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 selection and function of cell type-specific enhancers

Key Points

  • Enhancers are genetic elements that have major roles in determining cell type-specific gene expression patterns and responses to internal and external signals.

  • Mammalian genomes contain millions of enhancers, but only a subset of them are selected and activated in each cell type in the body.

  • Enhancer selection involves collaborative interactions between 'pioneer' or lineage-determining transcription factors. Such factors are able to prime enhancers in a cell type-specific manner by binding to closely spaced recognition motifs.

  • Signal-dependent transcription factors, such as nuclear receptors and nuclear factor-κB (NF-κB), primarily bind to regions of the genome that are primed in a cell type-specific manner by lineage-determining transcription factors. This enables such broadly expressed, signal-dependent transcription factors to regulate gene expression in a cell type-specific manner.

  • A proportion of the genome contains a very high density of marks of active enhancers. Such enhancer-rich genomic regions, which are called super-enhancers, are different in each cell type and regulate genes required for establishing cell identity and function.

  • An understanding of the mechanisms underlying the cell type-specific selection and function of enhancers will improve our understanding of the effects of natural genetic variation on complex phenotypes and diseases.

Abstract

The human body contains several hundred cell types, all of which share the same genome. In metazoans, much of the regulatory code that drives cell type-specific gene expression is located in distal elements called enhancers. Although mammalian genomes contain millions of potential enhancers, only a small subset of them is active in a given cell type. Cell type-specific enhancer selection involves the binding of lineage-determining transcription factors that prime enhancers. Signal-dependent transcription factors bind to primed enhancers, which enables these broadly expressed factors to regulate gene expression in a cell type-specific manner. The expression of genes that specify cell type identity and function is associated with densely spaced clusters of active enhancers known as super-enhancers. The functions of enhancers and super-enhancers are influenced by, and affect, higher-order genomic organization.

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: The anatomies of poised and active enhancers.
Figure 2: Cell type-specific enhancers are marked by specific epigenomic features and chromatin accessibility.
Figure 3: Cell type-specific enhancer selection and activation.
Figure 4: Enhancer activation and function.
Figure 5: The linear and 3D organization of enhancers in the nucleus.

Similar content being viewed by others

References

  1. Banerji, J., Rusconi, S. & Schaffner, W. Expression of a β-globin gene is enhanced by remote SV40 DNA sequences. Cell 27, 299–308 (1981).

    Article  CAS  PubMed  Google Scholar 

  2. 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).

    CAS  PubMed  Google Scholar 

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

    CAS  PubMed  Google Scholar 

  4. Carroll, J. S. et al. Genome-wide analysis of estrogen receptor binding sites. Nature Genet. 38, 1289–1297 (2006).

    CAS  PubMed  Google Scholar 

  5. Barish, G. D. et al. Bcl-6 and NF-κB cistromes mediate opposing regulation of the innate immune response. Genes Dev. 24, 2760–2765 (2010).

    CAS  PubMed  PubMed Central  Google Scholar 

  6. John, S. et al. Chromatin accessibility pre-determines glucocorticoid receptor binding patterns. Nature Genet. 43, 264–268 (2011).

    CAS  PubMed  Google Scholar 

  7. Heinz, S. et al. Simple combinations of lineage-determining transcription factors prime cis-regulatory elements required for macrophage and B cell identities. Mol. Cell 38, 576–589 (2010). This paper shows that the LDTF PU.1 binds to different loci in different cell types. The loci are marked by H3K4me1, act as beacons for SDTFs binding and drive cell type-specific cellular responses.

    CAS  PubMed  PubMed Central  Google Scholar 

  8. Lefterova, M. I. et al. Cell-specific determinants of peroxisome proliferator-activated receptor γ function in adipocytes and macrophages. Mol. Cell. Biol. 30, 2078–2089 (2010).

    CAS  PubMed  PubMed Central  Google Scholar 

  9. Nielsen, R. et al. Genome-wide profiling of PPARγ:RXR and RNA polymerase II occupancy reveals temporal activation of distinct metabolic pathways and changes in RXR dimer composition during adipogenesis. Genes Dev. 22, 2953–2967 (2008).

    CAS  PubMed  PubMed Central  Google Scholar 

  10. Ghisletti, S. et al. Identification and characterization of enhancers controlling the inflammatory gene expression program in macrophages. Immunity 32, 317–328 (2010). This paper shows that PU.1 is an LDTF necessary for priming signal-dependent enhancers in macrophages, which defines their transcriptional response to inflammatory stimuli.

    CAS  PubMed  Google Scholar 

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

    CAS  PubMed  Google Scholar 

  12. Woolfe, A. et al. Highly conserved non-coding sequences are associated with vertebrate development. PLoS Biol. 3, e7 (2005).

    PubMed  Google Scholar 

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

    CAS  PubMed  PubMed Central  Google Scholar 

  14. Visel, A. et al. ChIP–seq accurately predicts tissue-specific activity of enhancers. Nature 457, 854–858 (2009).

    CAS  PubMed  PubMed Central  Google Scholar 

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

    CAS  PubMed  PubMed Central  Google Scholar 

  16. Bernstein, B. E. et al. An integrated encyclopedia of DNA elements in the human genome. Nature 489, 57–74 (2012). This paper summarizes the work of the ENCODE consortium to annotate functional DNA elements in the human genome.

    Google Scholar 

  17. West, J. A. et al. Nucleosomal occupancy changes locally over key regulatory regions during cell differentiation and reprogramming. Nature Commun. 5, 4719 (2014).

    CAS  Google Scholar 

  18. Chan, H. M. & La Thangue, N. B. p300/CBP proteins: HATs for transcriptional bridges and scaffolds. J. Cell Sci. 114, 2363–2373 (2001).

    CAS  PubMed  Google Scholar 

  19. De Santa, F. et al. A large fraction of extragenic RNA Pol II transcription sites overlap enhancers. PLoS Biol. 8, e1000384 (2010).

    PubMed  PubMed Central  Google Scholar 

  20. Kim, T.-K. et al. Widespread transcription at neuronal activity-regulated enhancers. Nature 465, 182–187 (2010). References 19 and 20 report the existence of widespread transcription at enhancers, which correlates with the transcription of neighbouring genes.

    CAS  PubMed  PubMed Central  Google Scholar 

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

    CAS  PubMed  Google Scholar 

  22. Ernst, J. & Kellis, M. Discovery and characterization of chromatin states for systematic annotation of the human genome. Nature Biotech. 28, 817–825 (2010).

    CAS  Google Scholar 

  23. He, H. H. et al. Nucleosome dynamics define transcriptional enhancers. Nature Genet. 42, 343–347 (2010).

    CAS  PubMed  Google Scholar 

  24. Rada-Iglesias, A. et al. A unique chromatin signature uncovers early developmental enhancers in humans. Nature 470, 279–283 (2011).

    CAS  PubMed  Google Scholar 

  25. 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).

    CAS  PubMed  PubMed Central  Google Scholar 

  26. Zentner, G. E., Tesar, P. J. & Scacheri, P. C. Epigenetic signatures distinguish multiple classes of enhancers with distinct cellular functions. Genome Res. 21, 1273–1283 (2011).

    CAS  PubMed  PubMed Central  Google Scholar 

  27. Calo, E. & Wysocka, J. Modification of enhancer chromatin: what, how, and why? Mol. Cell 49, 825–837 (2013).

    CAS  PubMed  Google Scholar 

  28. Gottgens, B. et al. The scl +18/19 stem cell enhancer is not required for hematopoiesis: identification of a 5′ bifunctional hematopoietic-endothelial enhancer bound by Fli-1 and Elf-1. Mol. Cell. Biol. 24, 1870–1883 (2004).

    PubMed  PubMed Central  Google Scholar 

  29. Sanchez, M. et al. An SCL 3′ enhancer targets developing endothelium together with embryonic and adult haematopoietic progenitors. Development 126, 3891–3904 (1999).

    CAS  PubMed  Google Scholar 

  30. Delabesse, E. et al. Transcriptional regulation of the SCL locus: identification of an enhancer that targets the primitive erythroid lineage in vivo. Mol. Cell. Biol. 25, 5215–5225 (2005).

    CAS  PubMed  PubMed Central  Google Scholar 

  31. Zaret, K. S. et al. Pioneer factors, genetic competence, and inductive signaling: programming liver and pancreas progenitors from the endoderm. Cold Spring Harb. Symp. Quant. Biol. 73, 119–126 (2008).

    CAS  PubMed  PubMed Central  Google Scholar 

  32. Pham, T. H. et al. Mechanisms of in vivo binding site selection of the hematopoietic master transcription factor PU.1. Nucleic Acids Res. 41, 6391–6402 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  33. Adams, C. C. & Workman, J. L. Binding of disparate transcriptional activators to nucleosomal DNA is inherently cooperative. Mol. Cell. Biol. 15, 1405–1421 (1995).

    CAS  PubMed  PubMed Central  Google Scholar 

  34. Boyes, J. & Felsenfeld, G. Tissue-specific factors additively increase the probability of the all-or-none formation of a hypersensitive site. EMBO J. 15, 2496–2507 (1996).

    CAS  PubMed  PubMed Central  Google Scholar 

  35. Hoffman, B. G. et al. Locus co-occupancy, nucleosome positioning, and H3K4me1 regulate the functionality of FOXA2-, HNF4A-, and PDX1-bound loci in islets and liver. Genome Res. 20, 1037–1051 (2010).

    CAS  PubMed  PubMed Central  Google Scholar 

  36. Samstein, R. M. et al. Foxp3 exploits a pre-existent enhancer landscape for regulatory T cell lineage specification. Cell 151, 153–166 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  37. Shlyueva, D. et al. Hormone-responsive enhancer-activity maps reveal predictive motifs, indirect repression, and targeting of closed chromatin. Mol. Cell 54, 180–192 (2014).

    CAS  PubMed  Google Scholar 

  38. Vahedi, G. et al. STATs shape the active enhancer landscape of T cell populations. Cell 151, 981–993 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  39. Xu, J. et al. Combinatorial assembly of developmental stage-specific enhancers controls gene expression programs during human erythropoiesis. Dev. Cell. 23, 796–811 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  40. Soufi, A., Donahue, G. & Zaret, K. S. Facilitators and impediments of the pluripotency reprogramming factors' initial engagement with the genome. Cell 151, 994–1004 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  41. Scott, E. W., Simon, M. C., Anastasi, J. & Singh, H. Requirement of transcription factor PU.1 in the development of multiple hematopoietic lineages. Science 265, 1573–1577 (1994).

    CAS  PubMed  Google Scholar 

  42. Kazemian, M., Pham, H., Wolfe, S. A., Brodsky, M. H. & Sinha, S. Widespread evidence of cooperative DNA binding by transcription factors in Drosophila development. Nucleic Acids Res. 41, 8237–8252 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  43. Heinz, S. et al. Effect of natural genetic variation on enhancer selection and function. Nature 503, 487–492 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  44. Stefflova, K. et al. Cooperativity and rapid evolution of cobound transcription factors in closely related mammals. Cell 154, 530–540 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  45. Trompouki, E. et al. Lineage regulators direct BMP and Wnt pathways to cell-specific programs during differentiation and regeneration. Cell 147, 577–589 (2011).

    CAS  PubMed  PubMed Central  Google Scholar 

  46. Yanez-Cuna, J. O., Dinh, H. Q., Kvon, E. Z., Shlyueva, D. & Stark, A. Uncovering cis-regulatory sequence requirements for context-specific transcription factor binding. Genome Res. 22, 2018–2030 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  47. Mercer, E. M. et al. Multilineage priming of enhancer repertoires precedes commitment to the B and myeloid cell lineages in hematopoietic progenitors. Immunity 35, 413–425 (2011).

    CAS  PubMed  PubMed Central  Google Scholar 

  48. Mullen, A. C. et al. Master transcription factors determine cell-type-specific responses to TGF-β signaling. Cell 147, 565–576 (2011). References 36 and 48 show that SDTFs bind to open chromatin regions that are defined by combinations of LDTFs.

    CAS  PubMed  PubMed Central  Google Scholar 

  49. Carroll, J. S. et al. Chromosome-wide mapping of estrogen receptor binding reveals long-range regulation requiring the forkhead protein FoxA1. Cell 122, 33–43 (2005).

    CAS  PubMed  Google Scholar 

  50. Kaikkonen, M. U. et al. Remodeling of the enhancer landscape during macrophage activation is coupled to enhancer transcription. Mol. Cell 51, 310–325 (2013). This paper shows that the SDTF NF-κB contributes to the de novo priming (and activation) of enhancers in conjunction with LDTFs, and that transcription elongation is required for the methylation of the flanking H3K4 by MLL3 and MLL4.

    CAS  PubMed  PubMed Central  Google Scholar 

  51. Sullivan, A. L. et al. Serum response factor utilizes distinct promoter- and enhancer-based mechanisms to regulate cytoskeletal gene expression in macrophages. Mol. Cell. Biol. 31, 861–875 (2011).

    CAS  PubMed  Google Scholar 

  52. Ostuni, R. et al. Latent enhancers activated by stimulation in differentiated cells. Cell 152, 157–171 (2013). This paper describes signal-dependent selection of latent enhancers in macrophages and their persistence after signal termination as a molecular memory of prior activation.

    CAS  PubMed  Google Scholar 

  53. Voss, T. C. et al. Dynamic exchange at regulatory elements during chromatin remodeling underlies assisted loading mechanism. Cell 146, 544–554 (2011).

    CAS  PubMed  PubMed Central  Google Scholar 

  54. Gosselin, D. et al. Environment drives selection and function of enhancers controlling tissue-specific macrophage identities. Cell 159, 1327–1340 (2014).

    CAS  PubMed  PubMed Central  Google Scholar 

  55. Lavin, Y. et al. Tissue-resident macrophage enhancer landscapes are shaped by the local microenvironment. Cell 159, 1312–1326 (2014). References 54 and 55 demonstrate how specific tissue environments differentially influence the selection and function of enhancers in macrophage populations.

    CAS  PubMed  PubMed Central  Google Scholar 

  56. Herz, H. M., Hu, D. & Shilatifard, A. Enhancer malfunction in cancer. Mol. Cell 53, 859–866 (2014).

    CAS  PubMed  PubMed Central  Google Scholar 

  57. Wang, Z. et al. Genome-wide mapping of HATs and HDACs reveals distinct functions in active and inactive genes. Cell 138, 1019–1031 (2009).

    CAS  PubMed  PubMed Central  Google Scholar 

  58. Euskirchen, G. M. et al. Diverse roles and interactions of the SWI/SNF chromatin remodeling complex revealed using global approaches. PLoS Genet. 7, e1002008 (2011).

    CAS  PubMed  PubMed Central  Google Scholar 

  59. Morris, S. A. et al. Overlapping chromatin-remodeling systems collaborate genome wide at dynamic chromatin transitions. Nature Struct. Mol. Biol. 21, 73–81 (2014).

    CAS  Google Scholar 

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

    CAS  PubMed  PubMed Central  Google Scholar 

  61. Siersbaek, R. et al. Transcription factor cooperativity in early adipogenic hotspots and super-enhancers. Cell Rep. 7, 1443–1455 (2014).

    CAS  PubMed  Google Scholar 

  62. Rosenfeld, M. G., Lunyak, V. V. & Glass, C. K. Sensors and signals: a coactivator/corepressor/epigenetic code for integrating signal-dependent programs of transcriptional response. Genes Dev. 20, 1405–1428 (2006).

    CAS  PubMed  Google Scholar 

  63. Blazek, E., Mittler, G. & Meisterernst, M. The mediator of RNA polymerase II. Chromosoma 113, 399–408 (2005).

    CAS  PubMed  Google Scholar 

  64. Malik, S. & Roeder, R. G. The metazoan Mediator co-activator complex as an integrative hub for transcriptional regulation. Nature Rev. Genet. 11, 761–772 (2010).

    CAS  PubMed  Google Scholar 

  65. Vermeulen, M. et al. Selective anchoring of TFIID to nucleosomes by trimethylation of histone H3 lysine 4. Cell 131, 58–69 (2007).

    CAS  PubMed  Google Scholar 

  66. Dey, A., Chitsaz, F., Abbasi, A., Misteli, T. & Ozato, K. The double bromodomain protein Brd4 binds to acetylated chromatin during interphase and mitosis. Proc. Natl Acad. Sci. USA 100, 8758–8763 (2003).

    CAS  PubMed  PubMed Central  Google Scholar 

  67. Koch, F. et al. Transcription initiation platforms and GTF recruitment at tissue-specific enhancers and promoters. Nature Struct. Mol. Biol. 18, 956–963 (2011).

    CAS  Google Scholar 

  68. Zhang, W. et al. Bromodomain-containing protein 4 (BRD4) regulates RNA polymerase II serine 2 phosphorylation in human CD4+ T cells. J. Biol. Chem. 287, 43137–43155 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  69. Collis, P., Antoniou, M. & Grosveld, F. Definition of the minimal requirements within the human β-globin gene and the dominant control region for high level expression. EMBO J. 9, 233–240 (1990).

    CAS  PubMed  PubMed Central  Google Scholar 

  70. Lam, M. T. et al. Rev–Erbs repress macrophage gene expression by inhibiting enhancer-directed transcription. Nature 498, 511–515 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

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

    CAS  PubMed  PubMed Central  Google Scholar 

  72. Core, L. J. et al. Analysis of nascent RNA identifies a unified architecture of initiation regions at mammalian promoters and enhancers. Nature Genet. 46, 1311–1320 (2014).

    CAS  PubMed  Google Scholar 

  73. Wang, D. et al. Reprogramming transcription by distinct classes of enhancers functionally defined by eRNA. Nature 474, 390–394 (2011).

    CAS  PubMed  PubMed Central  Google Scholar 

  74. Kieffer-Kwon, K. R. et al. Interactome maps of mouse gene regulatory domains reveal basic principles of transcriptional regulation. Cell 155, 1507–1520 (2013).

    CAS  PubMed  Google Scholar 

  75. Bonn, S. et al. Tissue-specific analysis of chromatin state identifies temporal signatures of enhancer activity during embryonic development. Nature Genet. 44, 148–156 (2012).

    CAS  PubMed  Google Scholar 

  76. Almada, A. E., Wu, X., Kriz, A. J., Burge, C. B. & Sharp, P. A. Promoter directionality is controlled by U1 snRNP and polyadenylation signals. Nature 499, 360–363 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  77. Fong, Y. W. & Zhou, Q. Stimulatory effect of splicing factors on transcriptional elongation. Nature 414, 929–933 (2001).

    CAS  PubMed  Google Scholar 

  78. Muller-McNicoll, M. & Neugebauer, K. M. How cells get the message: dynamic assembly and function of mRNA-protein complexes. Nature Rev. Genet. 14, 275–287 (2013).

    PubMed  Google Scholar 

  79. Bieberstein, N. I., Carrillo Oesterreich, F., Straube, K. & Neugebauer, K. M. First exon length controls active chromatin signatures and transcription. Cell Rep. 2, 62–68 (2012).

    CAS  PubMed  Google Scholar 

  80. Kowalczyk, M. S. et al. Intragenic enhancers act as alternative promoters. Mol. Cell 45, 447–458 (2012).

    CAS  PubMed  Google Scholar 

  81. Derrien, T. et al. The GENCODE v7 catalog of human long noncoding RNAs: analysis of their gene structure, evolution, and expression. Genome Res. 22, 1775–1789 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  82. Schaukowitch, K. et al. Enhancer RNA facilitates NELF release from immediate early genes. Mol. Cell 56, 29–42 (2014).

    CAS  PubMed  PubMed Central  Google Scholar 

  83. Core, L. J., Waterfall, J. J. & Lis, J. T. Nascent RNA sequencing reveals widespread pausing and divergent initiation at human promoters. Science 322, 1845–1848 (2008).

    CAS  PubMed  PubMed Central  Google Scholar 

  84. Herz, H. M. et al. Enhancer-associated H3K4 monomethylation by Trithorax-related, the Drosophila homolog of mammalian Mll3/Mll4. Genes Dev. 26, 2604–2620 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  85. Lee, J. E. et al. H3K4 mono- and di-methyltransferase MLL4 is required for enhancer activation during cell differentiation. Elife 2, e01503 (2013).

    PubMed  PubMed Central  Google Scholar 

  86. Sims, R. J. 3rd & Reinberg, D. Histone H3 Lys 4 methylation: caught in a bind? Genes Dev. 20, 2779–2786 (2006).

    CAS  PubMed  Google Scholar 

  87. Plank, J. L. & Dean, A. Enhancer function: mechanistic and genome-wide insights come together. Mol. Cell 55, 5–14 (2014).

    CAS  PubMed  PubMed Central  Google Scholar 

  88. Schmidt, D. et al. A CTCF-independent role for cohesin in tissue-specific transcription. Genome Res. 20, 578–588 (2010).

    CAS  PubMed  PubMed Central  Google Scholar 

  89. Liu, W. et al. Brd4 and JMJD6-associated anti-pause enhancers in regulation of transcriptional pause release. Cell 155, 1581–1595 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  90. Nord, A. S. et al. Rapid and pervasive changes in genome-wide enhancer usage during mammalian development. Cell 155, 1521–1531 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

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

    CAS  PubMed  Google Scholar 

  92. Whyte, W. A. et al. Master transcription factors and mediator establish super-enhancers at key cell identity genes. Cell 153, 307–319 (2013). References 91 and 92 describe genomic regions near genes, which are highly enriched for marks of active enhancers and have essential roles in determining cell identity and function.

    CAS  PubMed  PubMed Central  Google Scholar 

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

    CAS  PubMed  PubMed Central  Google Scholar 

  94. Parker, S. C. et al. Chromatin stretch enhancer states drive cell-specific gene regulation and harbor human disease risk variants. Proc. Natl Acad. Sci. USA 110, 17921–17926 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  95. Cremer, T. et al. Chromosome territories — a functional nuclear landscape. Curr. Opin. Cell Biol. 18, 307–316 (2006).

    CAS  PubMed  Google Scholar 

  96. Lieberman-Aiden, E. et al. Comprehensive mapping of long-range interactions reveals folding principles of the human genome. Science 326, 289–293 (2009).

    CAS  PubMed  PubMed Central  Google Scholar 

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

    CAS  PubMed  PubMed Central  Google Scholar 

  98. Sanyal, A., Lajoie, B. R., Jain, G. & Dekker, J. The long-range interaction landscape of gene promoters. Nature 489, 109–113 (2012). References 97 and 98 use global 3C assays to interrogate the 3D organization of functional elements within the genome.

    CAS  PubMed  PubMed Central  Google Scholar 

  99. de Laat, W. & Grosveld, F. Spatial organization of gene expression: the active chromatin hub. Chromosome Res. 11, 447–459 (2003).

    CAS  PubMed  Google Scholar 

  100. Jin, F. et al. A high-resolution map of the three-dimensional chromatin interactome in human cells. Nature 503, 290–294 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  101. Ghavi-Helm, Y. et al. Enhancer loops appear stable during development and are associated with paused polymerase. Nature 512, 96–100 (2014).

    CAS  PubMed  Google Scholar 

  102. Vassetzky, Y., Hair, A. & Mechali, M. Rearrangement of chromatin domains during development in Xenopus. Genes Dev. 14, 1541–1552 (2000).

    CAS  PubMed  PubMed Central  Google Scholar 

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

    CAS  PubMed  Google Scholar 

  104. Nagano, T. et al. Single-cell Hi-C reveals cell-to-cell variability in chromosome structure. Nature 502, 59–64 (2013).

    CAS  PubMed  Google Scholar 

  105. Li, G. et al. Extensive promoter-centered chromatin interactions provide a topological basis for transcription regulation. Cell 148, 84–98 (2012). This study shows that promoters can function as enhancers.

    CAS  PubMed  PubMed Central  Google Scholar 

  106. Li, W. et al. Functional roles of enhancer RNAs for oestrogen-dependent transcriptional activation. Nature 498, 516–520 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  107. Bender, M. A. et al. The hypersensitive sites of the murine β-globin locus control region act independently to affect nuclear localization and transcriptional elongation. Blood 119, 3820–3827 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  108. 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).

    CAS  PubMed  Google Scholar 

  109. Rosenbauer, F. et al. Acute myeloid leukemia induced by graded reduction of a lineage-specific transcription factor, PU.1. Nature Genet. 36, 624–630 (2004).

    CAS  PubMed  Google Scholar 

  110. Hong, J. W., Hendrix, D. A. & Levine, M. S. Shadow enhancers as a source of evolutionary novelty. Science 321, 1314 (2008).

    CAS  PubMed  PubMed Central  Google Scholar 

  111. Barolo, S. Shadow enhancers: frequently asked questions about distributed cis-regulatory information and enhancer redundancy. Bioessays 34, 135–141 (2012).

    CAS  PubMed  Google Scholar 

  112. Lin, Y. C. et al. Global changes in the nuclear positioning of genes and intra- and interdomain genomic interactions that orchestrate B cell fate. Nature Immunol. 13, 1196–1204 (2012).

    CAS  Google Scholar 

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

    CAS  PubMed  Google Scholar 

  114. Pasquali, L. et al. Pancreatic islet enhancer clusters enriched in type 2 diabetes risk-associated variants. Nature Genet. 46, 136–143 (2014).

    CAS  PubMed  Google Scholar 

  115. 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).

    CAS  PubMed  PubMed Central  Google Scholar 

  116. Trynka, G. et al. Chromatin marks identify critical cell types for fine mapping complex trait variants. Nature Genet. 45, 124–130 (2013).

    CAS  PubMed  Google Scholar 

  117. Raj, T. et al. Polarization of the effects of autoimmune and neurodegenerative risk alleles in leukocytes. Science 344, 519–523 (2014).

    CAS  PubMed  PubMed Central  Google Scholar 

  118. Kasowski, M. et al. Extensive variation in chromatin states across humans. Science 342, 750–752 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  119. Kilpinen, H. et al. Coordinated effects of sequence variation on DNA binding, chromatin structure, and transcription. Science 342, 744–747 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  120. McVicker, G. et al. Identification of genetic variants that affect histone modifications in human cells. Science 342, 747–749 (2013). References 118–120 report on the effects of natural genetic variation in humans on the binding of transcription factors and histone modifications that are associated with enhancers and gene expression.

    CAS  PubMed  PubMed Central  Google Scholar 

  121. Degner, J. F. et al. DNase I sensitivity QTLs are a major determinant of human expression variation. Nature 482, 390–394 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  122. Gaffney, D. J. et al. Dissecting the regulatory architecture of gene expression QTLs. Genome Biol. 13, R7 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  123. Gaulton, K. J. et al. A map of open chromatin in human pancreatic islets. Nature Genet. 42, 255–259 (2010).

    CAS  PubMed  Google Scholar 

  124. Helgadottir, A. et al. A common variant on chromosome 9p21 affects the risk of myocardial infarction. Science 316, 1491–1493 (2007).

    CAS  PubMed  Google Scholar 

  125. Cowper-Sal lari, R. et al. Breast cancer risk-associated SNPs modulate the affinity of chromatin for FOXA1 and alter gene expression. Nature Genet. 44, 1191–1198 (2012).

    CAS  PubMed  Google Scholar 

  126. Kasowski, M. et al. Variation in transcription factor binding among humans. Science 328, 232–235 (2010).

    CAS  PubMed  PubMed Central  Google Scholar 

  127. Bauer, D. E. et al. An erythroid enhancer of BCL11A subject to genetic variation determines fetal hemoglobin level. Science 342, 253–257 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  128. Hsu, P. D., Lander, E. S. & Zhang, F. Development and applications of CRISPR–Cas9 for genome engineering. Cell 157, 1262–1278 (2014).

    CAS  PubMed  PubMed Central  Google Scholar 

Download references

Acknowledgements

The authors were primarily supported by the US National Institutes of Health (NIH) grants DK091183, CA17390 and DK063491, and the San Diego Center for Systems Biology. C.E.R. was supported by the American Heart Association (12POST11760017) and the NIH Pathway to Independence Award (1K99HL12348).

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Christopher K. Glass.

Ethics declarations

Competing interests

The authors declare no competing financial interests.

PowerPoint slides

Glossary

Primed enhancers

Enhancers that have been selected by lineage-determining transcription factors and collaborating transcription factors. They are marked with characteristic histone modifications such as histone H3 lysine 4 monomethylation (H3K4me1) and H3K4me2, but do not produce enhancer RNAs.

DNase I-hypersensitive

Pertaining to genomic sites where the chromatin was made more accessible (that is, hypersensitive) to digestion by DNase I owing to the binding of regulatory proteins.

Poised enhancers

Regulatory elements that are similar to primed enhancers but that are distinguished by the presence of histone H3 lysine 27 trimethylation (H3K27me3), which must be removed to allow the transition to an active enhancer state.

Active enhancers

Enhancers that are marked with histone H3 lysine 27 acetylation (H3K27ac) marks, in addition to the marks of poised enhancers. They produce enhancer RNAs, are bound by the Mediator complex and exert regulatory functions to increase the transcription of target genes.

Latent or de novo enhancer

An inactive enhancer that requires the binding of a combination of transcription factors, including signal- dependent transcription factors, for selection.

Mediator complex

A protein complex with important roles in transcription and the 3D organization of the genome. It integrates various intracellular signals to affect the formation of the pre-initiation complex, transcription initiation and transcription elongation.

Locus control regions

Regions that confer tissue-specific expression to linked transgenes irrespective of the transgene integration site in the genome. These regions display characteristics of both enhancers and insulators.

Exosome

A protein complex involved in the quality control, maturation and degradation of various RNA transcripts, both in the nucleus and cytoplasm.

Chromosome conformation capture

(3C). A method to probe the higher-order structure of the genome by capturing and sequencing DNA sites that are spatially close to each other in the nucleus.

Topologically associated domains

(TADs). Largely self-interacting genomic domains of submegabase size that are further organized into multimegabase-sized structures called nuclear compartments. Genes within TADs are co-regulated, and their expression patterns are highly correlated.

Chromatin hubs

Nuclear domains comprised of regulatory DNA elements (locus control regions, enhancers and promoters) and genes that enable correct gene expression. The smallest unit of a hub could be a topologically associated domain, and the largest could comprise an entire nuclear compartment.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Heinz, S., Romanoski, C., Benner, C. et al. The selection and function of cell type-specific enhancers. Nat Rev Mol Cell Biol 16, 144–154 (2015). https://doi.org/10.1038/nrm3949

Download citation

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/nrm3949

Search

Quick links

Nature Briefing

Sign up for the Nature Briefing newsletter — what matters in science, free to your inbox daily.

Get the most important science stories of the day, free in your inbox. Sign up for Nature Briefing