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:

Targeting androgen receptor phase separation to overcome antiandrogen resistance

Abstract

Patients with castration-resistant prostate cancer inevitably acquire resistance to antiandrogen therapies in part because of androgen receptor (AR) mutations or splice variants enabling restored AR signaling. Here we show that ligand-activated AR can form transcriptionally active condensates. Both structured and unstructured regions of AR contribute to the effective phase separation of AR and disordered N-terminal domain plays a predominant role. AR liquid–liquid phase separation behaviors faithfully report transcriptional activity and antiandrogen efficacy. Antiandrogens can promote phase separation and transcriptional activity of AR-resistant mutants in a ligand-independent manner. We conducted a phase-separation-based phenotypic screen and identified ET516 that specifically disrupts AR condensates, effectively suppresses AR transcriptional activity and inhibits the proliferation and tumor growth of prostate cancer cells expressing AR-resistant mutants. Our results demonstrate liquid–liquid phase separation as an emerging mechanism underlying drug resistance and show that targeting phase separation may provide a feasible approach for drug discovery.

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

Access options

Buy this article

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

Fig. 1: Ligand-stimulated AR translocated to nucleus and formed transcriptionally active condensates.
Fig. 2: All AR domains contribute to effective condensate formation while NTD plays a predominant role.
Fig. 3: AR phase separation mediates the mutation-induced resistance to antiandrogens.
Fig. 4: Identification of ET516 as a potential AR inhibitor from LLPS-based screen.
Fig. 5: ET516 specifically reduce AR transcriptional activity and inhibits growth of CRPC.

Similar content being viewed by others

Data availability

Data supporting the findings of this study are presented within the article, the accompanying source data files and the Supplementary Information. RNA sequencing data have been submitted to the Gene Expression Omnibus database under accession GSE185223. Three-dimensional rendered graphics have been uploaded to Figshare (https://doi.org/10.6084/m9.figshare.20484396). Source data are provided with this paper.

References

  1. Siegel, R. L., Miller, K. D. & Jemal, A. Cancer statistics, 2019. CA Cancer J. Clin. 69, 7–34 (2019).

    PubMed  Google Scholar 

  2. Schatten, H. in Molecular & Diagnostic Imaging in Prostate Cancer. Advances in Experimental Medicine and Biology Vol. 1126 (ed. Schatten, H.) 117–133 (Springer, 2019).

  3. Harris, W. P., Mostaghel, E. A., Nelson, P. S. & Montgomery, B. Androgen deprivation therapy: progress in understanding mechanisms of resistance and optimizing androgen depletion. Nat. Clin. Pract. Urol. 6, 76–85 (2009).

    CAS  PubMed  PubMed Central  Google Scholar 

  4. Watson, P. A., Arora, V. K. & Sawyers, C. L. Emerging mechanisms of resistance to androgen receptor inhibitors in prostate cancer. Nat. Rev. Cancer 15, 701–711 (2015).

    CAS  PubMed  PubMed Central  Google Scholar 

  5. Shen, M. M. & Abate-Shen, C. Molecular genetics of prostate cancer: new prospects for old challenges. Genes Dev. 24, 1967–2000 (2010).

    CAS  PubMed  PubMed Central  Google Scholar 

  6. Chen, C. D. et al. Molecular determinants of resistance to antiandrogen therapy. Nat. Med. 10, 33–39 (2004).

    PubMed  Google Scholar 

  7. Scher, H. I. & Sawyers, C. L. Biology of progressive, castration-resistant prostate cancer: directed therapies targeting the androgen-receptor signaling axis. J. Clin. Oncol. 23, 8253–8261 (2005).

    CAS  PubMed  Google Scholar 

  8. Grasso, C. S. et al. The mutational landscape of lethal castration-resistant prostate cancer. Nature 487, 239–243 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  9. Jenster, G. et al. Domains of the human androgen receptor involved in steroid binding, transcriptional activation, and subcellular localization. Mol. Endocrinol. 5, 1396–1404 (1991).

    CAS  PubMed  Google Scholar 

  10. Rundlett, S. E., Wu, X. P. & Miesfeld, R. L. Functional characterizations of the androgen receptor confirm that the molecular basis of androgen action is transcriptional regulation. Mol. Endocrinol. 4, 708–714 (1990).

    CAS  PubMed  Google Scholar 

  11. Simental, J. A., Sar, M., Lane, M. V., French, F. S. & Wilson, E. M. Transcriptional activation and nuclear targeting signals of the human androgen receptor. J. Biol. Chem. 266, 510–518 (1991).

    CAS  PubMed  Google Scholar 

  12. De Mol, E. et al. EPI-001, A compound active against castration-resistant prostate cancer, targets transactivation unit 5 of the androgen receptor. ACS Chem. Biol. 11, 2499–2505 (2016).

    PubMed  PubMed Central  Google Scholar 

  13. De Mol, E. et al. Regulation of androgen receptor activity by transient interactions of its transactivation domain with general transcription regulators. Structure 26, 145–152.e3 (2018).

    CAS  PubMed  Google Scholar 

  14. Reid, J., Kelly, S. M., Watt, K., Price, N. C. & McEwan, I. J. Conformational analysis of the androgen receptor amino-terminal domain involved in transactivation. J. Biol. Chem. 277, 20079–20086 (2002).

    CAS  PubMed  Google Scholar 

  15. Yoshida, T. et al. Antiandrogen bicalutamide promotes tumor growth in a novel androgen-dependent prostate cancer xenograft model derived from a bicalutamide-treated patient. Cancer Res. 65, 9611–9616 (2005).

    CAS  PubMed  Google Scholar 

  16. Li, Y. et al. Androgen receptor splice variants mediate enzalutamide resistance in castration-resistant prostate cancer cell lines. Cancer Res. 73, 483–489 (2013).

    CAS  PubMed  Google Scholar 

  17. Banani, S. F., Lee, H. O., Hyman, A. A. & Rosen, M. K. Biomolecular condensates: organizers of cellular biochemistry. Nat. Rev. Mol. Cell Biol. 18, 285–298 (2017).

    CAS  PubMed  PubMed Central  Google Scholar 

  18. Hnisz, D., Shrinivas, K., Young, R. A., Chakraborty, A. K. & Sharp, P. A. A phase separation model for transcriptional control. Cell 169, 13–23 (2017).

    CAS  PubMed  PubMed Central  Google Scholar 

  19. Boija, A. et al. Transcription factors activate genes through the phase-separation capacity of their activation domains. Cell 175, 1842–1855.e16 (2018).

    CAS  PubMed  Google Scholar 

  20. Alberti, S. & Dormann, D. Liquid–liquid phase separation in disease. Annu. Rev. Genet. 53, 171–194 (2019).

    CAS  PubMed  Google Scholar 

  21. Boija, A., Klein, I. A. & Young, R. A. Biomolecular condensates and cancer. Cancer Cell 39, 174–192 (2021).

    CAS  PubMed  PubMed Central  Google Scholar 

  22. Klein, I. A. et al. Partitioning of cancer therapeutics in nuclear condensates. Science 368, 1386–1392 (2020).

    CAS  PubMed  PubMed Central  Google Scholar 

  23. Bouchard, J. J. et al. Cancer mutations of the tumor suppressor SPOP disrupt the formation of active, phase-separated compartments. Mol. Cell 72, 19–36.e8 (2018).

    CAS  PubMed  PubMed Central  Google Scholar 

  24. Zhang, F. et al. Dynamic phase separation of the androgen receptor and its coactivators to regulate gene expression. Preprint at bioRxiv https://doi.org/10.1101/2021.03.27.437301 (2021).

  25. Ahmed, J., Meszaros, A., Lazar, T. & Tompa, P. DNA-binding domain as the minimal region driving RNA-dependent liquid–liquid phase separation of androgen receptor. Protein Sci. 30, 1380–1392 (2021).

    CAS  PubMed  PubMed Central  Google Scholar 

  26. Van Royen, M. E. et al. Compartmentalization of androgen receptor protein–protein interactions in living cells. J. Cell Biol. 177, 63–72 (2007).

    PubMed  PubMed Central  Google Scholar 

  27. Alberti, S., Gladfelter, A. & Mittag, T. Considerations and challenges in studying liquid–liquid phase separation and biomolecular condensates. Cell 176, 419–434 (2019).

    CAS  PubMed  PubMed Central  Google Scholar 

  28. Guo, Z. et al. A novel androgen receptor splice variant is up-regulated during prostate cancer progression and promotes androgen depletion-resistant growth. Cancer Res. 69, 2305–2313 (2009).

    CAS  PubMed  PubMed Central  Google Scholar 

  29. Hu, R. et al. Ligand-independent androgen receptor variants derived from splicing of cryptic exons signify hormone-refractory prostate cancer. Cancer Res. 69, 16–22 (2009).

    CAS  PubMed  PubMed Central  Google Scholar 

  30. Sun, S. et al. Castration resistance in human prostate cancer is conferred by a frequently occurring androgen receptor splice variant. J. Clin. Invest. 120, 2715–2730 (2010).

    CAS  PubMed  PubMed Central  Google Scholar 

  31. Shin, Y. et al. Spatiotemporal control of intracellular phase transitions using light-activated optodroplets. Cell 168, 159–171.e14 (2017).

    CAS  PubMed  Google Scholar 

  32. Brüggenwirth, H. T. et al. Substitution of Ala564 in the first zinc cluster of the deoxyribonucleic acid (DNA)-binding domain of the androgen receptor by Asp, Asn, or Leu exerts differential effects on DNA binding. Endocrinology 139, 103–110 (1998).

    PubMed  Google Scholar 

  33. He, B., Kemppainen, J. A. & Wilson, E. M. FXXLF and WXXLF sequences mediate the NH2-terminal interaction with the ligand binding domain of the androgen receptor. J. Biol. Chem. 275, 22986–22994 (2000).

    CAS  PubMed  Google Scholar 

  34. Hara, T. et al. Novel mutations of androgen receptor: a possible mechanism of bicalutamide withdrawal syndrome. Cancer Res. 63, 149–153 (2003).

    CAS  PubMed  Google Scholar 

  35. Lallous, N. et al. Functional analysis of androgen receptor mutations that confer anti-androgen resistance identified in circulating cell-free DNA from prostate cancer patients. Genome Biol. 17, 10 (2016).

    PubMed  PubMed Central  Google Scholar 

  36. Bohl, C. E., Gao, W., Miller, D. D., Bell, C. E. & Dalton, J. T. Structural basis for antagonism and resistance of bicalutamide in prostate cancer. Proc. Natl Acad. Sci. USA 102, 6201–6206 (2005).

    CAS  PubMed  PubMed Central  Google Scholar 

  37. Veldscholte, J. et al. A mutation in the ligand binding domain of the androgen receptor of human INCaP cells affects steroid binding characteristics and response to anti-androgens. Biochem. Biophys. Res. Commun. 173, 534–540 (1990).

    CAS  PubMed  Google Scholar 

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

    CAS  PubMed  PubMed Central  Google Scholar 

  39. Korpal, M. et al. An F876L mutation in androgen receptor confers genetic and phenotypic resistance to MDV3100 (enzalutamide). Cancer Discov. 3, 1030–1043 (2013).

    CAS  PubMed  Google Scholar 

  40. Balbas, M. D. et al. Overcoming mutation-based resistance to antiandrogens with rational drug design. eLife 2, e00499 (2013).

    PubMed  PubMed Central  Google Scholar 

  41. Zhu, G. et al. Pharmacological inhibition of SRC-1 phase separation suppresses YAP oncogenic transcription activity. Cell Res. 31, 1028–1031 (2021).

    CAS  PubMed  PubMed Central  Google Scholar 

  42. Wheeler, R. J. et al. Small molecules for modulating protein driven liquid-liquid phase separation in treating neurodegenerative disease. Preprint at bioRxiv https://doi.org/10.1101/721001 (2019).

  43. Risso-Ballester, J. et al. A condensate-hardening drug blocks RSV replication in vivo. Nature 595, 596–599 (2021).

    CAS  PubMed  Google Scholar 

  44. Andersen, R. J. et al. Regression of castrate-recurrent prostate cancer by a small-molecule inhibitor of the amino-terminus domain of the androgen receptor. Cancer Cell 17, 535–546 (2010).

    CAS  PubMed  Google Scholar 

  45. Myung, J.-K. et al. An androgen receptor N-terminal domain antagonist for treating prostate cancer. J. Clin. Invest. 123, 2948–2960 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  46. Zhu, J., Salvatella, X. & Robustelli, P. Small molecules targeting the disordered transactivation domain of the androgen receptor induce the formation of collapsed helical states. Preprint at bioRxiv https://doi.org/10.1101/2021.12.23.474012 (2021).

  47. Sheikhhassani, V. et al. Topological dynamics of an intrinsically disordered N‐terminal domain of the human androgen receptor. Protein Sci. 31, 1–17 (2022).

    Google Scholar 

  48. Henley, M. J. & Koehler, A. N. Advances in targeting ‘undruggable’ transcription factors with small molecules. Nat. Rev. Drug Discov. 20, 669–688 (2021).

    CAS  PubMed  Google Scholar 

  49. Naganathan, A. N. & Orozco, M. The native ensemble and folding of a protein molten-globule: functional consequence of downhill folding. J. Am. Chem. Soc. 133, 12154–12161 (2011).

    CAS  PubMed  Google Scholar 

  50. Kumar, R. & Litwack, G. Structural and functional relationships of the steroid hormone receptors’ N-terminal transactivation domain. Steroids 74, 877–883 (2009).

    CAS  PubMed  PubMed Central  Google Scholar 

  51. Shalem, O. et al. Genome-scale CRISPR–Cas9 knockout screening in human cells. Science 343, 84–87 (2014).

    CAS  PubMed  Google Scholar 

  52. Martínez, L., Andrade, R., Birgin, E. G. & Martínez, J. M. PACKMOL: a package for building initial configurations for molecular dynamics simulations. J. Comput. Chem. 30, 2157–2164 (2009).

    PubMed  Google Scholar 

  53. Miao, Y., Feher, V. A. & McCammon, J. A. Gaussian accelerated molecular dynamics: unconstrained enhanced sampling and free energy calculation. J. Chem. Theory Comput. 11, 3584–3595 (2015).

    CAS  PubMed  PubMed Central  Google Scholar 

  54. Cooper, P. A. & Shnyder, S. D. Use of the hollow fiber assay to evaluate agents that target the tumor neovasculature. Methods Mol. Biol. 1430, 375–386 (2016).

    CAS  PubMed  Google Scholar 

  55. Sharma, S. V., Haber, D. A. & Settleman, J. Cell line-based platforms to evaluate the therapeutic efficacy of candidate anticancer agents. Nat. Rev. Cancer 10, 241–253 (2010).

    CAS  PubMed  Google Scholar 

  56. Zhang, F. et al. Characterization of drug responses of mini patient-derived xenografts in mice for predicting cancer patient clinical therapeutic response. Cancer Commun. 38, 60 (2018).

    CAS  Google Scholar 

Download references

Acknowledgements

We thank J. Yuan for critical reading of the manuscript. We also thank Z. Wang and Y. Zhang for their efforts in crosslinking mass spectrometry. Thank D. Gao and Z. Liu for providing VCaP cells. We would like to acknowledge the staff members of the Large-scale Protein Preparation System at the National Facility for Protein Science in Shanghai (NFPS), Zhangjiang Laboratory, Shanghai Advanced Research Institute, Chinese Academy of Science, China for providing technical support and assistance in MST and nanoscale differentiation scanning fluorimetry data collection and analysis. This work was supported by the National Natural Science Foundation of China (21877123 to G.Z., 81803560 to L.S.).

Author information

Authors and Affiliations

Authors

Contributions

J.X., J.Z., and G.Z. conceived the project. J.X., H.H., and G.Z. performed most of the experiments and data analysis. H.H. performed the data analysis of image and RNA sequencing. W.K. assisted in plasmid construction and Z.L. helped in arranging compound characterization data. Z.G. and D.X. performed MD computer simulations. L.S. and G.L. helped with compound optimization. X.F. performed xenograft mouse model experiments. X.J. and Q.Z. designed and synthesized the compounds. J.X., H.H., J.Z., and G.Z. wrote the manuscript.

Corresponding authors

Correspondence to Jidong Zhu or Guangya Zhu.

Ethics declarations

Competing interests

J.Z. is a co-founder of Etern Biopharma. J.X., H.H., W.K., L.S., X.F., X.J., Q.Z., D.X., Z.G. and G.L. are employees of Etern Biopharma. The remaining authors declare no competing interests.

Peer review

Peer review information

Nature Chemical Biology thanks Andrew Cato, Iain McEwan and the other, anonymous, reviewer(s) for their contribution to the peer review of this work.

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 DHT-stimulated androgen receptor formed phase-separated condensates.

a. Immunofluorescence (IF) of AR in VCaP cells (AR amplification and some V7). Fluorescence signal is shown alone (green), merged with DAPI stain (blue) (Scale bar, 5 μm) and zoomed view of merged image (Scale bar, 350 nm). b. Live imaging of mEGFP tagged AR in VCaP cells. mEGFP signal is shown alone (green), mEGFP merged with Hoechst stain (blue). Scale bar, 5 μm. c. The immunoblot analysis of AR proteins in LNCaP and VCaP cells stably expressed with AR–mEGFP. d. Live imaging of mEGFP tagged AR in PC3 cells. mEGFP signal is shown alone (green), mEGFP merged with Hoechst stain (blue). Scale bar, 5 μm. e. Live imaging of mEGFP tagged AR in HEK293T cells. mEGFP signal is shown alone (green), mEGFP merged with Hoechst stain (blue). Scale bar, 5 μm. f. Live imaging of mEGFP tagged AR in H1299 cells. mEGFP signal is shown alone (green), mEGFP merged with Hoechst stain (right). Scale bar, 5 μm. g. Live imaging of mEGFP tagged AR in SF268 cells. mEGFP signal is shown alone (green), mEGFP merged with Hoechst stain (blue). Scale bar, 5 μm. h. Live imaging of mScarlet tagged AR in LNCaP cells. mScarlet signal is shown alone (magenta), mScarlet merged with Hoechst stain (blue). Scale bar, 5 μm. i. Schematic and sequencing validation of mEGFP knock-in strategy. j. Representative images of endogenous tagged AR–mEGFP treated with/without DHT stimulation. Scale bar, 5 μm (zoomed, 1 μm).

Source data

Extended Data Fig. 2 Active transcription markers can be detected in foci of AR splice variants.

a. Western blots showing the expression levels of cells expressing AR (full-length, FL) and AR (truncation)-EGFP-Flag. β-actin was used as control. Truncations containing LBD tend to have lower expression levels and we have adjusted the amounts before comparing the LLPS capability of various truncations. b. Time-lapsed images showing the FRAP experiments of AR(V7) puncta in LNCaP cells. Scale bar, 3 μm. (zoomed, 500 nm.) c. LNCaP cells transfected with AR(V7/V567es)–mEGFP were stained with anti-H3K27ac and anti-Pol II-S5P. Quantification of fluorescence intensity of AR(V7/V567es)–mEGFP and three markers along the line indicated in merged image were shown. Scale bar, 5 μm. Co-localization is assessed by Pearson’s correlation coefficient of the fluorescence intensity of each channel for all pixels and shown on the right of each group. NV7 = 82(H3K27ac), 95(PolII-S5P), NV567es = 104(H3K27ac), 120(PolII-S5P) cells. d. mNeonGreen MED1 knock-in HEK293T cells were transfected with AR(V7/V567es)-mTagBFP2 to examine the co-localization between MED1 and AR. Quantification of fluorescence intensity of AR(V7/V567es)-mTagBFP2 and mNeonGreen-MED1 along the line indicated in merged image were shown. Scale bar, 3 μm. Co-localization is assessed by Pearson’s correlation coefficient of the fluorescence intensity of each channel for all pixels and shown on the right of each group. NV567es = 91, NV7 = 89 cells.

Source data

Extended Data Fig. 3 Role of DBD in AR condensate formation and in vitro AR droplet formation.

a. Living-cell image of mEGFP labeled AR(WT) and AR(A574D) treated with DHT in LNCaP cells. Reconstructed images were shown and the puncta number and area of AR variants were analyzed. N = 100 cells. Scale bar, 2 μm. b. Time-lapsed images recording the fusion and fission events of AR(A574D) puncta in LNCaP cells. Scale bar, 5 μm. c. Images of AR(NTD), AR(DBD) and AR(LBD) in vitro phase separation. Each protein was mixed with a LLPS buffer containing 25 mM Tris pH8.0, 15%(w/v) PEG3350 and incubated for 10 min at 37 °C. Scale bar, 5 μm. d. Time-lapsed images of FRAP experiments in AR(NTD)-EGFP droplets. Scale bar, 2.5 μm. e. In vitro phase separation assays of AR(NTD)–mEGFP proteins under different concentrations of sodium chloride. Scale bar, 7.5 μm. f. Predictions of sequences mediating LLPS by catGRANULE score. G425-D496 showed the highest score. g. Western blots showing the expression levels of cells expressing AR (full-length, FL) and AR (deletions)-EGFP-Flag. GAPDH was used as control.

Source data

Extended Data Fig. 4 AR phase separation underlies mutation mediated-antiandrogen resistance.

a. Living-cell imaging of LNCaP cells expressing AR(WT) and AR(T878A)-EGFP treated with hydroxyflutamide (Hof) in the absence/presence of DHT. Scale bar, 3 μm. b. Box plot showing puncta number in LNCaP cells expressing AR(WT) and AR(T878A)-EGFP treated with hydroxyflutamide in the absence/presence of DHT (N = 71-200 cells). c. Transactivation reporter (ARE) assays of HEK293T cells expressing AR(WT) and AR(T878A) treated with hydroxyflutamide in the absence/presence of DHT. Data are presented as the mean ± SD, n = 3 biologically independent samples per group. d. Transactivation reporter (ARE) and puncta number statistical analyses of HEK293T cells expressing AR(WT) and AR(T878A) treated with varying concentrations of 17β-estradiol (E2) and progesterone. Data are presented as the mean ± SD, n = 3 biologically independent samples for E2 treatment and 2 for progesterone treatment. N = 200 cells for puncta analysis. e. Representative images of LNCaP cells expressing mEGFP tagged AR variants treated with E2 and progesterone were shown. Scale bar, 3 μm. f. Living-cell imaging of LNCaP cells expressing AR(WT), AR(T878A) and AR(W742C)-EGFP treated with enzalutamide (Enza) in the absence/presence of DHT. Scale bar, 3 μm. g. Box plot showing puncta number in LNCaP cells expressing AR(WT), AR(T878A) and AR(W742C)-EGFP treated with enzalutamide in the absence/presence of DHT (N = 73-200 cells). h. Transactivation reporter (ARE) assays of HEK293T cells expressing AR(WT), AR(T878A) and AR(W742C) treated with enzalutamide in the absence/presence of DHT. Data are presented as the mean ± SD, n = 3 biologically independent samples per group.

Source data

Extended Data Fig. 5 NTD inhibitor EPI001 exhibits weak inhibition on condensates of AR resistant mutations.

a. Living-cell imaging of LNCaP cells expressing AR(WT), AR(T878A), AR(W742C) and AR(F877L/T878A)-EGFP treated with EPI001 upon DHT stimulation. Scale bar, 2.5 μm. b. Box plot showing puncta number in LNCaP cells expressing AR(WT), AR(T878A), AR(W742C) and AR(F877L/T878A)-EGFP treated with EPI001 in the absence/presence of DHT (N = 50–100 cells). c. Transactivation reporter (ARE) assays of HEK293T cells expressing AR(WT), AR(T878A), AR(W742C) and AR(F877L/T878A) treated with EPI001 in the absence/presence of DHT. Data are presented as the mean ± SD, n = 3 biologically independent samples per group.

Source data

Extended Data Fig. 6 Identification of ET516 as a potential AR inhibitor.

a. Schematic representation of image-based LLPS screen combined with two parallel reporter assays. b. Three-dimensional scatter plot depicting the overall effects on both puncta, transactivation reporter and proliferation of compound library. The high content imaging features were first reduced by principal component analysis (PCA). The dissimilarity was calculated as spatial Euclidean distance to the center of vehicle group using first three components of principal component analysis from high content imaging features. The ARE/PSA reporters were averaged to one dimension. c. Principal component analysis of compound library screening using high content imaging data. All imaging features were extracted with Operetta CLS built-in methods. d. Pairwise scatter plot of ARE/PSA reporter and image-based puncta number (filtered by normalized CTG > 0.7). The linear regression line was added with 95% confidence interval (line range). Pearson correlation coefficient (R) and two-tailed p-value were also calculated and labeled. e. Compounds ranked by the antagonizing effects of AR activity. The inhibitory activity was calculated as normalized spatial distance to vehicle group by merging image level dissimilarity, transactivation reporter and proliferation data in (c). f. Immunoblot analysis of LNCaP cells expressing AR–mEGFP treated with 20 μM ET516 for indicated time points.

Source data

Extended Data Fig. 7 ET516 binds to a redistributed conformation of AF1.

a. Representative western blot showing thermostable AR in HEK293T cells expressing AR-EGFP following heat shock in the presence of 10 μM ET516 with 100 nM DHT. GAPDH was used as loading control. b. Melt and shift curve of AR in HEK293T cells expressing AR–mEGFP treated with ET516 (blue) and DMSO control (black) in the presence of 100 nM DHT. Data are represented as mean ± SEM, n = 3. c. Microscale thermophoresis experiments performed with AR (NTD, 1-555aa)–mEGFP proteins incubated with varying concentrations of ET516/Enzalutamide. mEGFP proteins were used as negative control. MO.Affinity Analysis software was used to fit the data. d. Steady-state fluorescence spectra of AR-AF1(144-450aa) protein(0.3 mg/mL) in 25 mM Tris-HCl pH8.0, 150 mM NaCl buffer treated with ET516. e. Thermal shift assay in AR AF-1(144-450aa, 0.3 mg/ml) incubated with ET516. The heating process increased from 25 to 81.5°C at a ramp rate of 1°C/min. Ratio of emission intensities (Em350nm/Em330nm) was recorded and plotted as a function of temperature. The fluorescence intensity ratio and first derivative were analyzed from PR.ThermControl. f. Principal component analysis (PCA) of Tau5_R2R3 structures extracted from GaMD trajectories. Tau5_R2R3 structures extracted from MD simulations at 2 ps intervals are merged and projected onto PC axes for the first two PCs. Color indicates the reweighted free energy using cumulant expansion to the second order.

Source data

Extended Data Fig. 8 ET516 specifically repress AR-dependent signaling and inhibits cell proliferation of prostate cancer expressing mutants resistant to antiandrogens.

a. GSEA enrichment plots comparing DHT(1 nM)-stimulated LNCaP cells to vehicle cells showing positive enrichment for gene set corresponding to AR target genes. b. Differential regulated signal pathway in ET516 treated LNCaP cells compared to DHT-stimulated control cells. LNCaP cells were treated with ET516 two hour prior to the DHT stimulation for 10 hours. c. GSEA enrichment plots comparing ET516 treated LNCaP cells to control cells showing no obvious enrichment for gene set corresponding to ER and p53 target genes compared to AR target genes. d. ET516 was subjected to a panel of transcription factor reporter assays. Data are presented as the mean ± SD, n = 3 replicates. e. A panel of condensates assays to test the specificity of ET516. Scale bar, 5 μm. ET516 was added to a series of phase-separated condensates for 2 hours and the puncta number of ET516/vehicle was analyzed. Data are presented as the mean ± SD, n = 4–7 biologically independent samples per group. In d-e, hydrocortisone(10 nM), E2(10 nM) and IL-6(4 ng/ml) were used to activate glucocorticoid receptor, estrogen receptor and Stat3, respectively. YAP condensates were induced with TEAD co-transfection. f. Images of LNCaP cells expressing AR(F877L/T878A) treated with DHT combined with DMSO (vehicle), ET516 and enzalutamide (Enza) for 8 days (Top). Brightfield images of LNCaP cells expressing AR(V7) treated with DMSO (vehicle), ET516 and enzalutamide (Enza) for 8 days (Bottom). Scale bar, 100 μm. g. Representation of hollow fiber assay for evaluating the efficacy of ET516, EPI001 and enzalutamide. h. Efficacy results of hollow fiber assay that treated with (ET516, 30 mg/kg, p.o., BID; EPI001, 30 mg/kg, p.o., BID; Enzalutamide, 30 mg/kg, p.o., QD) for 10 days, Data are presented as the mean ± SD, N = 6 mice each group.

Source data

Extended Data Fig. 9 AR V7 exhibits weaker LLPS capability and less sensitive to ET516 compared with full-length.

a. Fluorescence images showing the distribution of AR(FL)–mEGFP and two splice variants AR(V7/V567es)–mEGFP. Scale bar, 3 μm. Statistical analysis of the puncta number was shown on the right. N = 100 cells. b. Representative scheme showing the statistics of puncta fluorescence contrast and the analysis results of AR full length and splice variants. N = 150 cells. c. Box plot showing puncta number in LNCaP cells expressing AR(V7)–mEGFP treated with indicated concentrations of ET516 (N = 100 cells). d. Transactivation reporter (ARE) assays of HEK293T cells expressing AR(V7)–mEGFP treated with indicated concentrations of ET516. Data are presented as the mean ± SD, n = 3 biologically independent samples per group.

Source data

Supplementary information

Supplementary Information

Supplementary Figs 1 and 2, and Supplementary Notes 1 and 2.

Reporting Summary

Source data

Source Data Fig. 1

Unprocessed confocal image data and statistical analysis source data.

Source Data Fig. 2

Unprocessed confocal image data and statistical analysis source data

Source Data Fig. 3

Unprocessed confocal image data and statistical analysis source data.

Source Data Fig. 4

Unprocessed confocal image data and statistical analysis source data.

Source Data Fig. 5

Unprocessed confocal image data and statistical analysis source data.

Source Data Extended Data Fig. 1

Unprocessed confocal image data and unprocessed gel.

Source Data Extended Data Fig. 2

Unprocessed confocal image data, statistical analysis source data, and unprocessed gel.

Source Data Extended Data Fig. 3

Unprocessed confocal image data, statistical analysis source data, and unprocessed gel

Source Data Extended Data Fig. 4

Unprocessed confocal image data and statistical analysis source data.

Source Data Extended Data Fig. 5

Unprocessed confocal image data and statistical analysis source data.

Source Data Extended Data Fig. 6

Unprocessed gel and statistical analysis source data.

Source Data Extended Data Fig. 7

Unprocessed gel and statistical analysis source data.

Source Data Extended Data Fig. 8

Unprocessed confocal image data and statistical analysis source data.

Source Data Extended Data Fig. 9

Unprocessed confocal image data and statistical analysis source data.

Rights and permissions

Springer Nature or its licensor holds exclusive rights to this article under a publishing agreement with the author(s) or other rightsholder(s); author self-archiving of the accepted manuscript version of this article is solely governed by the terms of such publishing agreement and applicable law.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Xie, J., He, H., Kong, W. et al. Targeting androgen receptor phase separation to overcome antiandrogen resistance. Nat Chem Biol 18, 1341–1350 (2022). https://doi.org/10.1038/s41589-022-01151-y

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41589-022-01151-y

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