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.

  • Review Article
  • Published:

The role of enhancers in cancer

Key Points

  • Enhancer malfunction is a key process that drives the aberrant regulation of oncogenes in cancer.

  • Enhancer variants contribute more than any other known mechanism to heritable cancer predisposition.

  • Somatic mutations affecting enhancers are common in sporadic tumours. They include copy number alterations that increase the strength of an enhancer, structural rearrangements that direct enhancers to new targets and point mutations or small insertions or deletions (indels) that create new enhancers by altering transcription factor binding sites.

  • Epigenetic mechanisms also commonly alter enhancer activities in cancer.

  • During normal development and homeostasis, very strong enhancers or super-enhancers are found near lineage-specifying genes. During tumorigenesis, super-enhancers commonly form de novo near oncogenes.

  • Strong enhancer activities in cancer cells lock the growth regulatory network to an 'on' state, driving uncontrolled proliferation.

  • Enhancer malfunction can be targeted for cancer therapy. Although transcription is a general property of all cells, cancer cells are more dependent on increased transcription levels from enhancers and, thus, are more sensitive to enhancer inhibition than their normal counterparts.

Abstract

Enhancer elements function as the logic gates of the genetic regulatory circuitry. One of their most important functions is the integration of extracellular signals with intracellular cell fate information to generate cell type-specific transcriptional responses. Mutations occurring in cancer often misregulate enhancers that normally control the signal-dependent expression of growth-related genes. This misregulation can result from trans-acting mechanisms, such as activation of the transcription factors or epigenetic regulators that control enhancer activity, or can be caused in cis by direct mutations that alter the activity of the enhancer or its target gene specificity. These processes can generate tumour type-specific super-enhancers and establish a 'locked' gene regulatory state that drives the uncontrolled proliferation of cancer cells. Here, we review the role of enhancers in cancer, and their potential as therapeutic targets.

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

Figure 1: Regulatory elements controlling gene expression and the markers used to identify them.
Figure 2: The activity of enhancers and super-enhancers is cell- and tumour-type specific.
Figure 3: Germline variants and somatic mutations affecting enhancer activity differ in the magnitude of the resulting transcriptional change.
Figure 4: A model for tumour-type specificity of enhancers and oncogenic transcription factors.

Similar content being viewed by others

References

  1. Futreal, P. A. et al. A census of human cancer genes. Nat. Rev. Cancer 4, 177–183 (2004).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  2. Forbes, S. A. et al. COSMIC: mining complete cancer genomes in the Catalogue of Somatic Mutations in Cancer. Nucleic Acids Res. 39, D945–D950 (2011).

    Article  CAS  PubMed  Google Scholar 

  3. Lawrence, M. S. et al. Discovery and saturation analysis of cancer genes across 21 tumour types. Nature 505, 495–501 (2014).

    CAS  PubMed  PubMed Central  Google Scholar 

  4. Vogelstein, B. et al. Cancer genome landscapes. Science 339, 1546–1558 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

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

    Article  CAS  PubMed  Google Scholar 

  6. Dalla-Favera, R. et al. Human c-myc onc gene is located on the region of chromosome 8 that is translocated in Burkitt lymphoma cells. Proc. Natl Acad. Sci. USA 79, 7824–7827 (1982).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  7. Taub, R. et al. Translocation of the c-myc gene into the immunoglobulin heavy chain locus in human Burkitt lymphoma and murine plasmacytoma cells. Proc. Natl Acad. Sci. USA 79, 7837–7841 (1982).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  8. Ar-Rushdi, A. et al. Differential expression of the translocated and the untranslocated c-myc oncogene in Burkitt lymphoma. Science 222, 390–393 (1983).

    Article  CAS  PubMed  Google Scholar 

  9. Erikson, J., ar-Rushdi, A., Drwinga, H. L., Nowell, P. C. & Croce, C. M. Transcriptional activation of the translocated c-myc oncogene in Burkitt lymphoma. Proc. Natl Acad. Sci. USA 80, 820–824 (1983).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  10. Frohling, S. & Dohner, H. Chromosomal abnormalities in cancer. N. Engl. J. Med. 359, 722–734 (2008).

    Article  CAS  PubMed  Google Scholar 

  11. Dekker, J., Marti-Renom, M. A. & Mirny, L. A. Exploring the three-dimensional organization of genomes: interpreting chromatin interaction data. Nat. Rev. Genet. 14, 390–403 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  12. Schwarzer, W. & Spitz, F. The architecture of gene expression: integrating dispersed cis-regulatory modules into coherent regulatory domains. Curr. Opin. Genet. Dev. 27, 74–82 (2014).

    Article  CAS  PubMed  Google Scholar 

  13. Gorkin, D. U., Leung, D. & Ren, B. The 3D genome in transcriptional regulation and pluripotency. Cell Stem Cell 14, 762–775 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  14. Dowen, J. M. et al. Control of cell identity genes occurs in insulated neighborhoods in mammalian chromosomes. Cell 159, 374–387 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  15. Zabidi, M. A. et al. Enhancer-core-promoter specificity separates developmental and housekeeping gene regulation. Nature 518, 556–559 (2015).

    Article  CAS  PubMed  Google Scholar 

  16. Goto, T., Macdonald, P. & Maniatis, T. Early and late periodic patterns of even skipped expression are controlled by distinct regulatory elements that respond to different spatial cues. Cell 57, 413–422 (1989).

    Article  CAS  PubMed  Google Scholar 

  17. Harding, K., Hoey, T., Warrior, R. & Levine, M. Autoregulatory and gap gene response elements of the even-skipped promoter of Drosophila. EMBO J. 8, 1205–1212 (1989).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  18. Sharpe, J., Nonchev, S., Gould, A., Whiting, J. & Krumlauf, R. Selectivity, sharing and competitive interactions in the regulation of Hoxb genes. EMBO J. 17, 1788–1798 (1998).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  19. Buecker, C. & Wysocka, J. Enhancers as information integration hubs in development: lessons from genomics. Trends Genet. 28, 276–284 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  20. Perry, M. W., Boettiger, A. N., Bothma, J. P. & Levine, M. Shadow enhancers foster robustness of Drosophila gastrulation. Curr. Biol. 20, 1562–1567 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  21. Bell, O., Tiwari, V. K., Thomä, N. H. & Schübeler, D. Determinants and dynamics of genome accessibility. Nat. Rev. Genet. 12, 554–564 (2011).

    Article  CAS  PubMed  Google Scholar 

  22. Schübeler, D. Function and information content of DNA methylation. Nature 517, 321–326 (2015).

    Article  CAS  PubMed  Google Scholar 

  23. Furey, T. S. ChIP-seq and beyond: new and improved methodologies to detect and characterize protein–DNA interactions. Nat. Rev. Genet. 13, 840–852 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  24. Stormo, G. D. & Zhao, Y. Determining the specificity of protein–DNA interactions. Nat. Rev. Genet. 11, 751–760 (2010).

    Article  CAS  PubMed  Google Scholar 

  25. Levo, M. & Segal, E. In pursuit of design principles of regulatory sequences. Nat. Rev. Genet. 15, 453–468 (2014).

    Article  CAS  PubMed  Google Scholar 

  26. Laird, P. W. Principles and challenges of genomewide DNA methylation analysis. Nat. Rev. Genet. 11, 191–203 (2010).

    Article  CAS  PubMed  Google Scholar 

  27. Belton, J. M. et al. Hi-C: a comprehensive technique to capture the conformation of genomes. Methods 58, 268–276 (2012).

    Article  CAS  PubMed  Google Scholar 

  28. Pennacchio, L. A. et al. In vivo enhancer analysis of human conserved non-coding sequences. Nature 444, 499–502 (2006).

    Article  CAS  PubMed  Google Scholar 

  29. Segal, E., Raveh-Sadka, T., Schroeder, M., Unnerstall, U. & Gaul, U. Predicting expression patterns from regulatory sequence in Drosophila segmentation. Nature 451, 535–540 (2008).

    Article  CAS  PubMed  Google Scholar 

  30. Markstein, M., Markstein, P., Markstein, V. & Levine, M. S. Genome-wide analysis of clustered Dorsal binding sites identifies putative target genes in the Drosophila embryo. Proc. Natl Acad. Sci. USA 99, 763–768 (2002).

    Article  CAS  PubMed  Google Scholar 

  31. Hallikas, O. et al. Genome-wide prediction of mammalian enhancers based on analysis of transcription-factor binding affinity. Cell 124, 47–59 (2006).This study revealed the presence of multiple enhancers regulating the MYC and MYCN oncogenes.

    Article  CAS  PubMed  Google Scholar 

  32. Fullwood, M. J. et al. An oestrogen receptor α-bound human chromatin interactome. Nature 462, 58–64 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  33. Song, L. et al. Open chromatin defined by DNaseI and FAIRE identifies regulatory elements that shape cell-type identity. Genome Res. 21, 1757–1767 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  34. Shlyueva, D., Stampfel, G. & Stark, A. Transcriptional enhancers: from properties to genome-wide predictions. Nat. Rev. Genet. 15, 272–286 (2014).

    Article  CAS  PubMed  Google Scholar 

  35. Stadler, M. B. et al. DNA-binding factors shape the mouse methylome at distal regulatory regions. Nature 480, 490–495 (2011).

    Article  CAS  PubMed  Google Scholar 

  36. Heyn, H. et al. Epigenomic analysis detects aberrant super-enhancer DNA methylation in human cancer. Genome Biol. 17, 11 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  37. Thurman, R. E. et al. The accessible chromatin landscape of the human genome. Nature 489, 75–82 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  38. Hu, G. et al. H2A. Z facilitates access of active and repressive complexes to chromatin in embryonic stem cell self-renewal and differentiation. Cell Stem Cell 12, 180–192 (2013).

    Article  CAS  PubMed  Google Scholar 

  39. Jin, C. et al. H3.3/H2A. Z double variant-containing nucleosomes mark 'nucleosome-free regions' of active promoters and other regulatory regions. Nat. Genet. 41, 941–945 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  40. Yukawa, M. et al. Genome-wide analysis of the chromatin composition of histone H2A and H3 variants in mouse embryonic stem cells. PLoS ONE 9, e92689 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  41. 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 

  42. Creyghton, M. P. et al. Histone H3K27ac separates active from poised enhancers and predicts developmental state. Proc. Natl Acad. Sci. USA 107, 21931–21936 (2010).

    Article  PubMed  PubMed Central  Google Scholar 

  43. Roh, T. Y., Cuddapah, S. & Zhao, K. Active chromatin domains are defined by acetylation islands revealed by genome-wide mapping. Genes Dev. 19, 542–552 (2005).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  44. Seumois, G. et al. Epigenomic analysis of primary human T cells reveals enhancers associated with TH2 memory cell differentiation and asthma susceptibility. Nat. Immunol. 15, 777–788 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  45. Verzi, M. P. et al. Differentiation-specific histone modifications reveal dynamic chromatin interactions and partners for the intestinal transcription factor CDX2. Dev. Cell 19, 713–726 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  46. Andersson, R. et al. An atlas of active enhancers across human cell types and tissues. Nature 507, 455–461 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  47. Lam, M. T., Li, W., Rosenfeld, M. G. & Glass, C. K. Enhancer RNAs and regulated transcriptional programs. Trends Biochem. Sci. 39, 170–182 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  48. Whyte, W. A. et al. Master transcription factors and mediator establish super-enhancers at key cell identity genes. Cell 153, 307–319 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  49. Yan, J. et al. Transcription factor binding in human cells occurs in dense clusters formed around cohesin anchor sites. Cell 154, 801–813 (2013).

    Article  CAS  PubMed  Google Scholar 

  50. Hnisz, D. et al. Convergence of developmental and oncogenic signaling pathways at transcriptional super-enhancers. Mol. Cell 58, 362–370 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  51. Mansour, M. R. et al. Oncogene regulation. An oncogenic super-enhancer formed through somatic mutation of a noncoding intergenic element. Science 346, 1373–1377 (2014).This study shows the de novo generation of a super-enhancer in tumour cells through somatic mutations that introduce TF binding sites.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  52. Zhou, H. Y. et al. A Sox2 distal enhancer cluster regulates embryonic stem cell differentiation potential. Genes Dev. 28, 2699–2711 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  53. Li, Y. et al. CRISPR reveals a distal super-enhancer required for Sox2 expression in mouse embryonic stem cells. PLoS ONE 9, e114485 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  54. Akhtar-Zaidi, B. et al. Epigenomic enhancer profiling defines a signature of colon cancer. Science 336, 736–739 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  55. Aran, D., Sabato, S. & Hellman, A. DNA methylation of distal regulatory sites characterizes dysregulation of cancer genes. Genome Biol. 14, R21 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  56. Lovén, J. et al. Selective inhibition of tumor oncogenes by disruption of super-enhancers. Cell 153, 320–334 (2013).This study indicates that super-enhancers generated de novo in tumour cells can be selectively inhibited by the BET-bromodomain inhibitor JQ1.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

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

    Article  CAS  PubMed  Google Scholar 

  58. Jia, L. et al. Functional enhancers at the gene-poor 8q24 cancer-linked locus. PLoS Genet. 5, e1000597 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  59. Wasserman, N. F., Aneas, I. & Nobrega, M. A. An 8q24 gene desert variant associated with prostate cancer risk confers differential in vivo activity to a MYC enhancer. Genome Res. 20, 1191–1197 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  60. Ahmadiyeh, N. et al. 8q24 prostate, breast, and colon cancer risk loci show tissue-specific long-range interaction with MYC. Proc. Natl Acad. Sci. USA 107, 9742–9746 (2010).This study found that multiple tumour type-specific enhancers are found close to the MYC oncogene.

    Article  PubMed  PubMed Central  Google Scholar 

  61. Tuupanen, S. et al. The common colorectal cancer predisposition SNP rs6983267 at chromosome 8q24 confers potential to enhanced Wnt signaling. Nat. Genet. 41, 885–890 (2009).

    Article  CAS  PubMed  Google Scholar 

  62. Pomerantz, M. M. et al. The 8q24 cancer risk variant rs6983267 shows long-range interaction with MYC in colorectal cancer. Nat. Genet. 41, 882–884 (2009).References 61 and 62 show that SNPs identified by GWAS can alter the affinity of enhancers for oncogenic TFs.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  63. Welter, D. et al. The NHGRI GWAS Catalog, a curated resource of SNP-trait associations. Nucleic Acids Res. 42, D1001–D1006 (2014).

    Article  CAS  PubMed  Google Scholar 

  64. Hindorff, L. A. et al. Potential etiologic and functional implications of genome-wide association loci for human diseases and traits. Proc. Natl Acad. Sci. USA 106, 9362–9367 (2009).

    Article  PubMed  PubMed Central  Google Scholar 

  65. Manolio, T. A. Genomewide association studies and assessment of the risk of disease. N. Engl. J. Med. 363, 166–176 (2010).

    Article  CAS  PubMed  Google Scholar 

  66. Schaub, M. A., Boyle, A. P., Kundaje, A., Batzoglou, S. & Snyder, M. Linking disease associations with regulatory information in the human genome. Genome Res. 22, 1748–1759 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  67. Maurano, M. T. et al. Systematic localization of common disease-associated variation in regulatory DNA. Science 337, 1190–1195 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  68. Nicolae, D. L. et al. Trait-associated SNPs are more likely to be eQTLs: annotation to enhance discovery from GWAS. PLoS Genet. 6, e1000888 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  69. Hulur, I. et al. Enrichment of inflammatory bowel disease and colorectal cancer risk variants in colon expression quantitative trait loci. BMC Genomics 16, 138 (2015).

    Article  PubMed  PubMed Central  Google Scholar 

  70. Li, Q. et al. Integrative eQTL-based analyses reveal the biology of breast cancer risk loci. Cell 152, 633–641 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  71. Chen, Q.-R., Hu, Y., Yan, C., Buetow, K. & Meerzaman, D. Systematic genetic analysis identifies cis-eQTL target genes associated with glioblastoma patient survival. PLoS ONE 9, e105393 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  72. Albert, F. W. & Kruglyak, L. The role of regulatory variation in complex traits and disease. Nat. Rev. Genet. 16, 197–212 (2015).

    Article  CAS  PubMed  Google Scholar 

  73. Fortini, B. K. et al. Multiple functional risk variants in a SMAD7 enhancer implicate a colorectal cancer risk haplotype. PLoS ONE 9, e111914 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  74. Oldridge, D. A. et al. Genetic predisposition to neuroblastoma mediated by a LMO1 super-enhancer polymorphism. Nature 528, 418–421 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  75. Huang, Q. et al. A prostate cancer susceptibility allele at 6q22 increases RFX6 expression by modulating HOXB13 chromatin binding. Nat. Genet. 46, 126–135 (2014).

    Article  CAS  PubMed  Google Scholar 

  76. He, H. et al. Multiple functional variants in long-range enhancer elements contribute to the risk of SNP rs965513 in thyroid cancer. Proc. Natl Acad. Sci. USA 112, 6128–6133 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  77. Dunning, A. M. et al. Breast cancer risk variants at 6q25 display different phenotype associations and regulate ESR1. RMND1 and CCDC170. Nat. Genet. 48, 374–386 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  78. Amundadottir, L. T. et al. A common variant associated with prostate cancer in European and African populations. Nat. Genet. 38, 652–658 (2006).

    Article  CAS  PubMed  Google Scholar 

  79. Gudmundsson, J. et al. Genome-wide association study identifies a second prostate cancer susceptibility variant at 8q24. Nat. Genet. 39, 631–637 (2007).

    Article  CAS  PubMed  Google Scholar 

  80. Yeager, M. et al. Genome-wide association study of prostate cancer identifies a second risk locus at 8q24. Nat. Genet. 39, 645–649 (2007).

    Article  CAS  PubMed  Google Scholar 

  81. Yeager, M. et al. Identification of a new prostate cancer susceptibility locus on chromosome 8q24. Nat. Genet. 41, 1055–1057 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  82. Ghoussaini, M. et al. Multiple loci with different cancer specificities within the 8q24 gene desert. J. Natl Cancer Inst. 100, 962–966 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  83. Crowther-Swanepoel, D. et al. Common variants at 2q37.3, 8q24.21, 15q21.3 and 16q24.1 influence chronic lymphocytic leukemia risk. Nat. Genet. 42, 132–136 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  84. Tomlinson, I. et al. A genome-wide association scan of tag SNPs identifies a susceptibility variant for colorectal cancer at 8q24.21. Nat. Genet. 39, 984–988 (2007).

    Article  CAS  PubMed  Google Scholar 

  85. Zanke, B. W. et al. Genome-wide association scan identifies a colorectal cancer susceptibility locus on chromosome 8q24. Nat. Genet. 39, 989–994 (2007).

    Article  CAS  PubMed  Google Scholar 

  86. Varghese, J. S. & Easton, D. F. Genome-wide association studies in common cancers–what have we learnt? Curr. Opin. Genet. Dev. 20, 201–209 (2010).

    Article  CAS  PubMed  Google Scholar 

  87. Pomerantz, M. M. & Freedman, M. L. The genetics of cancer risk. Cancer J. 17, 416–422 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  88. Sur, I., Tuupanen, S., Whitington, T., Aaltonen, L. A. & Taipale, J. Lessons from functional analysis of genome-wide association studies. Cancer Res. 73, 4180–4184 (2013).

    Article  CAS  PubMed  Google Scholar 

  89. Beroukhim, R. et al. The landscape of somatic copy-number alteration across human cancers. Nature 463, 899–905 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  90. Hsu, P.-Y. et al. Amplification of distant estrogen response elements deregulates target genes associated with tamoxifen resistance in breast cancer. Cancer Cell 24, 197–212 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  91. Tuupanen, S. et al. Allelic imbalance at rs6983267 suggests selection of the risk allele in somatic colorectal tumor evolution. Cancer Res. 68, 14–17 (2008).The first study to show that a chromosomal region carrying a cancer-risk allele identified by GWAS is preferentially somatically amplified in cancer cells.

    Article  CAS  PubMed  Google Scholar 

  92. Sur, I. K. et al. Mice lacking a Myc enhancer that includes human SNP rs6983267 are resistant to intestinal tumors. Science 338, 1360–1363 (2012).This study showed that deletion of an enhancer element carrying a cancer-risk allele has a marked effect on tumour development.

    Article  CAS  PubMed  Google Scholar 

  93. Zhang, X. et al. Identification of focally amplified lineage-specific super-enhancers in human epithelial cancers. Nat. Genet. 48, 176–182 (2016).

    Article  CAS  PubMed  Google Scholar 

  94. Herranz, D. et al. A NOTCH1-driven MYC enhancer promotes T cell development, transformation and acute lymphoblastic leukemia. Nat. Med. 20, 1130–1137 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  95. Cauwelier, B. et al. Molecular cytogenetic study of 126 unselected T-ALL cases reveals high incidence of TCRβ locus rearrangements and putative new T-cell oncogenes. Leukemia 20, 1238–1244 (2006).

    Article  CAS  PubMed  Google Scholar 

  96. Northcott, P. A. et al. Enhancer hijacking activates GFI1 family oncogenes in medulloblastoma. Nature 511, 428–434 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  97. Hnisz, D. et al. Activation of proto-oncogenes by disruption of chromosome neighborhoods. Science 351, 1454–1458 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  98. Alexandrov, L. B. et al. Signatures of mutational processes in human cancer. Nature 500, 415–421 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  99. Katainen, R. et al. CTCF/cohesin-binding sites are frequently mutated in cancer. Nat. Genet. 47, 818–821 (2015).

    Article  CAS  PubMed  Google Scholar 

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

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  101. Puente, X. S. et al. Non-coding recurrent mutations in chronic lymphocytic leukaemia. Nature 526, 519–524 (2015).

    Article  CAS  PubMed  Google Scholar 

  102. Schuster-Böckler, B. & Lehner, B. Chromatin organization is a major influence on regional mutation rates in human cancer cells. Nature 488, 504–507 (2012).

    Article  CAS  PubMed  Google Scholar 

  103. Polak, P. et al. Reduced local mutation density in regulatory DNA of cancer genomes is linked to DNA repair. Nat. Biotechnol. 32, 71–75 (2014).

    Article  CAS  PubMed  Google Scholar 

  104. Polak, P. et al. Cell-of-origin chromatin organization shapes the mutational landscape of cancer. Nature 518, 360–364 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  105. Horn, S. et al. TERT promoter mutations in familial and sporadic melanoma. Science 339, 959–961 (2013).

    Article  CAS  PubMed  Google Scholar 

  106. Huang, F. W. et al. Highly recurrent TERT promoter mutations in human melanoma. Science 339, 957–959 (2013).References 105 and 106 show how somatic mutations in the non-coding regulatory genome alter TF binding sites and cause deregulated expression of TERT in cancer.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  107. Melton, C., Reuter, J. A., Spacek, D. V. & Snyder, M. Recurrent somatic mutations in regulatory regions of human cancer genomes. Nat. Genet. 47, 710–716 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  108. Weinhold, N., Jacobsen, A., Schultz, N., Sander, C. & Lee, W. Genome-wide analysis of noncoding regulatory mutations in cancer. Nat. Genet. 46, 1160–1165 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  109. Kinde, I. et al. TERT promoter mutations occur early in urothelial neoplasia and are biomarkers of early disease and disease recurrence in urine. Cancer Res. 73, 7162–7167 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  110. Nault, J. C. et al. High frequency of telomerase reverse-transcriptase promoter somatic mutations in hepatocellular carcinoma and preneoplastic lesions. Nat. Commun. 4, 2218 (2013).

    Article  CAS  PubMed  Google Scholar 

  111. Mertens, F., Johansson, B., Fioretos, T. & Mitelman, F. The emerging complexity of gene fusions in cancer. Nat. Rev. Cancer 15, 371–381 (2015).

    Article  CAS  PubMed  Google Scholar 

  112. Mitelman, F., Johansson, B. & Mertens, F. The impact of translocations and gene fusions on cancer causation. Nat. Rev. Cancer 7, 233–245 (2007).

    Article  CAS  PubMed  Google Scholar 

  113. Pomerantz, M. M. et al. The androgen receptor cistrome is extensively reprogrammed in human prostate tumorigenesis. Nat. Genet. 47, 1346–1351 (2015).A hallmark study showing how aberrant expression of a lineage-specific TF alters the binding landscape of an oncogenic TF.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  114. Solomon, D. A., Kim, J. S. & Waldman, T. Cohesin gene mutations in tumorigenesis: from discovery to clinical significance. BMB Rep. 47, 299–310 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  115. Plass, C. et al. Mutations in regulators of the epigenome and their connections to global chromatin patterns in cancer. Nat. Rev. Genet. 14, 765–780 (2013).

    Article  CAS  PubMed  Google Scholar 

  116. Bannister, A. J. & Kouzarides, T. Regulation of chromatin by histone modifications. Cell Res. 21, 381–395 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  117. Kagey, M. H. et al. Mediator and cohesin connect gene expression and chromatin architecture. Nature 467, 430–435 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  118. Bhaskara, S. et al. Hdac3 is essential for the maintenance of chromatin structure and genome stability. Cancer Cell 18, 436–447 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  119. Xu, H. et al. Rad21-cohesin haploinsufficiency impedes DNA repair and enhances gastrointestinal radiosensitivity in mice. PLoS ONE 5, e12112 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  120. Yang, L., Rau, R. & Goodell, M. A. DNMT3A in haematological malignancies. Nat. Rev. Cancer 15, 152–165 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  121. Huang, Y. & Rao, A. Connections between TET proteins and aberrant DNA modification in cancer. Trends Genet. 30, 464–474 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  122. Witte, T., Plass, C. & Gerhauser, C. Pan-cancer patterns of DNA methylation. Genome Med. 6, 66 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  123. Lu, C. et al. IDH mutation impairs histone demethylation and results in a block to cell differentiation. Nature 483, 474–478 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  124. Turcan, S. et al. IDH1 mutation is sufficient to establish the glioma hypermethylator phenotype. Nature 483, 479–483 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  125. Figueroa, M. E. et al. Leukemic IDH1 and IDH2 mutations result in a hypermethylation phenotype, disrupt TET2 function, and impair hematopoietic differentiation. Cancer Cell 18, 553–567 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  126. Herman, J. G. et al. Silencing of the VHL tumor-suppressor gene by DNA methylation in renal carcinoma. Proc. Natl Acad. Sci. USA 91, 9700–9704 (1994).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  127. Esteller, M. et al. Promoter hypermethylation and BRCA1 inactivation in sporadic breast and ovarian tumors. J. Natl Cancer Inst. 92, 564–569 (2000).

    Article  CAS  PubMed  Google Scholar 

  128. Blattler, A. & Farnham, P. J. Cross-talk between site-specific transcription factors and DNA methylation states. J. Biol. Chem. 288, 34287–34294 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  129. Flavahan, W. A. et al. Insulator dysfunction and oncogene activation in IDH mutant gliomas. Nature 529, 110–114 (2016).

    Article  CAS  PubMed  Google Scholar 

  130. Toyota, M. et al. CpG island methylator phenotype in colorectal cancer. Proc. Natl Acad. Sci. USA 96, 8681–8686 (1999).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  131. Webster, D. E. et al. Enhancer-targeted genome editing selectively blocks innate resistance to oncokinase inhibition. Genome Res. 24, 751–760 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  132. Tak, Y. G. et al. Effects on the transcriptome upon deletion of a distal element cannot be predicted by the size of the H3K27Ac peak in human cells. Nucleic Acids Res. 44, 4123–4133 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  133. Diffner, E. et al. Activity of a heptad of transcription factors is associated with stem cell programs and clinical outcome in acute myeloid leukemia. Blood 121, 2289–2300 (2013).

    Article  CAS  PubMed  Google Scholar 

  134. Lehmann-Werman, R. et al. Identification of tissue-specific cell death using methylation patterns of circulating DNA. Proc. Natl Acad. Sci. USA 113, E1826–E1834 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  135. Sun, K. et al. Plasma DNA tissue mapping by genome-wide methylation sequencing for noninvasive prenatal, cancer, and transplantation assessments. Proc. Natl Acad. Sci. USA 112, E5503–E5512 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  136. Snyder, M. W., Kircher, M., Hill, A. J., Daza, R. M. & Shendure, J. Cell-free DNA comprises an in vivo nucleosome footprint that informs its tissues-of-origin. Cell 164, 57–68 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  137. Chipumuro, E. et al. CDK7 inhibition suppresses super-enhancer-linked oncogenic transcription in MYCN-driven cancer. Cell 159, 1126–1139 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  138. Delmore, J. E. et al. BET bromodomain inhibition as a therapeutic strategy to target c-Myc. Cell 146, 904–917 (2011).This report showed the efficacy of JQ1 in inhibiting MYC transcription.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  139. Filippakopoulos, P. et al. Selective inhibition of BET bromodomains. Nature 468, 1067–1073 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  140. West, A. C. & Johnstone, R. W. New and emerging HDAC inhibitors for cancer treatment. J. Clin. Invest. 124, 30–39 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  141. Richon, V. M., Sandhoff, T. W., Rifkind, R. A. & Marks, P. A. Histone deacetylase inhibitor selectively induces p21WAF1 expression and gene-associated histone acetylation. Proc. Natl Acad. Sci. USA 97, 10014–10019 (2000).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  142. Insinga, A. et al. Inhibitors of histone deacetylases induce tumor-selective apoptosis through activation of the death receptor pathway. Nat. Med. 11, 71–76 (2005).

    Article  CAS  PubMed  Google Scholar 

  143. Rhodes, J. M. et al. Positive regulation of c-Myc by cohesin is direct, and evolutionarily conserved. Dev. Biol. 344, 637–649 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  144. McEwan, M. V., Eccles, M. R. & Horsfield, J. A. Cohesin is required for activation of MYC by estradiol. PLoS ONE 7, e49160 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  145. Pelish, H. E. et al. Mediator kinase inhibition further activates super-enhancer-associated genes in AML. Nature 526, 273–276 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  146. Nebbioso, A. et al. Tumor-selective action of HDAC inhibitors involves TRAIL induction in acute myeloid leukemia cells. Nat. Med. 11, 77–84 (2005).

    Article  CAS  PubMed  Google Scholar 

  147. Xu, W. S., Perez, G., Ngo, L., Gui, C. Y. & Marks, P. A. Induction of polyploidy by histone deacetylase inhibitor: a pathway for antitumor effects. Cancer Res. 65, 7832–7839 (2005).

    Article  CAS  PubMed  Google Scholar 

  148. Gui, C. Y., Ngo, L., Xu, W. S., Richon, V. M. & Marks, P. A. Histone deacetylase (HDAC) inhibitor activation of p21WAF1 involves changes in promoter-associated proteins, including HDAC1. Proc. Natl Acad. Sci. USA 101, 1241–1246 (2004).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  149. Rathert, P. et al. Transcriptional plasticity promotes primary and acquired resistance to BET inhibition. Nature 525, 543–547 (2015).This paper shows that changes in enhancer usage can cause resistance to JQ1-mediated inhibition of MYC expression.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  150. Kumar, K. et al. GLI2-dependent c-MYC upregulation mediates resistance of pancreatic cancer cells to the BET bromodomain inhibitor JQ1. Sci. Rep. 5, 9489 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  151. Fong, C. Y. et al. BET inhibitor resistance emerges from leukaemia stem cells. Nature 525, 538–542 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  152. Holohan, C., Van Schaeybroeck, S., Longley, D. B. & Johnston, P. G. Cancer drug resistance: an evolving paradigm. Nat. Rev. Cancer 13, 714–726 (2013).

    Article  CAS  PubMed  Google Scholar 

  153. Huang, M., Shen, A., Ding, J. & Geng, M. Molecularly targeted cancer therapy: some lessons from the past decade. Trends Pharmacol. Sci. 35, 41–50 (2014).

    Article  CAS  PubMed  Google Scholar 

  154. Kress, T. R., Sabo, A. & Amati, B. MYC: connecting selective transcriptional control to global RNA production. Nat. Rev. Cancer 15, 593–607 (2015).

    Article  CAS  PubMed  Google Scholar 

  155. Bolden, J. E. et al. Inducible in vivo silencing of Brd4 identifies potential toxicities of sustained BET protein inhibition. Cell Rep. 8, 1919–1929 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  156. Lettice, L. A. et al. A long-range Shh enhancer regulates expression in the developing limb and fin and is associated with preaxial polydactyly. Hum. Mol. Genet. 12, 1725–1735 (2003).

    Article  CAS  PubMed  Google Scholar 

  157. ENCODE Project Consortium. An integrated encyclopedia of DNA elements in the human genome. Nature 489, 57–74 (2012).

  158. Heintzman, N. D. et al. Distinct and predictive chromatin signatures of transcriptional promoters and enhancers in the human genome. Nat. Genet. 39, 311–318 (2007).

    Article  CAS  PubMed  Google Scholar 

  159. Ong, C.-T. & Corces, V. G. CTCF: an architectural protein bridging genome topology and function. Nat. Rev. Genet. 15, 234–246 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  160. Dixon, J. R. et al. Topological domains in mammalian genomes identified by analysis of chromatin interactions. Nature 485, 376–380 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

Download references

Acknowledgements

The authors wish to thank M. Taipale and E. Kaasinen for careful reading of the manuscript and helpful comments.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Jussi Taipale.

Ethics declarations

Competing interests

The authors declare no competing financial interests.

PowerPoint slides

Glossary

Variant histones

Non-allelic variant forms of the canonical histone proteins, such as H2AX and H3.3, that can be incorporated into nucleosomes in a replication-independent manner and serve to regulate multiple processes, including transcription and DNA repair.

Expression quantitative trait loci

(eQTLs). Regions of the genome that account for quantitative variability in the expression of a given gene. These can either be linked (in cis) or unlinked (in trans) to the target gene. They are mapped by measuring the expression levels of genes and correlating them with the individual genotypes.

DNase I hypersensitive sites

Active regions of the genome that are depleted of nucleosomes. The open chromatin conformation of these regions makes them hypersensitive to digestion by the enzyme DNase I.

Effect sizes

Quantitative measures of the magnitude of a given effect. For example, in genome-wide association studies (GWAS), the effect size is given as an odds ratio, describing the relative increase in cancer risk of individuals carrying the risk variant.

Odds ratio

The association between the presence or absence of a variant (for example, a particular allele of a single nucleotide polymorphism) and the presence or absence of a trait (for example, cancer) in the population. The odds of occurrence of the variant is determined in groups of subjects with and without the trait; the odds ratio is the ratio of the odds of occurrence of the variant in people with the trait to the odds of occurrence of the variant in people without the trait.

tag SNPs

Genotyped single nucleotide polymorphisms (SNPs) that are used to predict the genotypes of other nearby SNPs. The SNP that contributes to cancer predisposition can either be the tag SNP itself, or more commonly, another SNP that is in linkage disequilibrium with the tag SNP.

Insulated chromosome neighbourhoods

Large DNA loops containing genes and enhancers, in which the expression of the genes is controlled by the enhancers within the loop, but is insulated from influence by enhancers residing outside the loop. It is thought that these loops are formed by an interaction between two CCCTC-binding factor (CTCF) and cohesin proteins bound at the loop boundaries.

Microsatellite instability

Microsatellite DNA corresponds to short stretches of highly repetitive 1–5 bp sequences. Some cancer cells display high mutation rates at such microsatellite DNA repeats. This microsatellite instability is a consequence of mutational or epigenetic inactivation of genes involved in DNA mismatch repair.

Cyclin-dependent kinase 7

(CDK7). Part of the multi-protein transcription factor IIH complex that is required for general transcription. Within this complex, CDK7 phosphorylates the carboxy-terminal domain of RNA polymerase II.

BET-bromodomain protein

Member of a protein family that contains conserved bromodomains that recognize acetyl-lysine on proteins, including histones. Within this group, bromodomain protein 2 (BRD2), BRD3, BRD4 and BRDT constitute a family of bromodomain and extra-terminal (BET) proteins that regulate transcription.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Sur, I., Taipale, J. The role of enhancers in cancer. Nat Rev Cancer 16, 483–493 (2016). https://doi.org/10.1038/nrc.2016.62

Download citation

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/nrc.2016.62

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