogen receptor in recurrent breast cancer is associated with poor response to endocrine therapy. J. Clin. Oncol. 14, 2584–2589 (1996).

    PubMed  Google Scholar 

  44. Musgrove, E. A. & Sutherland, R. L. Biological determinants of endocrine resistance in breast cancer. Nat. Rev. Cancer 9, 631–643 (2009).

    CAS  PubMed  Google Scholar 

  45. Lindström, L. S. et al. Clinically used breast cancer markers such as estrogen receptor, progesterone receptor, and human epidermal growth factor receptor 2 are unstable throughout tumor progression. J. Clin. Oncol. 30, 2601–2608 (2012).

    PubMed  Google Scholar 

  46. Dieci, M. V. et al. Discordance in receptor status between primary and recurrent breast cancer has a prognostic impact: a single-institution analysis. Ann. Oncol. 24, 101–108 (2013).

    CAS  PubMed  Google Scholar 

  47. Caumanns, J. J., Wisman, G. B. A., Berns, K., van der Zee, A. G. J. & de Jong, S. ARID1A mutant ovarian clear cell carcinoma: a clear target for synthetic lethal strategies. Biochim. Biophys. Acta Rev. Cancer 1870, 176–184 (2018).

    CAS  PubMed  Google Scholar 

  48. Bitler, B. G. et al. Synthetic lethality by targeting EZH2 methyltransferase activity in ARID1A-mutated cancers. Nat. Med. 21, 231–238 (2015).

    CAS  PubMed  Google Scholar 

  49. Doench, J. G. et al. Optimized sgRNA design to maximize activity and minimize off-target effects of CRISPR-Cas9. Nat. Biotechnol. 34, 184–191 (2016).

    CAS  PubMed  PubMed Central  Google Scholar 

  50. Toska, E. et al. PI3K inhibition activates SGK1 via a feedback loop to promote chromatin-based regulation of ER-dependent gene expression. Cell Rep. 27, 294–306.e5 (2019).

    CAS  PubMed  PubMed Central  Google Scholar 

  51. 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).

    CAS  PubMed  PubMed Central  Google Scholar 

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

    CAS  PubMed  PubMed Central  Google Scholar 

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

    CAS  PubMed  PubMed Central  Google Scholar 

  54. Feng, J., Liu, T., Qin, B., Zhang, Y. & Liu, X. S. Identifying ChIP-seq enrichment using MACS. Nat. Protoc. 7, 1728–1740 (2012).

    CAS  PubMed  Google Scholar 

  55. Lawrence, M. et al. Software for computing and annotating genomic ranges. PLoS Comput. Biol. 9, e1003118 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  56. 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).

    PubMed  PubMed Central  Google Scholar 

  57. Philip, M. et al. Chromatin states define tumour-specific T cell dysfunction and reprogramming. Nature 545, 452–456 (2017).

    CAS  PubMed  PubMed Central  Google Scholar 

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

    PubMed  PubMed Central  Google Scholar 

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

    CAS  PubMed  Google Scholar 

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

    CAS  PubMed  Google Scholar 

  61. Chen, C. W. et al. DOT1L inhibits SIRT1-mediated epigenetic silencing to maintain leukemic gene expression in MLL-rearranged leukemia. Nat. Med. 21, 335–343 (2015).

    CAS  PubMed  PubMed Central  Google Scholar 

  62. Bailey, T. L. & Machanick, P. Inferring direct DNA binding from ChIP-seq. Nucleic Acids Res. 40, e128 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  63. Ramírez, F., Dündar, F., Diehl, S., Grüning, B. A. & Manke, T. deepTools: a flexible platform for exploring deep-sequencing data. Nucleic Acids Res. 42, W187–W191 (2014).

    PubMed  PubMed Central  Google Scholar 

  64. Shen, R. & Seshan, V. E. FACETS: allele-specific copy number and clonal heterogeneity analysis tool for high-throughput DNA sequencing. Nucleic Acids Res. 44, e131 (2016).

    PubMed  PubMed Central  Google Scholar 

  65. Weinreb, I. et al. Hotspot activating PRKD1 somatic mutations in polymorphous low-grade adenocarcinomas of the salivary glands. Nat. Genet. 46, 1166–1169 (2014).

    CAS  PubMed  Google Scholar 

Download references

Acknowledgements

We thank the Center for Epigenetics Research at MSKCC for help with the ATAC-seq and ChIP-seq assays, and E. de Stanchina and the Antitumor Assessment Core Facility for help with xenografts establishment. We also thank the Baselga and Scaltriti laboratory members for helpful advice and discussions. We thank A. Del from the Department of Pathology at MSKCC for the procurement of the FFPE slides. This work has been supported by National Institutes of Health grant nos. P30 CA008748 and RO1CA190642-01A1, the Breast Cancer Research Foundation grant no. BCRF-17-013. E.T. and M. Scaltriti are supported by a kind gift from B. Smith. E.L. is supported by grant no. NCI K00CA212478. This work was also supported by grants from Stand Up to Cancer (Cancer Drug Combination Convergence Team) grant no. SU2C 2015-004, the V Foundation grant no. D2015-036 and the National Science Foundation grant no. PHY-1545853 (G.X. and M. Scaltriti). This work was also funded by a U54 award grant no. CA209975-01 to C.S.L. E.C. is a recipient of an MSK Society Scholar Prize.

Author information

Authors and Affiliations

Authors

Contributions

G.X., S.C., C.S.L., J.B. and E.T. conceived the project. G.X., E.C., J.E.O., Y. Cai, C.C., A.D.P., M.W., Y.Cheng, J.P., F.W., M. Sallaku, A.Z. and E.T. performed the experiments. S.C. performed all the computational analyses and statistical calculations supervised by C.S.L. L.F. and P.S. performed the survival analyses supervised by J.S.R., E.L., C.K.C., A.R.D., R.K. and X.Q. assisted with the initial computational analyses. H.Z. helped with the mouse in vivo experiment. K.J. assisted with patient selection. P.R. performed the nested control study, and assisted with patient sample procurement and survival analyses. P.R. and C.S. also performed the patient clinical annotation. J.S.R. viewed the FFPE slides, performed the laser microdissection and provided intellectual support. C.K. supervised the SWI/SNF complex ChIP-seq, helped with the SWI/SNF complex ChIP-seq data interpretation and provided intellectual insights. R.L.L. and M. Scaltriti contributed intellectual insights regarding study design and manuscript writing. E.T., G.X., M. Scaltriti, C.S.L. and J.B. wrote the manuscript with help from all the authors.

Corresponding authors

Correspondence to Christina S. Leslie, José Baselga or Eneda Toska.

Ethics declarations

Competing interests

M.Scaltriti has received research funds from Puma Biotechnology, Daiichi Sankyo, Immunomedics, TargImmune Therapeutics and Menarini Ricerche, is a cofounder of Medendi Medical Travel and is on the advisory board of Menarini Ricerche. C.K. is a scientific founder, fiduciary Board of Directors member, Scientific Advisory Board member, shareholder and consultant for Foghorn Therapeutics. R.L.L. is on the supervisory board of QIAGEN and is a scientific advisor to Loxo Oncology, Imago, C4 Therapeutics and Isoplexis, each including an equity interest. He receives research support from and consulted for Celgene and Roche, has received research support from Prelude Therapeutics and has consulted for Incyte, Novartis, MorphoSys and Janssen. He has received honoraria from Eli Lilly and Amgen for invited lectures and from Gilead Sciences for grant reviews. J.B. is an employee and shareholder of AstraZeneca, Board of Directors member of Foghorn Therapeutics and is a past board member of Varian Medical Systems, Bristol‐Myers Squibb, Grail, Aura Biosciences and Infinity Pharmaceuticals. He has performed consulting and/or advisory work for Grail, PMV Pharma, ApoGen Biotechnologies, Juno, Eli Lilly, Seragon Pharmaceuticals, Novartis and Northern Biologics. He has stock or other ownership interests in PMV Pharma, Grail, Juno, Varian Medical Systems, Foghorn Therapeutics, Aura Biosciences, Infinity Pharmaceuticals and ApoGen Biotechnologies, as well as Tango Therapeutics and Venthera, of which he is a cofounder. He has previously received honoraria or travel expenses from Roche, Novartis and Eli Lilly. P.R. has received consultation fees from Novartis and institutional research funds from Grail and Illumina. J.S.R. is a consultant of Goldman Sachs and Repare Therapeutics, a member of the Scientific Advisory Board of VolitionRx and Paige (Artificial Intelligence) and an ad hoc member of the Scientific Advisory Board of Ventana Medical Systems, Roche, Genentech, Novartis and InviCRO, outside of the scope of the submitted work. E.T. has received honoraria from AstraZeneca for invited lectures. No potential conflicts of interests were disclosed by the other authors.

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 Enrichment of mutations of core subunits of the SWI/SNF complex in HR+ HER2- breast cancer.

(a) Mutation enrichment based on IMPACT study. (b) Mutation enrichment based on TCGA and METABRIC studies.

Extended Data Fig. 2 Loss of SWI/SNF complex subunits mediate resistance to endocrine therapy.

(a) In vitro proliferation of ARID1A knockout (KO) MCF7 cells as measured by cell quantification. (b) Cell cycle distributions as measured by FACS analyses of control and ARID1A KO MCF7 cells. Error bars=mean ±SEM, n=2 biologically independent samples, center values are means. P values were calculated using two-way ANOVA test; all P values > 0.2. N.S=non-significant. (c) Cell quantification of ARID1A KO vs. control cells upon fulvestrant treatment (100nM). (d) In vitro proliferation assay in ARID1A KO vs. control cells upon a dose response of the ER degrader GDC0927. The experiments were repeated thrice with similar results. (e) Cell quantification of ARID1A KO vs. control cells under estrogen (E2) depleted media vs. full media. (f) In vitro proliferation assay of ARID1A KO vs. control cells in estrogen depleted media and full media. The experiments were repeated thrice with similar results. (g) Cropped western blot of SMARCB1 or SMARCE1 KO (sg1-sg5) vs. control MCF7 cells. (h) In vitro proliferation assay in SMARCB1 or SMARCE1 KO vs. control MCF7 cells upon treatment with fulvestrant (100nM). The experiments were repeated three times with similar results. (i) The ratio of RFP+ SMARCB1 or SMARCE1 (sg1-sg5) knockout cells to GFP+ control cells (sgNT-GFP) upon DMSO or fulvestrant treatment (100nM) for 8 days as measured by flow cytometry. For (a), (c), (e), (i), error bars=mean ±SEM, n=3 biologically independent samples, center values are means. P values, Student’s two-sided t test.

Source data

Extended Data Fig. 3 ARID1A knockout leads to equal chromatin accessibility changes in DMSO or fulvestrant setting.

(a) Cropped western blot with indicated antibodies in MCF7 cells. (b) Pie chart of peak distributions to various genic parts. (c) ATAC-seq analysis revealed 59,000 peaks in total; 33% in intergenic regions, ~30% in promoter regions, and 35% in intron regions. Violin plot shows probability density of peaks across the samples. (d) Heatmap of differential peaks in control vs. ARID1A KO (knockout) upon DMSO or fulvestrant (fulv) treatment (absolute log2 fold change > 0.5, Benjamini-Hochberg adjusted P < 0.05). (e) ChIP-qPCR analyses of TEAD4 binding in control and ARID1A KO MCF7. Error bars=mean ±SEM, n=2 biologically independent samples, center values are means. P values, Student’s two-sided t test. (f) Cropped western blot of TEAD4 in control and ARID1A KO MCF7. (g) Crystal violet assay of TEAD4 knockdown cells in control and ARID1A KO upon DMSO or fulvestrant (100nM). (h) Cropped western blot of GATA3 overexpression in MCF7 (n=3). (i) In vitro proliferation of GATA3 overexpressed cells in control and ARID1A KO setting upon DMSO or fulvestrant treatment (100nM). The experiments were repeated thrice with similar results. (j) Cell quantification of GATA3 overexpressed cells in control and ARID1A KO setting upon DMSO or fulvestrant treatment (100nM). Error bars=mean ±SEM, n=3 biologically independent samples, center values are means. P values, Student’s two-sided t test. (k) Learned coefficients of transcription factors motifs that gain or lose enrichment in control vs. ARID1A KO in DMSO or fulvestrant (n=15 samples).

Source data

Extended Data Fig. 4 ARID1A loss mediates a basal-like gene expression.

(a) Volcano plot; x-axis is log2 fold change and y-axis represents -log10(P); n=18 samples, statistical by DESeq2. (b) mRNA levels of luminal and basal-like/stemness markers in control and ARID1A KO cells. Error bars=mean ±SEM, n=2 biologically independent samples, center values are means. *P value<0.05, ** P value<0.01, Student’s two-sided t test. (c) Cropped western blot of indicated antibodies. (d) mRNA levels of aforementioned markers in MCF7 upon addition of doxycycline (DOX) knockdown of ARID1A. Error bars=mean ±SEM, n=2 biologically independent samples, center values are means. *P value<0.05, ** P value<0.01, Student’s two-sided t test. Also shown are ARID1A and vinculin levels. (e and f) Cropped western blot with indicated antibodies in BT474 or MDA-MB-361 cells expressing sgNT and two sgRNAs against ARID1A. (g and h) Enrichment of basal-like signatures in BT474 (Charafe breast cancer luminal vs. basal down) or MDA-MB-361 (Huper breast basal vs. luminal up) upon ARID1A KO. (n=6 per cell type, nominal P values and FDR adjusted P values were calculated using GSEA package.) (i and j) Enrichment of basal-like and estrogen response signatures in MCF7 after SMARCB1 or SMARCE1 knockout; n=8 per gene knockout, nominal P values and FDR adjusted P values were calculated using GSEA package. (k) Enrichment of basal-like signatures in ARID1A wild type vs. biallelic loss of ARID1A of patient sample pairs (*, FDR < 0.25;). n=2 for each patient pair, nominal P values and FDR adjusted P values were calculated using GSEA package).

Source data

Extended Data Fig. 5 SWI/SNF binding to chromatin but not complex assembly is lost upon ARID1A loss.

(a) BAF155-2 and BRG1-2 at BAF155/BRG1 binding sites in control and ARID1A KO MCF7 cells (n=1). (b) Box plot representing mean signal across differential BAF155-2 or BRG1-2 after ARID1A KO at BAF155/BRG1 sites. (c) Cropped western blots of co-immunoprecipitation of BRG1 with subunits of the SWI/SNF complex in control and ARID1A KO MCF7. (d) Plot of the fold change between control and ARID1A KO of ATAC-seq sites vs. similar fold change of BAF155/BRG1 sites; n=14838 peaks, R and P values calculated using spearman correlation from ggpubr package in R. (e) BAF155-2 and BRG1-2 at differential accessible sites in control and ARID1A KO MCF7. (f) Box plot representing mean signal across differential BAF155-2 or BRG1-2 after ARID1A KO at lost accessible sites. (g) ChIP-qPCR analysis of ER, FOXA1, and GATA3 in shared loci in control and ARID1A KO cells. (h) ChIP-qPCR analysis of FOS, JUN, and IgG control. (i) ChIP-seq tracks of BRG1 and BAF155 in control and ARID1A KO cells (n=1). For (g) and (h), error bars=mean ±SEM, n=3 biologically independent samples, center values are means. P values, Student’s two-sided t test. For the box plots P-values, Mann-Whitney U test (Wilcoxon rank-sum test, two-sided) and effect size (rosenthal’s coefficient) are shown. The log2FC which is calculated as log2 (mean KO / mean Control) is also indicated (n=6). Box shows 25th, median and 75th percentiles with whiskers extending to ± 1.5 * IQR.

Source data

Extended Data Fig. 6 ARID1A regulates the expression of nuclear hormone receptors in breast cancer.

(a) ECDF plot of log2 fold changes in gene expression between ARID1A knockout and control for genes nearest to the TSS-distal SWI/SNF binding sites at GRHL1, FOXA1, FOS, JUN, GATA3, and ER motifs loci. P values were measured by the Mann-Whitney U test (Wilcoxon rank-sum test, two-sided) and effect size (rosenthal’s coefficients. The log2FC (fold change) values which are calculated as log2 (mean KO / mean Control) are also indicated (n=9). (b) Expression of ER canonical targets in control and ARID1A knockout MCF7 cells. Error bars=mean ±SEM, n=3 biologically independent samples, center values are means. P values, Student’s two-sided t test. (c) Cropped western blot of AR+ TNBC breast cancer cells BT549 and HCC70 with the indicated antibodies. (d) and (e) Cropped western blot with the indicated antibodies of control and ARID1A knockout BT549 or HCC70. (f) and (g) GSEA of androgen response in BT549 and HCC70 after ARID1A knockout; n=8 for each cell line, nominal P values and FDR adjusted P values were calculated using GSEA package.

Source data

Supplementary information

Supplementary Information

Supplementary Note

Reporting Summary

Supplementary Tables

Supplementary Tables 1–8

Source data

Source Data Fig. 1

Unprocessed western blots

Source Data Fig. 3

Unprocessed western blots

Source Data Extended Data Fig. 2

Unprocessed western blots

Source Data Extended Data Fig. 3

Unprocessed western blots

Source Data Extended Data Fig. 4

Unprocessed western blots

Source Data Extended Data Fig. 5

Unprocessed western blots

Source Data Extended Data Fig. 6

Unprocessed western blots

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Xu, G., Chhangawala, S., Cocco, E. et al. ARID1A determines luminal identity and therapeutic response in estrogen-receptor-positive breast cancer. Nat Genet 52, 198–207 (2020). https://doi.org/10.1038/s41588-019-0554-0

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41588-019-0554-0

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