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.

The androgen receptor is a tumor suppressor in estrogen receptor–positive breast cancer

Abstract

The role of the androgen receptor (AR) in estrogen receptor (ER)-α-positive breast cancer is controversial, constraining implementation of AR-directed therapies. Using a diverse, clinically relevant panel of cell-line and patient-derived models, we demonstrate that AR activation, not suppression, exerts potent antitumor activity in multiple disease contexts, including resistance to standard-of-care ER and CDK4/6 inhibitors. Notably, AR agonists combined with standard-of-care agents enhanced therapeutic responses. Mechanistically, agonist activation of AR altered the genomic distribution of ER and essential co-activators (p300, SRC-3), resulting in repression of ER-regulated cell cycle genes and upregulation of AR target genes, including known tumor suppressors. A gene signature of AR activity positively predicted disease survival in multiple clinical ER-positive breast cancer cohorts. These findings provide unambiguous evidence that AR has a tumor suppressor role in ER-positive breast cancer and support AR agonism as the optimal AR-directed treatment strategy, revealing a rational therapeutic opportunity.

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: Androgen inhibits estrogen-stimulated proliferation in clinical breast tissues.
Fig. 2: Activation of AR redistributes ER and p300 on chromatin to repress cell cycle genes.
Fig. 3: AR inhibits growth of patient-derived and cell-line xenograft models of endocrine-resistant ER+ breast cancer.
Fig. 4: AR agonism in the context of CDK4/6 inhibition.
Fig. 5: Clinical significance of AR activity in primary ER+ breast cancers.

Similar content being viewed by others

Nadia Harbeck, Frédérique Penault-Llorca, … Fatima Cardoso

Data availability

All ChIP-seq and RNA-seq data are deposited in the Gene Expression Omnibus under accession number GSE123770. Source data are provided with this paper.

References

  1. Somboonporn, W. & Davis, S. R. Testosterone effects on the breast: implications for testosterone therapy for women. Endocr. Rev. 25, 374–388 (2004).

    Article  CAS  PubMed  Google Scholar 

  2. Hickey, T. E., Robinson, J. L., Carroll, J. S. & Tilley, W. D. Minireview: the androgen receptor in breast tissues: growth inhibitor, tumor suppressor, oncogene? Mol. Endocrinol. 26, 1252–1267 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  3. Goldenberg, I. S. Testosterone propionate therapy in breast cancer. JAMA 188, 1069–1072 (1964).

    Article  CAS  PubMed  Google Scholar 

  4. Kennedy, B. J. Fluoxymesterone therapy in advanced breast cancer. N. Engl. J. Med. 259, 673–675 (1958).

    Article  CAS  PubMed  Google Scholar 

  5. Jordan, V. C. & Brodie, A. M. Development and evolution of therapies targeted to the estrogen receptor for the treatment and prevention of breast cancer. Steroids 72, 7–25 (2007).

    Article  CAS  PubMed  Google Scholar 

  6. Clarke, R., Tyson, J. J. & Dixon, J. M. Endocrine resistance in breast cancer: an overview and update. Mol. Cell. Endocrinol. 418, 220–234 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  7. Lonning, P. E. Additive endocrine therapy for advanced breast cancer – back to the future. Acta Oncol. 48, 1092–1101 (2009).

    Article  CAS  PubMed  Google Scholar 

  8. Ricciardelli, C. et al. The magnitude of androgen receptor positivity in breast cancer is critical for reliable prediction of disease outcome. Clin. Cancer Res. 24, 2328–2341 (2018).

    Article  CAS  PubMed  Google Scholar 

  9. Narayanan, R., Coss, C. C. & Dalton, J. T. Development of selective androgen receptor modulators (SARMs). Mol. Cell. Endocrinol. 465, 134–142 (2018).

    Article  CAS  PubMed  Google Scholar 

  10. Chia, K. et al. Non-canonical AR activity facilitates endocrine resistance in breast cancer. Endocr. Relat. Cancer 26, 251–264 (2019).

    Article  CAS  PubMed  Google Scholar 

  11. Cochrane, D. R. et al. Role of the androgen receptor in breast cancer and preclinical analysis of enzalutamide. Breast Cancer Res. 16, R7 (2014).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  12. D’Amato, N. C. et al. Cooperative dynamics of AR and ER activity in breast cancer. Mol. Cancer Res. 14, 1054–1067 (2016).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  13. De Amicis, F. et al. Androgen receptor overexpression induces tamoxifen resistance in human breast cancer cells. Breast Cancer Res. Treat. 121, 1–11 (2010).

    Article  CAS  PubMed  Google Scholar 

  14. Elston, C. W. & Ellis, I. O. Pathological prognostic factors in breast cancer. I. The value of histological grade in breast cancer: experience from a large study with long-term follow-up. Histopathology 19, 403–410 (1991).

    Article  CAS  PubMed  Google Scholar 

  15. Curtis, C. et al. The genomic and transcriptomic architecture of 2,000 breast tumours reveals novel subgroups. Nature 486, 346–352 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  16. Centenera, M. M. et al. A patient-derived explant (PDE) model of hormone-dependent cancer. Mol. Oncol. 12, 1608–1622 (2018).

    Article  PubMed  PubMed Central  Google Scholar 

  17. Mohammed, H. et al. Progesterone receptor modulates ERα action in breast cancer. Nature 523, 313–317 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  18. Peters, A. A. et al. Androgen receptor inhibits estrogen receptor-α activity and is prognostic in breast cancer. Cancer Res. 69, 6131–6140 (2009).

    Article  CAS  PubMed  Google Scholar 

  19. Santagata, S. et al. Taxonomy of breast cancer based on normal cell phenotype predicts outcome. J. Clin. Invest. 124, 859–870 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  20. Poulin, R., Baker, D. & Labrie, F. Androgens inhibit basal and estrogen-induced cell proliferation in the ZR-75-1 human breast cancer cell line. Breast Cancer Res. Treat. 12, 213–225 (1988).

    Article  CAS  PubMed  Google Scholar 

  21. Birrell, S. N. et al. Androgens induce divergent proliferative responses in human breast cancer cell lines. J. Steroid Biochem. Mol. Biol. 52, 459–467 (1995).

    Article  CAS  PubMed  Google Scholar 

  22. Moore, N. L., Hanson, A. R., Ebrahimie, E., Hickey, T. E. & Tilley, W. D. Anti-proliferative transcriptional effects of medroxyprogesterone acetate in estrogen receptor positive breast cancer cells are predominantly mediated by the progesterone receptor. J. Steroid Biochem. Mol. Biol. 199, 105548 (2020).

    Article  CAS  PubMed  Google Scholar 

  23. Heintzman, N. D. et al. Histone modifications at human enhancers reflect global cell-type-specific gene expression. Nature 459, 108–112 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  24. Drabsch, Y. et al. Mechanism of and requirement for estrogen-regulated MYB expression in estrogen-receptor-positive breast cancer cells. Proc. Natl Acad. Sci. USA 104, 13762–13767 (2007).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  25. Ramsay, R. G. & Gonda, T. J. MYB function in normal and cancer cells. Nat. Rev. Cancer 8, 523–534 (2008).

    Article  CAS  PubMed  Google Scholar 

  26. Buluwela, L. et al. Inhibiting estrogen responses in breast cancer cells using a fusion protein encoding estrogen receptor-α and the transcriptional repressor PLZF. Gene Ther. 12, 452–460 (2005).

    Article  CAS  PubMed  Google Scholar 

  27. Pike, J. et al. Silencing of androgen-regulated genes using a fusion of AR with the PLZF transcriptional repressor. Oncogene 23, 7561–7570 (2004).

    Article  CAS  PubMed  Google Scholar 

  28. Wang, X. et al. Reduced expression of tocopherol-associated protein (TAP/Sec14L2) in human breast cancer. Cancer Invest. 27, 971–977 (2009).

    Article  CAS  PubMed  Google Scholar 

  29. Xiao, W. et al. Suppression of prostate tumor growth by U19, a novel testosterone-regulated apoptosis inducer. Cancer Res. 63, 4698–4704 (2003).

    CAS  PubMed  Google Scholar 

  30. Weinert, B. T. et al. Time-resolved analysis reveals rapid dynamics and broad scope of the CBP/p300 acetylome. Cell 174, 231–244 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  31. Murakami, S., Nagari, A. & Kraus, W. L. Dynamic assembly and activation of estrogen receptor α enhancers through coregulator switching. Genes Dev. 31, 1535–1548 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  32. Lasko, L. M. et al. Discovery of a selective catalytic p300/CBP inhibitor that targets lineage-specific tumours. Nature 550, 128–132 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  33. Yi, P. et al. Structure of a biologically active estrogen receptor-coactivator complex on DNA. Mol. Cell 57, 1047–1058 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  34. Yu, X. et al. Structural insights of transcriptionally active, full-length androgen receptor coactivator complexes. Mol. Cell https://doi.org/10.1016/j.molcel.2020.06.031 (2020).

  35. Hu, D. G. et al. Androgen and estrogen receptors in breast cancer coregulate human UDP-glucuronosyltransferases 2B15 and 2B17. Cancer Res. 76, 5881–5893 (2016).

    Article  CAS  PubMed  Google Scholar 

  36. Tran, C. et al. Development of a second-generation antiandrogen for treatment of advanced prostate cancer. Science 324, 787–790 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  37. DeRose, Y. S. et al. Tumor grafts derived from women with breast cancer authentically reflect tumor pathology, growth, metastasis and disease outcomes. Nat. Med. 17, 1514–1520 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  38. Finlay-Schultz, J. et al. Breast cancer suppression by progesterone receptors is mediated by their modulation of estrogen receptors and RNA polymerase III. Cancer Res. 77, 4934–4946 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  39. Jeselsohn, R., Buchwalter, G., De Angelis, C., Brown, M. & Schiff, R. ESR1 mutations—a mechanism for acquired endocrine resistance in breast cancer. Nat. Rev. Clin. Oncol. 12, 573–583 (2015).

    Article  CAS  Google Scholar 

  40. Lundberg, A. et al. The long-term prognostic and predictive capacity of cyclin D1 gene amplification in 2305 breast tumours. Breast Cancer Res. 21, 34 (2019).

    Article  PubMed  PubMed Central  Google Scholar 

  41. Hicks, D. G. et al. Loss of breast cancer metastasis suppressor 1 protein expression predicts reduced disease-free survival in subsets of breast cancer patients. Clin. Cancer Res. 12, 6702–6708 (2006).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  42. Finlay-Schultz, J. et al. New generation breast cancer cell lines developed from patient-derived xenografts. Breast Cancer Res. 22, 68 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  43. Sflomos, G. et al. A preclinical model for ERα-positive breast cancer points to the epithelial microenvironment as determinant of luminal phenotype and hormone response. Cancer Cell 29, 407–422 (2016).

    Article  CAS  PubMed  Google Scholar 

  44. Turner, N. C. et al. Overall survival with palbociclib and fulvestrant in advanced breast cancer. N. Engl. J. Med. 379, 1926–1936 (2018).

    Article  CAS  PubMed  Google Scholar 

  45. Li, Z. et al. Loss of the FAT1 tumor suppressor promotes resistance to CDK4/6 inhibitors via the Hippo pathway. Cancer Cell 34, 893–905 (2018).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  46. Portman, N. et al. Overcoming CDK4/6 inhibitor resistance in ER-positive breast cancer. Endocr. Relat. Cancer 26, R15–R30 (2019).

    Article  CAS  PubMed  Google Scholar 

  47. Ur-Rehman, S., Gao, Q., Mitsopoulos, C. & Zvelebil, M. ROCK: a resource for integrative breast cancer data analysis. Breast Cancer Res. Treat. 139, 907–921 (2013).

    Article  CAS  PubMed  Google Scholar 

  48. Loi, S. et al. Definition of clinically distinct molecular subtypes in estrogen receptor–positive breast carcinomas through genomic grade. J. Clin. Oncol. 25, 1239–1246 (2007).

    Article  CAS  PubMed  Google Scholar 

  49. Schmidt, S. F., Larsen, B. D., Loft, A. & Mandrup, S. Cofactor squelching: artifact or fact? Bioessays 38, 618–626 (2016).

    Article  PubMed  Google Scholar 

  50. Guertin, M. J., Zhang, X., Coonrod, S. A. & Hager, G. L. Transient estrogen receptor binding and p300 redistribution support a squelching mechanism for estradiol-repressed genes. Mol. Endocrinol. 28, 1522–1533 (2014).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  51. Lydon, J. P. & O’Malley, B. W. Minireview: steroid receptor coactivator-3: a multifarious coregulator in mammary gland metastasis. Endocrinology 152, 19–25 (2011).

    Article  CAS  PubMed  Google Scholar 

  52. Lim, E. et al. Pushing estrogen receptor around in breast cancer. Endocr. Relat. Cancer 23, T227–T241 (2016).

    Article  CAS  PubMed  Google Scholar 

  53. Abdel-Fatah, T. M. A. et al. SPAG5 as a prognostic biomarker and chemotherapy sensitivity predictor in breast cancer: a retrospective, integrated genomic, transcriptomic, and protein analysis. Lancet Oncol. 17, 1004–1018 (2016).

    Article  CAS  PubMed  Google Scholar 

  54. Schindelin, J. et al. FIJI: an open-source platform for biological-image analysis. Nat. Methods 9, 676–682 (2012).

    Article  CAS  PubMed  Google Scholar 

  55. Carroll, J. S., Prall, O. W., Musgrove, E. A. & Sutherland, R. L. A pure estrogen antagonist inhibits cyclin E-Cdk2 activity in MCF-7 breast cancer cells and induces accumulation of p130-E2F4 complexes characteristic of quiescence. J. Biol. Chem. 275, 38221–38229 (2000).

    Article  CAS  PubMed  Google Scholar 

  56. Cazet, A. S. et al. Targeting stromal remodeling and cancer stem cell plasticity overcomes chemoresistance in triple negative breast cancer. Nat. Commun. 9, 2897 (2018).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  57. Vijayaraghavan, S. et al. CDK4/6 and autophagy inhibitors synergistically induce senescence in Rb positive cytoplasmic cyclin E negative cancers. Nat. Commun. 8, 15916 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  58. Law, A. M. K. et al. Andy’s algorithms: new automated digital image analysis pipelines for FIJI. Sci. Rep. 7, 15717 (2017).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  59. Singh, A. A. et al. Optimized ChIP-seq method facilitates transcription factor profiling in human tumors. Life Sci. Alliance 2, e201800115 (2019).

    Article  PubMed  Google Scholar 

  60. Wingett, S. W. & Andrews, S. FastQ Screen: a tool for multi-genome mapping and quality control. F1000Res 7, 1338 (2018).

    Article  PubMed  PubMed Central  Google Scholar 

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

    Article  CAS  PubMed  PubMed Central  Google Scholar 

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

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  63. Li, H. et al. The Sequence Alignment/Map format and SAMtools. Bioinformatics 25, 2078–2079 (2009).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

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

    Article  CAS  PubMed  Google Scholar 

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

    Article  PubMed  PubMed Central  CAS  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. Li, H. & Durbin, R. Fast and accurate short read alignment with Burrows–Wheeler transform. Bioinformatics 25, 1754–1760 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  68. Ross-Innes, C. S. et al. Differential oestrogen receptor binding is associated with clinical outcome in breast cancer. Nature 481, 389–393 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  69. Robinson, J. T. et al. Integrative genomics viewer. Nat. Biotechnol. 29, 24–26 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  70. Ramírez, F. et al. deepTools2: a next generation web server for deep-sequencing data analysis. Nucleic Acids Res. 44, W160–W165 (2016).

    Article  PubMed  PubMed Central  Google Scholar 

  71. 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 

  72. Ji, H. et al. An integrated software system for analyzing ChIP-chip and ChIP-seq data. Nat. Biotechnol. 26, 1293–1300 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

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

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  74. Kal, A. J. et al. Dynamics of gene expression revealed by comparison of serial analysis of gene expression transcript profiles from yeast grown on two different carbon sources. Mol. Biol. Cell 10, 1859–1872 (1999).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  75. Baggerly, K. A., Deng, L., Morris, J. S. & Aldaz, C. M. Differential expression in SAGE: accounting for normal between-library variation. Bioinformatics 19, 1477–1483 (2003).

    Article  CAS  PubMed  Google Scholar 

  76. Conway, T. et al. Xenome—a tool for classifying reads from xenograft samples. Bioinformatics 28, i172–i178 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

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

    Article  CAS  PubMed  Google Scholar 

  78. Li, B. & Dewey, C. N. RSEM: accurate transcript quantification from RNA-Seq data with or without a reference genome. BMC Bioinf. 12, 323 (2011).

    Article  CAS  Google Scholar 

  79. Robinson, M. D. & Oshlack, A. A scaling normalization method for differential expression analysis of RNA-seq data. Genome Biol. 11, R25 (2010).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  80. Law, C. W., Chen, Y., Shi, W. & Smyth, G. K. Voom: precision weights unlock linear model analysis tools for RNA-seq read counts. Genome Biol. 15, R29 (2014).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  81. Smyth, G. K. Linear models and empirical Bayes methods for assessing differential expression in microarray experiments. Stat. Appl. Genet. Mol. Biol. https://doi.org/10.2202/1544-6115.1027 (2004).

  82. 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  PubMed  PubMed Central  CAS  Google Scholar 

  83. McCarthy, D. J., Chen, Y. & Smyth, G. K. Differential expression analysis of multifactor RNA-seq experiments with respect to biological variation. Nucleic Acids Res. 40, 4288–4297 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  84. Ritchie, M. E. et al. limma powers differential expression analyses for RNA-sequencing and microarray studies. Nucleic Acids Res. 43, e47 (2015).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  85. 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  PubMed  PubMed Central  Google Scholar 

  86. Tishchenko, I., Milioli, H. H., Riveros, C. & Moscato, P. Extensive transcriptomic and genomic analysis provides new insights about luminal breast cancers. PLoS ONE 11, e0158259 (2016).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

Download references

Acknowledgements

The authors thank C. Hoffmann, N. Dean and the teams supporting them for consenting and recruiting tissue donors, as well as tissue donors themselves. We thank A. Welm for PDX model provision; J. Robinson, M. Pickering, Z. Kikhtyak and N. Ryan for experimental assistance; and our institutional animal care facilities and genomics cores. This work was supported by grants from the National Health and Medical Research Council of Australia (W.D.T., J.S.C., T.E.H., E.L., L.A.S., E.E., S.N.B., C.P. and A.S.; ID 1084416, ID 1130077; K.M.C.; ID 1093801), the National Breast Cancer Foundation (NBCF; W.D.T.; ID PS-15-041; C.E.C.; ID ECF-17-002) and a Movember & National Breast Cancer Foundation Collaboration Initiative grant (MNBCF-17-012 to W.D.T., J.S.C., E.L., T.E.H. and L.A.S.), Cancer Australia (W.D.T., E.L., T.E.H., J.S.C.; ID 1107170), Cancer Institute of New South Wales (H.H.M.; ID ECF-17-1156) and The Hospital Research Foundation (T.E.H., L.A.S., W.D.T.). T.E.H. held fellowship awards from the US Department of Defense Breast Cancer Research Program (ID BC103923) and the Royal Adelaide Hospital Research Foundation (Adelaide, Australia). W.Z. and S.S. were supported by the Movember Foundation (ID NKI01), Dutch Cancer Society/Alpe D’HuZes (ID 10084) and the Netherlands Organisation for Scientific Research Vidi grant (ID 91716401). C.A.S. and J.F.S. were supported by NIHR01CA140985 and the Breast Cancer Research Foundation. J.S.C. and C.P. are supported by CRUK Funding (grant no. A17197 to J.S.C.). E.L. is supported by an NBCF Endowed Chair. K.M.C. is an NHMRC PhD Scholar. T.E.H. is currently supported by an NBCF Fellowship (IIRS-19-009) and the University of Adelaide.

Author information

Authors and Affiliations

Authors

Contributions

T.E.H., E.L., J.S.C. and W.D.T. conceptualized and designed experiments and acquired funding. Experimental work was conducted by S.J., K.M.C., G.L.L., M.H., S.S., T.M.A.-F., J.F.S. and T.E.H. Data analyses were performed by T.E.H., L.A.S., K.M.C., H.H.M., D.R., G.L.L., S.J., E.E., S.S., T.M.A.-F. and C.P. Reagents and resources were provided by C.E.C., I.O.E., T.M.A.-F., C.P. and S.N.B. Experiments were supervised by T.E.H., E.L., A.S., W.Z., C.P., C.A.S., J.S.C. and W.D.T. The manuscript was written and compiled by T.E.H., L.A.S., J.S.C. and W.D.T., with contributions from other authors. All authors reviewed the manuscript.

Corresponding author

Correspondence to Wayne D. Tilley.

Ethics declarations

Competing interests

S.N.B. is a major shareholder in Havah Therapeutics Pty Ltd. Research support was received from AstraZeneca (W.Z.), Astellas Pharma (W.Z.) and GTx Inc. (W.D.T. and C.P.). All other authors declare no competing interests.

Additional information

Peer review information Javier Carmona was the primary editor on this article and managed its editorial process and peer review in collaboration with the rest of the editorial team.

Publisher’s note Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Extended data

Extended Data Fig. 1 AR protein levels are prognostic in ER-positive breast cancer.

Immunogenic AR protein levels predict breast cancer-specific survival in the Nottingham Tenovus Primary Breast Cancer Series14. Kaplan–Meier plots depicting the association between immunogenic AR positivity (dichotomized by high ≥ 78% and low < 78% nuclear staining based on an optimal cut-point criterion8) and breast cancer-specific survival. a, All cases; b, ER-positive cases (defined by immunogenic ER positivity of >1% as per current ASCO guidelines); c, High-risk (defined by Nottingham Prognostic Index ≥ 3.4) ER-positive cases treated with adjuvant endocrine therapy (Tamoxifen; Tam); d, ER-positive cases not treated with adjuvant Tamoxifen. Survival probabilities were estimated using a univariate Cox proportional hazards model. Differences between survival rates were tested for significance using the log-rank test. Hazard ratios (HR), 95% confidence intervals (95% CI), and p values are provided within each graph. All tests were two-sided.

Extended Data Fig. 2 Anti-proliferative effects of androgen in patient-derived explants (PDEs) of breast tissue exposed to estrogen.

a, Quantification of BrdU immunohistochemistry in ER + malignant patient-derived explants (PDEs; n = 9) treated with estrogen (E2, 10 nM) or E2 plus androgen (DHT, 10 nM). Data represented as mean ± SEM. Data was analyzed using a two-tailed Wilcoxon matched-pairs signed rank test (W = -45.0, r = 0.1149, **p = 0.0039). b, Correlation between immunohistochemical quantification of both BrdU- and Ki67-positive cells shown in (a) and (Fig. 1b). At least 1,000 cells, or all visible epithelial cells, whichever threshold was reached first, were examined in n = 9 independent cases of ER + malignant PDEs, treated with either E2 (left panel) or E2 + DHT (right panel). Data was analyzed by linear regression (trend line) and a two-tailed Pearson’s correlation test (E2 treatment; r = 0.697 [0.061, 0.931]; E2 + DHT treatment; r = 0.912 [0.627, 0.982]. P values are provided within each graph. c,d, RNA-seq data derived from PDEs of (c) normal breast tissue (n = 17 cases) and (d) ER + invasive breast cancer (n = 8 cases) treated with E2 (10 nM) or E2 + DHT (10 nM each). Heat maps depict the top 200 differentially expressed genes. e, Gene set enrichment plots for the Hallmark G2M Checkpoint gene set derived from RNA-seq data representing normal breast tissues (left panel) or ER + breast cancers (right panel).

Extended Data Fig. 3 AR-mediated changes in ER binding at ER and AR target genes.

a, Motif analysis of E2-stimulated ER binding sites associated with Fig. 2a,b, which are altered in the presence of androgen (DHT) in ZR-75-1 cells. GRHL (grainy-head-like motif); ERE (estrogen response element); ARE (androgen response element). b, Cropped immunoblot showing effective siRNA-mediated knock-down of AR (left panel) and ER (right panel) in ZR-75-1 cells treated with estrogen (E; 10 nM) alone or in combination with androgen (E + D; 10 nM each) for 4 hours. β-Actin was used as a loading control. Blots are representative of three experiments corresponding to independent passages of cells. c, Left panels: Genome browser images showing ER and AR ChIP-seq signals (associated with Fig. 2b) at E2-stimulated ER binding sites (BCL2, PGR) that are reduced in the presence of androgen (DHT, 10 nM). Right panels: Corresponding ER ChIP-PCR for BCL2 (upper panel) and PGR (lower panel) following AR knockdown and subsequent treatment with estrogen (E2, 10 nM) alone or in combination with androgen (E2 + DHT, 10 nM each) for 4 hours in ZR-75-1 cells. Data was analyzed using a one-sided unpaired t-test. Upper panel test details: t = 5.320, 0.2325, and 0.2274, for siCon (p = 0.0030), siAR (1) (p = 0.4138), and siAR (2) (p = 0.4915), respectively. Lower panel test details: t = 2.074, 0.2437, and 0.3767 for siCon (p = 0.0534), siAR (1) (p = 0.4097), and siAR (2) (p = 0.3628), respectively; df = 4 for each test. d, Left panels: Genome browser images showing AR and ER ChIP-seq signals (associated with Fig. 2b) at DHT-stimulated AR binding sites (ZBTB16, SEC14L2). Right panels: Corresponding AR ChIP-qPCR for ZBTB16 (upper panel) and SEC14L2 (lower panel) following ER knockdown and subsequent treatment with estrogen (E2, 10 nM) alone or in combination with androgen (E2 + DHT, 10 nM each) for 4 hours in ZR-75-1 cells. Each genome track in (c) and (d) depicts the average read density of all ChIP-seq replicates for the designated receptor and treatment condition. Data was analyzed using a one-sided unpaired t-test. Test details: Upper-left panel; t = 4.587, 1.102, and 0.6163, for siCon (p = 0.0051), siER (1) (p = 0.1662), and siER (2) (p = 0.2885), respectively. Upper right panel; t = 2.596, 4.443, and 6.906, for siCon (p = 0.0301), siAR (1) (p = 0.0057), and siAR (2) (p = 0.0012), respectively. Lower-left panel; t = 4.301, 1.221, and 1.198, for siCon (p = 0.0063), siER (1) (p = 0.1446), and siER (2) (p = 0.1485), respectively. Lower-right panel; t = 2.321, 2.504, and 1.678, for siCon (p = 0.0405), siAR (1) (p = 0.0332), and siAR (2) (p = 0.0843), respectively; df = 4 for each test. ChIP-PCR data is represented as mean ± SEM of three independent passages of cells. Asterisks denote statistical significance; NS = not significant. e, Average read density plots for ER, AR, and H3K27ac at ER binding sites within 100 kb of genes down-regulated in E2 + DHT versus E2 treatment arms in ZR-75-1 cells. f, Heatmap of RT-qPCR analysis of ER target gene expression in ZR-75-1 and T-47D cells treated for 6 hours with Vehicle (Veh), DHT (10 nM), E2 (10 nM) or E2 + DHT (ED, 10 nM each). g, Average read density plots for AR, ER, and H3K27ac at AR binding sites within 100 kb of genes up-regulated by E2 + DHT versus E2 in ZR-75-1 cells. h, Heatmap of RT-qPCR analysis of AR target gene expression in ZR-75-1 and T-47D cells treated for 6 hours with Vehicle (Veh), DHT (10 nM), E2 (10 nM) or E2 + DHT (ED, 10 nM each). Heatmap data shown in (f) and (h) represents the average normalized gene expression of four independent passages of cells.

Extended Data Fig. 4 AR-mediated modification of the ER cistrome requires an agonist ligand and AR DNA binding capacity.

a, Genome browser image of ER and AR chromatin binding events after treatment for 4 hours with either estrogen (E2, 10 nM) or estrogen plus androgen (E2 + DHT, 10 nM each) in ZR-75-1 cells at the MYB gene locus based on ChIP-seq data from Fig. 2. Each genome track depicts the average read density of all ChIP-seq replicates for the designated receptor and treatment condition. Enh., Enhancer. Tracks are scaled to facilitate visualization of the Intron 1 locus. b, ZR-75-1 in vitro ChIP-qPCR validation data for ER, H3K27ac and AR at three loci associated with the MYB gene locus as depicted in (a). Bar graphs depict mean ± SEM of 3 independent passages of cells. Data was analyzed using a one-sided unpaired t-test. ER ChIP; t = 1.345, 1.17, and 0.5662 for Enh. 1 (p = 0.1249), Intron 1 (p < 0.0001), and Enh. 2 (p = 0.3101), respectively. H3K27ac ChIP; t = 0.3320, 2.310, and 0.5664 for Enh. 1 (p = 0.3783), Intron 1 (p = 0.0410), and Enh. 2 (p = 0.3007), respectively. AR ChIP; t = 3.555, 1.480, and 3.993 for for Enh. 1 (p = 0.0118), Intron 1 (p = 0.1064), and Enh. 2 (p = 0.0085), respectively; df = 4 for each test. Grey asterisks denote statistical significance. c, Average read density plot of ER ChIP-seq enrichment at ER binding sites proximal (<100 kb) to genes down-regulated in E2 + DHT versus E2 treatment arms in ZR-75-1 cells, as per Extended Data Fig. 3e. Data shows ER enrichment in ZR-75-1 cells following in vitro treatment with Vehicle (Veh), estrogen (E2, 1 nM), an AR agonist (DHT, 1 nM), an AR antagonist (Bicalutamide; Bic 1 µM), alone or with indicated combinations. The data independently recapitulates the DHT-induced reduction in the E2-stimulated ER signal observed in Extended Data Fig. 3c,e at 10 nM doses of hormones, and using an independent ER antibody; Millipore 06-935. Treatment with an AR antagonist does not induce loss of ER at these loci. d, Genome browser images of ER binding sites in ZR-75-1 cells at ER target genes (BCL2, PGR), showing reduced binding following treatment with an AR agonist (DHT) but not an AR antagonist (Bicalutamide). Each genome track depicts the average read density of all ChIP-seq replicates for the designated receptor and treatment condition. e, Growth curve showing that estrogen (E2, 1 nM)-stimulated in vitro growth of ZR-75-1 cells is inhibited by an AR agonist (DHT, 1 nM) but not an AR antagonist (Bicalutamide, 1 µM). Data represents mean ± SD of three replicate cell culture wells per condition, and is a representative of two independent experiments. f-i, Experiments in MCF7 breast cancer cells transformed to overexpress a constitutively active, truncated AR (1-707 aa; AR) or a mutant AR that is unable to bind DNA (ARmut). AR expression is induced upon treatment with doxycycline (Dox). f Cropped immunoblot showing Dox-induced expression of truncated AR in cells treated with Vehicle or estrogen (E2, 1 nM), and resultant ER expression. β-Actin was used as a loading control. Blots are representative of three experiments corresponding to independent passages of cells. g Heatmap of RT-qPCR data for two ER target genes (MYB, CCND1) and two AR target genes (SEC14L2; ZBTB16) in cells treated with Vehicle (Veh) or estrogen (E2, 1 nM). Data represents the average normalized gene expression from three independent passages of cells replicates. h ER ChIP-qPCR at loci associated with ER target genes (MYB, CCND1) (top panels) and an AR target gene (ZBTB16). AR ChIP-qPCR is also shown for the AR target gene. Data was analyzed by a two-way ANOVA followed by Tukey’s multiple comparisons test. Upper-left panel; F = 125.9, 11.75, and 3.143 for presence of E2 (p < 0.0001), AR status (p = 0.0015), and their interaction (p = 0.0799), respectively. Asterisks denote statistical significance, where **p = 0.0060, NSp = 0.9998. Upper-right panel; F = 81.79, 13.85, and 1.565 for presence of E2 (p < 0.0001), AR status (p = 0.0008), and their interaction (p = 0.2489), respectively; where **p = 0.0094, NSp > 0.9999. Lower-left panel; F = 140.8, 284.6, and 94.59 for presence of E2 (p < 0.0001), AR status (p < 0.0001), and their interaction (p < 0.0001), respectively; where ****p < 0.0001. Lower-right panel; F = 0.1975, 147.1, and 0.09898 for presence of E2 (p = 0.6647), AR status (p < 0.0001), and their interaction (p = 0.9065), respectively; where ****p < 0.0001. df = 4 for each test. Bar graphs depict mean ± SEM of 3 independent passages of cells. NS = Not significant; **p < 0.01; ****p < 0.0001. i, Growth curve showing that induction of constitutively-active AR inhibits E2-stimulated growth but induction of the constitutively-active ARmut does not. Data represents mean ± SD of three replicate cell culture wells per condition, and is a representative of two independent experiments.

Extended Data Fig. 5 Activation of AR relocates p300 from ER to AR target genes.

a, Replicate data for p300 ChIP-seq in vitro experiments in ZR-75-1 cells associated with Fig. 2e. Venn diagrams show the overlap of three independent experiments representing consecutive passages of cells treated with vehicle (Veh), estrogen (E2; 10 nM) or estrogen plus androgen (E2 + DHT; 10 nM each). Peaks present in at least 2 of 3 replicates were used to generate a consensus cistrome, indicated below the Venn diagrams, for further comparative analyses. b, Overlap of consensus p300 cistromes under E2 or E2 + DHT hormone treatments, after subtracting peaks present under basal (Veh) conditions to generate as set of hormone-regulated peaks. c, Consensus p300 ChIP-seq data from (b), associated with Fig. 2e, showing average read density plots (top panels) and heatmaps (bottom panels), illustrating changes in hormone-regulated p300 chromatin binding sites following treatment with E2 + DHT. d, Average read density plots for p300 binding at ER binding sites (ERBS) proximal (<100 kb) to genes down-regulated by androgen under estrogenic conditions (left panel), and AR binding sites (ARBS) proximal (<100 kb) to genes up-regulated under the same conditions (right panel). e, Example genome browser images showing averaged p300 ChIP-seq signals at binding sites associated with an ER target gene (PGR; left panel) and AR target gene (SEC14L2; right panel) in ZR-75-1 cells. Data represents the average signal of three replicates. f, p300 ChIP-qPCR at enhancers of ER-regulated cell cycle genes (as per Fig. 2g) in T-47D cells treated in vitro under designated hormone conditions. Data was analyzed by a two-way ANOVA (F = 93.45, 80.19, and 14.78 for hormone treatment (p < 0.0001), test site (p < 0.0001), and their interaction (p < 0.0001), respectively; df = 30). Data represented as mean ± SEM of 3 independent passages of cells. Post-hoc analyzes were performed using Tukey’s multiple comparisons test, where asterisks denote statistical significance; *p = 0.0275; *** p = 0.0040; ****p < 0.0001. g, Heatmap of RT-qPCR data for genes associated with Fig. 2g assessed in ZR-75-1 and T-47D cells treated in vitro with estrogen (E2, 10 nM) alone or in the presence of androgen (E2 + DHT, 10 nM each). Data represents the average normalized gene expression of four experiments conducted on independent passages of cells.

Extended Data Fig. 6 Activation of AR relocates SRC-3 from ER to AR target genes; correlating with p300 chromatin localization.

a, Replicate data for SRC-3 ChIP-seq experiments. Venn diagrams show the overlap of three independent experiments representing consecutive passages of ZR-75-1 cells treated with vehicle (Veh), estrogen (E2; 10 nM) or estrogen plus androgen (E2 + DHT; 10 nM each). Peaks present in at least 2 of 3 replicates were used to generate a consensus cistrome, indicated below the Venn diagrams, for further comparative analyses. b, Overlap of consensus SRC-3 cistromes under E2 or E2 + DHT hormone treatments. c, Two-factor log ratio (M) plot displaying DHT-induced changes in E2-stimulated p300 and SRC-3 enrichment at consensus p300 binding sites in ZR-75-1 cells (as per Fig. 2e). Point color denotes p300 consensus peak occupancy; hormone-responsive E2-unique (pink), hormone-responsive E2 + DHT-unique (purple) and basal (black; plotted at rear). Point co-ordinates denote enrichment scores at consensus binding sites, derived from an average of three replicates. Example p300 binding sites associated with ER (yellow-orange) and AR (orange-red) target genes are highlighted. d, SRC-3 ChIP-seq data showing read density plots (left panel) and genome browser images (right panels), illustrating E2 + DHT-induced changes in SRC-3 enrichment at ER binding sites (ERBS) proximal (<100 kb) to genes down-regulated by androgen under estrogenic conditions. Data represents an average of three replicates. e, Average read density plots for SRC-3 binding at AR binding sites (ARBS) proximal (<100 kb) to genes up-regulated by androgen under estrogenic conditions (left panel). Right panels: Genome browser images displaying averaged SRC-3 ChIP-seq signals at example ARBS; SEC14L2 and ZBTB16. Data represents an average of three replicates. f, SRC-3 ChIP-PCR at enhancers of ER-regulated cell cycle genes (as per Fig. 2g; Extended Data Fig. 5f) in ZR-75-1 cells (upper panel) and T-47D cells (lower panel), treated in vitro under designated hormone conditions. Data was analyzed by a two-way ANOVA. Upper panel F = 235.4, 61.41, and 7.338 for hormone treatment (p < 0.0001), test site (p < 0.0001), and their interaction (p < 0.0001), respectively; df =30. Lower panel F = 33.84, 29.62, and 5.343 for hormone treatment (p < 0.0001), test site (p < 0.0001), and their interaction (p = 0.0003), respectively; df = 30. ChIP-PCR data is represented as mean ± SEM of 3 independent passages of cells. Post-hoc analyzes were performed using Tukey’s multiple comparisons test, where asterisks denote statistical significance; *p < 0.05; ** p < 0.01; *** p < 0.001; ****p < 0.0001. Exact p-values for upper panel: MYC p = 0.0051 for E2 vs. E2 + DHT; BCL2 p = 0.0066 for E2 vs. E2 + DHT; FOXM1 p = 0.296 for E2 vs. E2 + DHT; and p < 0.0001 for all other comparisons indicated by ****. Lower panel: MYB p = 0.0002 for E2 vs. E2 + DHT; CCND1 p = 0.0371 for E2 vs. E2 + DHT; FOXM1 p = 0.0198 for E2 vs. E2 + DHT; p < 0.0001 for all other comparisons that are indicated by ****.

Extended Data Fig. 7 AR agonism sustains growth inhibition in a ZR-75-1 xenograft tumor model.

a, Growth curves for ZR-75-1 xenograft tumors treated with E2 alone (n = 6) or in combination with AR agonists (DHT or Enobosarm (SARM); n = 6 and 7 tumors, respectively) or an AR antagonist (Enzalutamide; n = 6). Data represented as mean ± SEM. b, Representative IHC for ER, PR (encoded by ER target gene PGR), AR, and SEC14L2 (encoded by AR target gene SEC14L2) for ZR-75-1 xenograft tumors harvested 5 days post-treatment. Scale bars represent 100 µm. c, Quantification of Ki67-positivity from n = 4 (E2 + Veh, E2 + SARM and E2 + Enz) or n = 5 (E2 + DHT) ZR-75-1 xenograft tumors harvested 5 days post-treatment. Boxes in the graph show minimum and maximum (bottom and top lines, respectively) and mean (line within the boxes). Data was analyzed using a two-tailed, unpaired t-test (t = 17.36, 11.37, and 0.7941; df = 7, 6, and 7; for E2 + Veh vs. E2 + DHT (p < 0.0001), E2 + SARM (p < 0.0001), and E2 + Enza (p = 0.4574), respectively). Asterisks denote statistical significance; ****p < 0.0001, d, Ki67 IHC images in each treatment arm associated with (c). Scale bars represent 100 µm. Images are representative of n = 4 (E2 + Veh, E2 + SARM and E2 + Enz) or n = 5 (E2 + DHT) independent tumors. e, RNA-seq heatmap of all differentially expressed genes (FDR < 0.05) in 5 day-treated ZR-75-1 tumors demonstrates clustering of AR agonists (DHT and SARM) away from Vehicle (Veh) and the AR antagonist Enzalutamide (Enz). f, Independent RT-qPCR validation of select ER and AR target genes. Data represents an average normalized gene expression of 4 (E2 + Veh, E2 + Enz, E2 + SARM) or 5 (E2 + DHT) replicate tumors. NS, not significant.

Extended Data Fig. 8 ZR-75-1 xenograft sequencing data representing tumors harvested 5 days post-AR agonist treatment in vivo.

a, Read density plots (upper panel) and heatmaps (lower panel) of ER ChIP-seq data from ZR-75-1 xenograft tumors, illustrating the loss of ER binding upon activation of AR by agonists (SARM, DHT). Data is presented as an average of 4 (E2 + Veh) or 5 (E2 + SARM, E2 + DHT) replicate tumors. b, Two-factor log-ratio (M) plot showing SARM-induced changes in ER and AR enrichment at consensus ER (yellow-orange), AR (blue), or Shared (that is AR and ER co-occupied; grey-brown; plotted at rear) binding sites seen across the Vehicle, SARM, and DHT treatment arms. Example binding sites near ER target genes and AR target genes are highlighted in pink and purple, respectively. Enrichment scores were calculated from an average of 4 (ER, E2) or 5 (ER, E2 + SARM, AR E2, E2 + SARM) tumors. c, Read density plots (upper panel) and heatmap (lower panels) depictions of ER ChIP-seq presented in (a,b). d, Genome browser images of ER and AR binding in ZR-75-1 xenograft tumors treated with E2 or E2 + SARM at loci associated with ER target genes (MYB, CCND1; left panels) and AR target genes (SEC14L2, ZBTB16; right panels). Data represents an average of 4-5 replicates as described in (b). e, Normalized enrichment scores for signature gene sets correlated with E2 or E2 + SARM treatments, derived using RNA-seq data associated with Extended Data Fig. 7e.

Extended Data Fig. 9 AR modulation in patient-derived xenograft (PDX) models of ER+ endocrine-resistant breast cancer.

a, IHC showing baseline expression of ER, AR, progesterone receptor (PR), Ki67, and human cytokeratin 8/18 (CK8/18) in HCI-005 and GAR15-13 PDX models. Scale bars represent 50 µm. b, Growth curves showing inhibition of E2-stimulated PDX tumor growth (UCD4; UCD65) by co-treatment with DHT at time of tumor inoculation. Tumor volumes at ethical end-point were analyzed using a two-tailed, unpaired t-test, for UCD4; vehicle (n = 6) vs. DHT (n = 6 tumors each; t = 4.213, df = 10, p = 0.0018) and UCD65; vehicle (n = 12) vs. DHT (n = 5; t = 5.572, df = 15, p < 0.0001). c, AR agonist Enobosarm (SARM) inhibits in vivo growth of the UCD65 PDX. Tumors were grown in the presence of E2 to ~100 mm3 and subsequently treated with E2 + SARM (n = 10) or E2 n = 7), indicated by the red arrow. Graph on right shows quantification of tumor mass at end-point for E2 vs. E2 + SARM (n = 9 tumors each), analyzed by a two-tailed, unpaired t-test (t = 2.635, df = 16, p = 0.018). HER2, human epidermal growth factor receptor 2 (where ‘HER2-’ indicates lack of overexpression or gene amplification). d, Quantification of Ki67 positivity in PDX model HCI-005, from Fig. 3b, harvested 5 days post-treatment (t = 32.43 and 14.30 for E2 vs. E2 + DHT and E2 + SARM, respectively; df = 6 and p < 0.0001 for each test). At least 1,000 cells, or all visible epithelial cells, whichever threshold was reached first, were examined over n = 4 independent tumors per condition. e, HCI-005 PDX ER ChIP-seq data depicting loss of E2-stimulated ER chromatin binding at the MYC gene locus following 5 days treatment with AR agonists; Enobosarm (SARM) and DHT. f, HCI-005 PDX ER ChIP-seq data from Fig. 3b, showing average read density plots (top panels) and heatmaps (bottom panels) to illustrate genome-wide changes in E2-stimulated ER chromatin binding sites following 5 days treatment with AR agonists; Enobosarm (SARM) or DHT. g, Growth curve representing the UCD4 PDX cell line42 treated with estrogen (E2, 1 nM) alone or in combination with an AR agonist (DHT, 1 nM), AR antagonist (Enzalutamide; Enz, 1 µM) or the AR agonist and antagonist combined (Combo). Graph depicts mean± SEM of twelve replicate cell culture wells. h, Quantification of Ki67 positivity and i, corresponding representative images of Ki67 IHC in PDX model GAR15-13 from Fig. 3f, harvested 5 days post-treatment, with 5 tumors analyzed per condition by a two-tailed, unpaired t-test. Test details: Veh vs. DHT (t = 9.231, df = 8, p < 0.0001); Veh vs. SARM (t = 4.361, df = 8, p = 0.0024); Veh vs. Enz (t = 0.7958, df = 8, p = 0.4491). Scale bars in (i) represent 50 µm. j, Growth curve representing GAR15-13 PDX tumors treated with either Vehicle (n = 6), AR agonist (DHT, n = 4), or AR antagonist (Enzalutamide; Enza, n = 5) alone or in combination (DHT + Enza, n = 5). k, Heatmap of RT-qPCR data representing the average normalized gene expression from GAR15-13 and HCI-005 PDX tumors harvested 5 days post-treatment (GAR15-13: Veh, n = 6; DHT, n = 8; Enza, n = 8; DHT + Enza, n = 8) and (HCI-005: Veh, n = 5; DHT, n = 6; Enza, n = 7; DHT + Enza, n = 5). All graphed data depicts mean ± SEM. Asterisks denote statistical significance; *p < 0.05; **p < 0.01; ****p < 0.0001.

Extended Data Fig. 10 Data associated with the AR signature genes in Fig. 5.

a, Hallmark gene sets enriched in the AR signature. The signature was separated into up-regulated (n = 59) and down-regulated (n = 83) genes and enriched gene sets (FDR < 0.05) identified by GSEA. b, Average read density plots for H3K27ac binding in HCI-005 PDX tumors, found at ER chromatin binding sites (left panel) and AR chromatin binding sites (right panel) proximal (<100 kb) to AR signature genes, associated with Fig. 5c. Data illustrates AR agonist (SARM, DHT)-dependent loss of H3K27ac binding at ER binding sites (ERBS) associated with down-regulated AR signature genes, and AR agonist-dependent gain of H3K27ac binding at AR binding sites (ARBS) associated with up-regulated AR signature genes. c, Genome browser image of H3K27ac binding in HCI-005 PDX tumors at a locus associated with an ER target gene down-regulated by AR agonists (PGR; left panel), and an AR target gene up-regulated by AR agonists (ZBTB16; right panel) as found in the AR signature. Data represents an average of three (E2 + DHT) or four (E2, E2 + SARM) replicate tumors.

Supplementary information

Supplementary Information

Supplementary Figs. 1–4, unprocessed western blots related to Extended Data and Supplementary Figs. and raw data exemplifying FACS analysis.

Reporting Summary

Supplementary Table 1

Data associated with Fig. 1 and Extended Data Figs. 1 and 2, collated over seven tabs, including the regression model analysis for 20 years of breast cancer–specific survival for both all and ER-positive breast cancer cases, clinical information associated with patient-derived explants used for Ki67 quantification and RNA-seq, and clinical cohort data associated with breast tissue microarray.

Supplementary Table 2

Data associated with Fig. 2, providing a list of genes differentially expressed by dihydrotestosterone (DHT) in ZR-75-1 cells, as determined by RNA-seq, and MACS2 peak calling statistics from AR and ER ChIP-seq with ZR-75-1 tumors.

Supplementary Table 3

Data associated with development of the AR signature, providing a list of genes differentially expressed by DHT in an E2-dependent ZR-75-1 xenograft model and two PDX models (E2-dependent HCI-005 and E2-independent GAR15-13). A list of genes found in the AR gene signature is provided alongside results of the SigCheck analysis.

Supplementary Table 4

Primer sequences used in ChIP–PCR and RT–PCR experiments.

Source data

Source Data Fig. 4

Unprocessed western blots.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Hickey, T.E., Selth, L.A., Chia, K.M. et al. The androgen receptor is a tumor suppressor in estrogen receptor–positive breast cancer. Nat Med 27, 310–320 (2021). https://doi.org/10.1038/s41591-020-01168-7

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41591-020-01168-7

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