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:

Prostate cancer reactivates developmental epigenomic programs during metastatic progression

Abstract

Epigenetic processes govern prostate cancer (PCa) biology, as evidenced by the dependency of PCa cells on the androgen receptor (AR), a prostate master transcription factor. We generated 268 epigenomic datasets spanning two state transitions—from normal prostate epithelium to localized PCa to metastases—in specimens derived from human tissue. We discovered that reprogrammed AR sites in metastatic PCa are not created de novo; rather, they are prepopulated by the transcription factors FOXA1 and HOXB13 in normal prostate epithelium. Reprogrammed regulatory elements commissioned in metastatic disease hijack latent developmental programs, accessing sites that are implicated in prostate organogenesis. Analysis of reactivated regulatory elements enabled the identification and functional validation of previously unknown metastasis-specific enhancers at HOXB13, FOXA1 and NKX3-1. Finally, we observed that prostate lineage-specific regulatory elements were strongly associated with PCa risk heritability and somatic mutation density. Examining prostate biology through an epigenomic lens is fundamental for understanding the mechanisms underlying tumor progression.

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: The AR cistrome and genome-wide H3K27ac are systematically reprogrammed during PCa progression, and AR relocates to epigenetically premarked ‘sentinel’ sites.
Fig. 2: Regulatory sites activated in mCRPC coincide with prostate developmental programs.
Fig. 3: Functionally relevant mCRPC enhancers are identified by integrating genetic and epigenetic datasets.
Fig. 4: Genetic variation in PCa is enriched in prostate lineage-specific chromatin states.

Similar content being viewed by others

Data availability

Further information and requests for resources and reagents should be directed to and will be fulfilled by the lead contact, Matthew Freedman (mfreedman@partners.org). We incorporated all the epigenomic data generated in this study into a publicly accessible resource for investigators, which is available at: http://genome.ucsc.edu/cgi-bin/hgTracks?db=hg19&hubUrl=https://de.cyverse.org/anon-files/iplant/home/dfcipc/trackhub/hub.txt. All sequencing data generated for the study has been deposited in GEO (GSE130408).

References

  1. Baca, S. C. et al. Punctuated evolution of prostate cancer genomes. Cell 153, 666–677 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  2. 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 

  3. Kunz, M. et al. RNA-seq analysis identifies different transcriptomic types and developmental trajectories of primary melanomas. Oncogene 37, 6136–6151 (2018).

    CAS  PubMed  Google Scholar 

  4. Chen, H. et al. A pan-cancer analysis of enhancer expression in nearly 9000 patient samples. Cell 173, 386–399.e12 (2018).

    CAS  PubMed  PubMed Central  Google Scholar 

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

    CAS  PubMed  PubMed Central  Google Scholar 

  6. Pomerantz, M. M. et al. The androgen receptor cistrome is extensively reprogrammed in human prostate tumorigenesis. Nat. Genet. 47, 1346–1351 (2015).

    CAS  PubMed  PubMed Central  Google Scholar 

  7. Roe, J.-S. et al. Enhancer reprogramming promotes pancreatic cancer metastasis. Cell 170, 875–888.e20 (2017).

    CAS  PubMed  PubMed Central  Google Scholar 

  8. Jones, P. A. & Baylin, S. B. The epigenomics of cancer. Cell 128, 683–692 (2007).

    CAS  PubMed  PubMed Central  Google Scholar 

  9. Dongre, A. & Weinberg, R. A. New insights into the mechanisms of epithelial-mesenchymal transition and implications for cancer. Nat. Rev. Mol. Cell Biol. 20, 69–84 (2019).

    CAS  PubMed  Google Scholar 

  10. Feldman, B. J. & Feldman, D. The development of androgen-independent prostate cancer. Nat. Rev. Cancer 1, 34–45 (2001).

    CAS  PubMed  Google Scholar 

  11. Wang, Q. et al. Androgen receptor regulates a distinct transcription program in androgen-independent prostate cancer. Cell 138, 245–256 (2009).

    CAS  PubMed  PubMed Central  Google Scholar 

  12. Stelloo, S., Bergman, A. M. & Zwart, W. Androgen receptor enhancer usage and the chromatin regulatory landscape in human prostate cancers. Endocr. Relat. Cancer 26, R267–R285 (2019).

    CAS  PubMed  Google Scholar 

  13. Stelloo, S. et al. Androgen receptor profiling predicts prostate cancer outcome. EMBO Mol. Med. 7, 1450–1464 (2015).

    CAS  PubMed  PubMed Central  Google Scholar 

  14. Kron, K. J. et al. TMPRSS2–ERG fusion co-opts master transcription factors and activates NOTCH signaling in primary prostate cancer. Nat. Genet. 49, 1336–1345 (2017).

    CAS  PubMed  Google Scholar 

  15. Chen, Z. et al. Diverse AR-V7 cistromes in castration-resistant prostate cancer are governed by HoxB13. Proc. Natl Acad. Sci. USA 115, 6810–6815 (2018).

    CAS  PubMed  PubMed Central  Google Scholar 

  16. Cato, L. et al. ARv7 represses tumor-suppressor genes in castration-resistant prostate cancer. Cancer Cell 35, 401–413.e6 (2019).

    CAS  PubMed  PubMed Central  Google Scholar 

  17. Stelloo, S. et al. Integrative epigenetic taxonomy of primary prostate cancer. Nat. Commun. 9, 4900 (2018).

    PubMed  PubMed Central  Google Scholar 

  18. Nguyen, H. M. et al. LuCaP prostate cancer patient-derived xenografts reflect the molecular heterogeneity of advanced disease and serve as models for evaluating cancer therapeutics. Prostate 77, 654–671 (2017).

    CAS  PubMed  PubMed Central  Google Scholar 

  19. Kundaje, A. et al. Integrative analysis of 111 reference human epigenomes. Nature 518, 317–330 (2015).

    CAS  PubMed  PubMed Central  Google Scholar 

  20. Guo, C. et al. Epcam, CD44, and CD49f distinguish sphere-forming human prostate basal cells from a subpopulation with predominant tubule initiation capability. PLoS ONE 7, e34219 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  21. Taylor, B. S. et al. Integrative genomic profiling of human prostate cancer. Cancer Cell 18, 11–22 (2010).

    CAS  PubMed  PubMed Central  Google Scholar 

  22. Yu, Y. P. et al. Whole-genome methylation sequencing reveals distinct impact of differential methylations on gene transcription in prostate cancer. Am. J. Pathol. 183, 1960–1970 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  23. Hanrahan, K. et al. The role of epithelial–mesenchymal transition drivers ZEB1 and ZEB2 in mediating docetaxel-resistant prostate cancer. Mol. Oncol. 11, 251–265 (2017).

    CAS  PubMed  PubMed Central  Google Scholar 

  24. Dai, Y. et al. Copy number gain of ZEB1 mediates a double-negative feedback loop with miR-33a-5p that regulates EMT and bone metastasis of prostate cancer dependent on TGF-β signaling. Theranostics 9, 6063–6079 (2019).

    CAS  PubMed  PubMed Central  Google Scholar 

  25. McLean, C. Y. et al. GREAT improves functional interpretation of cis-regulatory regions. Nat. Biotechnol. 28, 495–501 (2010).

    CAS  PubMed  PubMed Central  Google Scholar 

  26. Simons, B. W. et al. Wnt signaling though beta-catenin is required for prostate lineage specification. Dev. Biol. 371, 246–255 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  27. Lee, S. H. et al. Wnt/β-catenin-responsive cells in prostatic development and regeneration. Stem Cells 33, 3356–3367 (2015).

    PubMed  PubMed Central  Google Scholar 

  28. Kruithof-de Julio, M. et al. Canonical Wnt signaling regulates Nkx3.1 expression and luminal epithelial differentiation during prostate organogenesis. Dev. Dyn. 242, 1160–1171 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  29. Branam, A. M. et al. TCDD inhibition of canonical Wnt signaling disrupts prostatic bud formation in mouse urogenital sinus. Toxicol. Sci. 133, 42–53 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  30. Huang, L. et al. The role of Wnt5a in prostate gland development. Dev. Biol. 328, 188–199 (2009).

    CAS  PubMed  PubMed Central  Google Scholar 

  31. Toivanen, R. & Shen, M. M. Prostate organogenesis: tissue induction, hormonal regulation and cell type specification. Development 144, 1382–1398 (2017).

    CAS  PubMed  PubMed Central  Google Scholar 

  32. Patten, D. K. et al. Enhancer mapping uncovers phenotypic heterogeneity and evolution in patients with luminal breast cancer. Nat. Med. 24, 1469–1480 (2018).

    CAS  PubMed  PubMed Central  Google Scholar 

  33. Watson, C. J. & Khaled, W. T. Mammary development in the embryo and adult: a journey of morphogenesis and commitment. Development 135, 995–1003 (2008).

    CAS  PubMed  Google Scholar 

  34. Pritchard, C. et al. Conserved gene expression programs integrate mammalian prostate development and tumorigenesis. Cancer Res. 69, 1739–1747 (2009).

    CAS  PubMed  Google Scholar 

  35. Takeda, D. Y. et al. A somatically acquired enhancer of the androgen receptor is a noncoding driver in advanced prostate cancer. Cell 174, 422–432.e13 (2018).

    CAS  PubMed  PubMed Central  Google Scholar 

  36. Viswanathan, S. R. et al. Structural alterations driving castration-resistant prostate cancer revealed by linked-read genome sequencing. Cell 174, 433–447.e19 (2018).

    CAS  PubMed  PubMed Central  Google Scholar 

  37. Quigley, D. A. et al. Genomic hallmarks and structural variation in metastatic prostate cancer. Cell 174, 758–769.e9 (2018).

    CAS  PubMed  PubMed Central  Google Scholar 

  38. Bhatia-Gaur, R. et al. Roles for Nkx3.1 in prostate development and cancer. Genes Dev. 13, 966–977 (1999).

    CAS  PubMed  PubMed Central  Google Scholar 

  39. Dutta, A. et al. Identification of an NKX3.1-G9a-UTY transcriptional regulatory network that controls prostate differentiation. Science 352, 1576–1580 (2016).

    CAS  PubMed  PubMed Central  Google Scholar 

  40. Tsherniak, A. et al. Defining a cancer dependency map. Cell 170, 564–576.e16 (2017).

    CAS  PubMed  PubMed Central  Google Scholar 

  41. Horoszewicz, J. S. et al. LNCaP model of human prostatic carcinoma. Cancer Res. 43, 1809–1818 (1983).

    CAS  PubMed  Google Scholar 

  42. Economides, K. D. & Capecchi, M. R. Hoxb13 is required for normal differentiation and secretory function of the ventral prostate. Development 130, 2061–2069 (2003).

    CAS  PubMed  Google Scholar 

  43. Gao, N. et al. The role of hepatocyte nuclear factor-3α (forkhead box A1) and androgen receptor in transcriptional regulation of prostatic genes. Mol. Endocrinol. 17, 1484–1507 (2003).

    CAS  PubMed  Google Scholar 

  44. Hankey, W., Chen, Z. & Wang, Q. Shaping chromatin states in prostate cancer by pioneer transcription factors. Cancer Res. 80, 2427–2436 (2020).

    PubMed  PubMed Central  Google Scholar 

  45. Schumacher, F. R. et al. Association analyses of more than 140,000 men identify 63 new prostate cancer susceptibility loci. Nat. Genet. 50, 928–936 (2018).

    CAS  PubMed  PubMed Central  Google Scholar 

  46. Abeshouse, A. et al. The molecular taxonomy of primary prostate cancer. Cell 163, 1011–1025 (2015).

    CAS  Google Scholar 

  47. Robinson, D. et al. Integrative clinical genomics of advanced prostate cancer. Cell 161, 1215–1228 (2015).

    CAS  PubMed  PubMed Central  Google Scholar 

  48. Ernst, J. & Kellis, M. Chromatin-state discovery and genome annotation with ChromHMM. Nat. Protoc. 12, 2478–2492 (2017).

    CAS  PubMed  PubMed Central  Google Scholar 

  49. Michailidou, K. et al. Association analysis identifies 65 new breast cancer risk loci. Nature 551, 92–94 (2017).

    PubMed  PubMed Central  Google Scholar 

  50. Mazrooei, P. et al. Cistrome partitioning reveals convergence of somatic mutations and risk variants on master transcription regulators in primary prostate tumors. Cancer Cell 36, 674–689.e6 (2019).

    CAS  PubMed  Google Scholar 

  51. Morova, T. et al. Androgen receptor-binding sites are highly mutated in prostate cancer. Nat. Commun. 11, 832 (2020).

    CAS  PubMed  PubMed Central  Google Scholar 

  52. Zhou, S. et al. Noncoding mutations target cis-regulatory elements of the FOXA1 plexus in prostate cancer. Nat. Commun. 11, 441 (2020).

    CAS  PubMed  PubMed Central  Google Scholar 

  53. Waddington, C. H. The Strategy of the Genes: a Discussion of Some Aspects of Theoretical Biology (George Allen & Unwin, 1957).

  54. Goldberg, A. D., Allis, C. D. & Bernstein, E. Epigenetics: a landscape takes shape. Cell 128, 635–638 (2007).

    CAS  PubMed  Google Scholar 

  55. Tam, W. L. & Weinberg, R. A. The epigenetics of epithelial–mesenchymal plasticity in cancer. Nat. Med. 19, 1438–1449 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  56. Beard, J. Embryological aspects and etiology of carcinoma. Lancet 159, 1758–1761 (1902).

    Google Scholar 

  57. Markert, C. L. Neoplasia: a disease of cell differentiation. Cancer Res. 28, 1908–1914 (1968).

    CAS  PubMed  Google Scholar 

  58. Schaeffer, E. M. et al. Androgen-induced programs for prostate epithelial growth and invasion arise in embryogenesis and are reactivated in cancer. Oncogene 27, 7180–7191 (2008).

    CAS  PubMed  PubMed Central  Google Scholar 

  59. Stergachis, A. B. et al. Developmental fate and cellular maturity encoded in human regulatory DNA landscapes. Cell 154, 888–903 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  60. Gupta, P. B. et al. The melanocyte differentiation program predisposes to metastasis after neoplastic transformation. Nat. Genet. 37, 1047–1054 (2005).

    CAS  PubMed  PubMed Central  Google Scholar 

  61. Latil, M. et al. Cell-type-specific chromatin states differentially prime squamous cell carcinoma tumor-initiating cells for epithelial to mesenchymal transition. Cell Stem Cell 20, 191–204.e5 (2017).

    CAS  PubMed  Google Scholar 

  62. Krebs, A. M. et al. The EMT-activator Zeb1 is a key factor for cell plasticity and promotes metastasis in pancreatic cancer. Nat. Cell Biol. 19, 518–529 (2017).

    CAS  PubMed  Google Scholar 

  63. Sun, Y. et al. Androgen deprivation causes epithelial–mesenchymal transition in the prostate: implications for androgen-deprivation therapy. Cancer Res. 72, 527–536 (2012).

    CAS  PubMed  Google Scholar 

  64. Montanari, M. et al. Epithelial–mesenchymal transition in prostate cancer: an overview. Oncotarget 8, 35376–35389 (2017).

    PubMed  PubMed Central  Google Scholar 

  65. Graham, T. R. et al. Insulin-like growth factor-I-dependent up-regulation of ZEB1 drives epithelial-to-mesenchymal transition in human prostate cancer cells. Cancer Res. 68, 2479–2488 (2008).

    CAS  PubMed  Google Scholar 

  66. Figiel, S. et al. Clinical significance of epithelial–mesenchymal transition markers in prostate cancer. Hum. Pathol. 61, 26–32 (2017).

    CAS  PubMed  Google Scholar 

  67. Morton, A. R. et al. Functional enhancers shape extrachromosomal oncogene amplifications. Cell 179, 1330–1341.e13 (2019).

    CAS  PubMed  PubMed Central  Google Scholar 

  68. Gusev, A. et al. Atlas of prostate cancer heritability in European and African-American men pinpoints tissue-specific regulation. Nat. Commun. 7, 10979 (2016).

    CAS  PubMed  PubMed Central  Google Scholar 

  69. Lindblad-Toh, K. et al. A high-resolution map of human evolutionary constraint using 29 mammals. Nature 478, 476–482 (2011).

    CAS  PubMed  PubMed Central  Google Scholar 

  70. Yang, J. et al. Recurrent mutations at estrogen receptor binding sites alter chromatin topology and distal gene expression in breast cancer. Genome Biol. 19, 190 (2018).

    PubMed  PubMed Central  Google Scholar 

  71. Mao, P. et al. ETS transcription factors induce a unique UV damage signature that drives recurrent mutagenesis in melanoma. Nat. Commun. 9, 2626 (2018).

    PubMed  PubMed Central  Google Scholar 

  72. Gonzalez-Perez, A., Sabarinathan, R. & Lopez-Bigas, N. Local determinants of the mutational landscape of the human genome. Cell 177, 101–114 (2019).

    CAS  PubMed  Google Scholar 

  73. Fraser, M. et al. Genomic hallmarks of localized, non-indolent prostate cancer. Nature 541, 359–364 (2017).

    CAS  PubMed  Google Scholar 

  74. Jacob, F. Evolution and tinkering. Science 196, 1161–1166 (1977).

    CAS  PubMed  Google Scholar 

  75. Dang, C. V., Reddy, E. P., Shokat, K. M. & Soucek, L. Drugging the ‘undruggable’ cancer targets. Nat. Rev. Cancer 17, 502–508 (2017).

    CAS  PubMed  PubMed Central  Google Scholar 

  76. Chen, A. & Koehler, A. N. Drug discovery. Tying up a transcription factor. Science 347, 713–714 (2015).

    CAS  PubMed  Google Scholar 

  77. Morrow, J. J. et al. Positively selected enhancer elements endow osteosarcoma cells with metastatic competence. Nat. Med. 24, 176–185 (2018).

    CAS  PubMed  PubMed Central  Google Scholar 

  78. Suvà, M. L., Riggi, N. & Bernstein, B. E. Epigenetic reprogramming in cancer. Science 339, 1567–1570 (2013).

    PubMed  Google Scholar 

  79. de Thé, H. Differentiation therapy revisited. Nat. Rev. Cancer 18, 117–127 (2018).

    PubMed  Google Scholar 

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

    PubMed  Google Scholar 

  81. 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 

  82. Corces, M. R. et al. An improved ATAC-seq protocol reduces background and enables interrogation of frozen tissues. Nat. Methods 14, 959–962 (2017).

    CAS  PubMed  PubMed Central  Google Scholar 

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

    PubMed  PubMed Central  Google Scholar 

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

    PubMed  PubMed Central  Google Scholar 

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

    CAS  PubMed  PubMed Central  Google Scholar 

  86. Krueger, F. & Andrews, S. R. Bismark: a flexible aligner and methylation caller for bisulfite-seq applications. Bioinformatics 27, 1571–1572 (2011).

    CAS  PubMed  PubMed Central  Google Scholar 

  87. Hansen, K. D., Langmead, B. & Irizarry, R. A. BSmooth: from whole genome bisulfite sequencing reads to differentially methylated regions. Genome Biol. 13, R83 (2012).

    PubMed  PubMed Central  Google Scholar 

  88. Lawrence, M., Gentleman, R. & Carey, V. rtracklayer: an R package for interfacing with genome browsers. Bioinformatics 25, 1841–1842 (2009).

    CAS  PubMed  PubMed Central  Google Scholar 

  89. Korthauer, K., Chakraborty, S., Benjamini, Y. & Irizarry, R. A. Detection and accurate false discovery rate control of differentially methylated regions from whole genome bisulfite sequencing. Biostatistics 20, 367–383 (2019).

    PubMed  Google Scholar 

  90. Qin, Q. et al. ChiLin: a comprehensive ChIP–seq and DNase-seq quality control and analysis pipeline. BMC Bioinformatics 17, 404 (2016).

    PubMed  PubMed Central  Google Scholar 

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

    PubMed  PubMed Central  Google Scholar 

  92. Hollander, M. & Wolfe, D. A. Nonparametric Statistical Methods (Wiley, 1973).

  93. Siepel, A. et al. Evolutionarily conserved elements in vertebrate, insect, worm, and yeast genomes. Genome Res. 15, 1034–1050 (2005).

    CAS  PubMed  PubMed Central  Google Scholar 

  94. Siepel, A. & Haussler, D. Combining phylogenetic and hidden Markov models in biosequence analysis. J. Comput. Biol. 11, 413–428 (2004).

    CAS  PubMed  Google Scholar 

  95. Finucane, H. K. et al. Partitioning heritability by functional annotation using genome-wide association summary statistics. Nat. Genet. 47, 1228–1235 (2015).

    CAS  PubMed  PubMed Central  Google Scholar 

  96. Marchal, C. et al. Genome-wide analysis of replication timing by next-generation sequencing with E/L Repli-seq. Nat. Protoc. 13, 819–839 (2018).

    CAS  PubMed  PubMed Central  Google Scholar 

  97. Gehring, J. S., Fischer, B., Lawrence, M. & Huber, W. SomaticSignatures: inferring mutational signatures from single-nucleotide variants. Bioinformatics 31, 3673–3675 (2015).

    CAS  PubMed  PubMed Central  Google Scholar 

Download references

Acknowledgements

We thank M. Brown (DFCI) and members of the Center for Functional Cancer Epigenetics at DFCI for useful discussions and technical assistance. We also thank the NKI Core Facility Molecular Pathology and Biobanking for technical assistance and tissue processing, and the NKI Genomics Core Facility for the Illumina sequencing analyses. We thank K. Schuurman, D. Sondheim and J. Conner for technical support. We also thank P. Nelson and C. Pritchard for their contributions to the dataset. This work was supported by R. and N. Milikowsky (to M.M.P.); a Prostate Cancer Foundation Challenge Award (to M.M.P. and M.L.F.); NIH grant nos. R01GM107427 and R01CA193910 (to M.L.F.); the H.L. Snyder Medical Research Foundation (to M.L.F.); Department of Defense (DOD) grant no.W81XWH-19-1-0565 (to M.L.F., M.M.P. and W.Z.); a VIDI grant from the Netherlands Organisation for Scientific Research (to W.Z.); the Dutch Cancer Society/Alpe d’HuZes (10084) and Oncode Institute (to W.Z.); NIH grant no. K08 13 CA218530 (to D.Y.T.); the Jean Perkins Foundation; the Prostate Cancer Foundation; the STOP Cancer Foundation; Department of Defense (DOD) grant no. W81XWH-14-1-0273; National Cancer Institute/NIH grant no. P50CA092131 (to I.P.G.); the PNW Prostate Cancer SPORE no. P50 CA097186; DOD grant no. W81XWH-17-1-0415; and grant no. P01 CA163227. The IPCR supported the establishment and generation of the LuCaP PDXs models. We thank the patients who generously donated the tissue that made this research possible.

Author information

Authors and Affiliations

Authors

Contributions

M.M.P., H.W.L., W.Z. and M.L.F. conceptualized the study. M.M.P., X.Q., Y.Z., D.Y.T., K.D.K., S.C.B., A.G., T.M.S., G.H., S.R.V., J.-H.S., B.P., C.G., S.A.G., K.L., G.M., B.E., L.E., H.W.L., W.Z. and M.L.F. developed the methodology. X.Q., Y.Z., D.Y.T., K.D.K., S.C.B., A.G., T.M.S., G.H., S.R.V., L.L., M.A.S.F., J.R., R.I.C., W.P. and G.-S.M.L. carried out the formal analysis. Y.Z., D.Y.T., J.-H.S., S.A.A., C.A.B., E.P.O., M.S.C., J.S., R.L., D.R.S., A.F.-T., P.C., W.P. and G.-S.M.L. carried out the investigation. M.M.P., A.M.B., H.G.v.d.P., E.C., I.P.G., B.Z., H.M.N., T.C., W.Z. and M.L.F. obtained the resources. M.M.P., S.C.B., A.G., H.W.L., W.Z. and M.L.F. wrote the original draft. M.M.P., X.Q., Y.Z., D.Y.T., K.D.K., S.C.B., A.G., T.M.S., T.C., L.E., E.C., H.W.L., W.Z. and M.L.F. reviewed and edited the draft. M.M.P., X.Q. and H.W.L., M.L.F. visualized the study. M.M.P., H.W.L., W.Z. and M.L.F. supervised the study.

Corresponding authors

Correspondence to Wilbert Zwart or Matthew L. Freedman.

Ethics declarations

Competing interests

The authors declare no competing interests.

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 Co-occupancy of AR and H3K27Ac at met-ARBS.

a, Heatmaps for AR and H3K27Ac ChIP-seq signal intensity at met-ARBS. Each horizontal line represents a four kilobase (kb) locus. Shade of red reflects average binding intensity at that site across all subjects in the normal prostate, primary tumor and mCRPC cohorts. b, H3K27Ac ChIP-seq signal intensity across tissue types at the 17,655 met-ARBS. The curves depict overall signal in each of the three tissue types. Signal significantly higher in mCRPC compared with primary prostate tumor and normal prostate tissue (Kolmogorov-Smirnov test, D^- = 0.74, p-value < 2.2e-16).

Extended Data Fig. 2 Genes that are down-regulated in metastasis compared to primary tumor are enriched for primary tumor-specific H3K27Ac ChIP-seq peaks.

Each dot represents a gene. Red dots are genes with a primary tumor-specific H3K27Ac peak (that is, sites with H3K27Ac signal in primary tumor and absent in mCRPC) in the transition start site (p-value <0.00001 for association between primary tumor-specific H3K27Ac and transcriptional down-regulation in mCRPC).

Extended Data Fig. 3 Reprogrammed AR binding sites in primary prostate tumors and in mCRPC are epigenetically pre-marked in earlier states.

a, Heat map indicating HOXB13 and FOXA1 ChIP-seq signal intensity in normal prostate epithelium and primary prostate tumor in the NKI dataset. At left, the 9,179 AR sites enriched in primary tumor relative to normal prostate epithelium (T-ARBS). At right, the 17,655 AR sites enriched in mCRPC relative to primary tumor tissue (met-ARBS). Each horizontal line represents a four kilobase (kb) locus. Shade of red reflects binding intensity. b, AR ChIP-seq binding intensity across clinical tissue subtypes in T-ARBS and met-ARBS. c, Average DNA methylation signal at T-ARBS across prostate tumor (red curve) and normal prostate (blue curve) at T-ARBS (top) and met-ARBS (bottom).

Extended Data Fig. 4 GREAT analysis characterizing mCRPC-enriched epigenetic sites.

a, GREAT analysis characterizing the gene ontology biological terms most significantly associated with genes proximal to the 17,655 met-ARBS. b, GREAT analysis characterizing the gene ontology biological terms most significantly associated with genes proximal to the subset 17,655 met-ARBS that are co-occupied by H3K27Ac. Terms associated with genitourinary development are highlighted in yellow. c, The biological terms most significantly associated with genes proximal to the 2,683 AR sites enriched in primary tumor compared to mCRPC. d, GREAT analysis of the MSigDB pathway terms most significantly associated with genes proximal to met-ARBS.

Extended Data Fig. 5 Across 27 human adult tissues and 10 fetal tissues, the met-K27ac cistrome is most strongly associated with fetal urogenital sinus.

a, Tissue type listed at left (adult tissues are followed by their Roadmap Epigenomics Project identification codes). Multiple biologic replicates were performed and included here. Urogenital sinus sample was performed in replicate. Heat map indicates H3K27Ac binding intensity at the 16,047 met-K27ac sites across a 4 kilobase (kb) interval. b, Heat map for subset of met-K27ac sites that are co-occupied by AR.

Extended Data Fig. 6 Association between fetal and mature prostate murine gene expression and met-K27ac sites.

Gene expression in mouse prostate embryonic (red) and post-natal (blue) tissue34 at (a) the 50 most differential H3K27Ac sites between mCRPC and localized PCa in humans that reside within transcriptional start sites; (b) the 100 most differential H3K27Ac sites; (c) the 500 most differential H3K27Ac sites; and (d) at a randomly selected set of 500 genes that do not overlap with met-K27ac sites. Expression levels were performed in three replicates and measured relative to embryonic day 14 (y-axis). The x-axis shows embryonic days 15, 16 and 17 then post-natal days 7, 30 and 90. Box plots depict median, 25th–75th percentile interval and extremes in gene expression.

Extended Data Fig. 7 Enhancers of FOXA1 in mCRPC are identified by integrating genetic and epigenetic datasets.

a, At top, color-coded tracks in a 183 kilobase (kb) region derived from the segments ranked in Fig. 3. Tracks depict the intensity of ChIP-seq signal averaged across all DFCI normal prostate, primary prostate tumor and mCRPC specimens, respectively. FOXA1 is visualized in the Genes track. HiChIP track depicts chromatin looping in the LNCaP cell line. Blue bars show H3K27Ac sites meeting criteria for mCRPC enrichment (met-K27ac). Orange bars depict the locus against which guide RNAs (gRNAs) were designed (Methods). b, Functional interrogation of candidate metastasis-specific enhancers. Left, LNCaP FOXA1 expression in two controls (no gRNA and gRNA targeting unrelated gene HPRT1) and after transduction with each individual gRNA depicted in (a). Middle and right, LNCaP cell proliferation over the course of four days after control conditions of transduction with one of the three FOXA1 region gRNAs. Each shape represents an independent experiment, center line indicates mean, error bars indicate ± s.d. Using student’s t-test – n.s not significant, *p < 0.05, **p < 0.01, ***p < 0.001.

Extended Data Fig. 8 Enhancer of NKX3-1 in mCRPC is identified by integrating genetic and epigenetic datasets.

a, At top, color-coded tracks in the 2,456 kb region depict the intensity of ChIP-seq signal averaged across all DFCI normal prostate, primary prostate tumor and mCRPC specimens, respectively. NKX3-1 is visualized in the Genes track. HiChIP track depicts chromatin looping in the LNCaP cell line. Blue bars show H3K27Ac sites meeting criteria for mCRPC enrichment (met-K27ac). Orange bars depict the locus against which guide RNAs (gRNAs) were designed (Methods). Below, magnification of an 85 kb region where met-K27ac and HiChIP signal were strongest. b, Functional interrogation of the candidate metastasis-specific enhancer. LNCaP NKX3-1 expression in two controls (no gRNA and gRNA targeting unrelated gene HPRT1) and after transduction with gRNAs depicted in (a). Data represent the average and standard deviation of three biological replicates and significance determined by unpaired Student’s t test. * p < 0.001.

Extended Data Fig. 9 Prostate cancer and breast cancer risk heritability attributable to germline variation within prostate tumor chromatin states.

a, Prostate cancer heritability attributable to each prostate cancer chromHMM state. b, Breast cancer heritability attributable to each prostate cancer chromHMM state. %SNPs: percentage of single nucleotide polymorphisms residing within a chromatin state; %h2: proportion of prostate cancer risk heritability; se: standard error; Enrichment: heritability based on overall proportion of SNPs within the chromatin state. c, Q-Q Plot of PCa risk GWAS statistics in lineage specific and non-specific features. Lineage specific promoters, enhancers, and all other variants shown in green, orange, and black respectively. Variants with Chi-squared statistic > 80 were removed, as recommend by LD-score regression to mitigate outliers. Across all variants, mean Chi-squared statistic was 1.6 (s.e. 0.04), 1.7 (s.e. 0.07), and 1.2 (s.e. 0.003) for variants in promoters, enhancers, and all variants.

Extended Data Fig. 10 Prostate cancer somatic mutations are enriched at prostate lineage specific sites.

a, Rank-ordered terms in a linear model of somatic mutation density in prostate cancer. Using 210 prostate cancer whole genome sequences from the International Cancer Genome Consortium, the number of donors with one or more mutations per 200 bp window was modeled as a poisson distribution determined by a linear combination of the listed factors. Beta coefficients for each term were calculated and are reported as standardized Z-scores to allow comparison. ChromHMM states are highlighted in gray. See Methods for details and a listing of datasets used in the model. b, SNV distribution at FOXA1 binding sites in prostate tumor tissue. c, SNV distribution at FOXA1 binding sites with no overlapping AR peak in prostate tumors (left), at intersection of FOXA1 and AR tumor peaks (center), and at AR tumor binding sites without overlapping FOXA1 peaks. P-values compare differential enrichment by Pearson’s chi-square test of mutation counts at the peak (±250 bp) and shoulder regions (-1000 to -250 and 250 to 1000) of the TF binding sites. d, SNV distribution at met-ARBS.

Supplementary information

Supplementary Information

Supplementary Figs. 1–6

Reporting Summary

Supplementary Tables

Supplementary Tables 1–8

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Pomerantz, M.M., Qiu, X., Zhu, Y. et al. Prostate cancer reactivates developmental epigenomic programs during metastatic progression. Nat Genet 52, 790–799 (2020). https://doi.org/10.1038/s41588-020-0664-8

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41588-020-0664-8

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