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:

Increased enhancer–promoter interactions during developmental enhancer activation in mammals

Abstract

Remote enhancers are thought to interact with their target promoters via physical proximity, yet the importance of this proximity for enhancer function remains unclear. Here we investigate the three-dimensional (3D) conformation of enhancers during mammalian development by generating high-resolution tissue-resolved contact maps for nearly a thousand enhancers with characterized in vivo activities in ten murine embryonic tissues. Sixty-one percent of developmental enhancers bypass their neighboring genes, which are often marked by promoter CpG methylation. The majority of enhancers display tissue-specific 3D conformations, and both enhancer–promoter and enhancer–enhancer interactions are moderately but consistently increased upon enhancer activation in vivo. Less than 14% of enhancer–promoter interactions form stably across tissues; however, these invariant interactions form in the absence of the enhancer and are likely mediated by adjacent CTCF binding. Our results highlight the general importance of enhancer–promoter physical proximity for developmental gene activation in mammals.

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: Identification of enhancer-centric chromatin interactions in ten mouse embryonic tissues.
Fig. 2: Properties of promoters that are skipped by remote enhancers.
Fig. 3: Enhancers are required for expression of interacting genes.
Fig. 4: Tissue specificity of developmental enhancer interactions.
Fig. 5: Imaging E–P interactions in the developing mouse embryo.
Fig. 6: Properties of tissue-invariant E–P chromatin interactions.

Similar content being viewed by others

Data availability

Sequencing data generated in this study are available at the Gene Expression Omnibus repository under accession number GSE217078. Several mouse embryonic ChIP–seq, DNase-seq, bisulfite-seq and RNA-seq data for different tissues at E11.5 were downloaded from ENCODE (https://www.encodeproject.org/). The CTCF ChIP–seq datasets used for comparison were downloaded from GEO (https://www.ncbi.nlm.nih.gov/geo/) under accession numbers GSM5501396, GSM5501397 and GSM5501398. Enhancer interaction profiles are available at https://www.kvonlab.org/data/echic. Source data are provided with this paper.

Code availability

Public software and packages were used following the developer’s manuals. The custom code used for data analysis has been deposited at GitHub (https://github.com/kvonlab/Chen_et_al_2024) and Zenodo (https://doi.org/10.5281/zenodo.10594800)100.

References

  1. Lettice, L. A. et al. Disruption of a long-range cis-acting regulator for Shh causes preaxial polydactyly. Proc. Natl Acad. Sci. USA 99, 7548–7553 (2002).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  2. Long, H. K. et al. Loss of extreme long-range enhancers in human neural crest drives a craniofacial disorder. Cell Stem Cell 27, 765–783 (2020).

    CAS  Google Scholar 

  3. 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 

  4. Long, H. K., Prescott, S. L. & Wysocka, J. Ever-changing landscapes: transcriptional enhancers in development and evolution. Cell 167, 1170–1187 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  5. Schoenfelder, S. & Fraser, P. Long-range enhancer–promoter contacts in gene expression control. Nat. Rev. Genet. 20, 437–455 (2019).

    Article  CAS  PubMed  Google Scholar 

  6. 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 

  7. Nora, E. P. et al. Spatial partitioning of the regulatory landscape of the X-inactivation centre. Nature 485, 381–385 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  8. Lupiáñez, D. G. et al. Disruptions of topological chromatin domains cause pathogenic rewiring of gene–enhancer interactions. Cell 161, 1012–1025 (2015).

    Article  PubMed  PubMed Central  Google Scholar 

  9. Lettice, L. A. et al. Enhancer-adoption as a mechanism of human developmental disease. Hum. Mutat. 32, 1492–1499 (2011).

    Article  CAS  PubMed  Google Scholar 

  10. 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 

  11. Bulger, M. & Groudine, M. Looping versus linking: toward a model for long-distance gene activation. Genes Dev. 13, 2465–2477 (1999).

    Article  CAS  PubMed  Google Scholar 

  12. Furlong, E. E. M. & Levine, M. Developmental enhancers and chromosome topology. Science 361, 1341–1345 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  13. Oudelaar, A. M. & Higgs, D. R. The relationship between genome structure and function. Nat. Rev. Genet. 22, 154–168 (2021).

    Article  CAS  PubMed  Google Scholar 

  14. Andrey, G. et al. A switch between topological domains underlies HoxD genes collinearity in mouse limbs. Science 340, 1234167 (2013).

    Article  PubMed  Google Scholar 

  15. Tolhuis, B., Palstra, R. J., Splinter, E., Grosveld, F. & de Laat, W. Looping and interaction between hypersensitive sites in the active β-globin locus. Mol. Cell 10, 1453–1465 (2002).

    Article  CAS  PubMed  Google Scholar 

  16. Madsen, J. G. S. et al. Highly interconnected enhancer communities control lineage-determining genes in human mesenchymal stem cells. Nat. Genet. 52, 1227–1238 (2020).

    Article  CAS  PubMed  Google Scholar 

  17. Rubin, A. J. et al. Lineage-specific dynamic and pre-established enhancer–promoter contacts cooperate in terminal differentiation. Nat. Genet. 49, 1522–1528 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  18. Deng, W. et al. Reactivation of developmentally silenced globin genes by forced chromatin looping. Cell 158, 849–860 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

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

    Article  CAS  PubMed  Google Scholar 

  20. Andrey, G. et al. Characterization of hundreds of regulatory landscapes in developing limbs reveals two regimes of chromatin folding. Genome Res. 27, 223–233 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

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

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  22. Montavon, T. et al. A regulatory archipelago controls Hox genes transcription in digits. Cell 147, 1132–1145 (2011).

    Article  CAS  PubMed  Google Scholar 

  23. Alexander, J. M. et al. Live-cell imaging reveals enhancer-dependent Sox2 transcription in the absence of enhancer proximity. eLife 8, e41769 (2019).

    Article  PubMed  PubMed Central  Google Scholar 

  24. Benabdallah, N. S. et al. Decreased enhancer–promoter proximity accompanying enhancer activation. Mol. Cell 76, 473–484 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  25. Gómez Acuña, L. I., Flyamer, I., Boyle, S., Friman, E. T. & Bickmore, W. A. Transcription decouples estrogen-dependent changes in enhancer–promoter contact frequencies and physical proximity. Preprint at bioRxiv https://doi.org/10.1101/2023.03.29.534720 (2023).

  26. Zaugg, J. B. et al. Current challenges in understanding the role of enhancers in disease. Nat. Struct. Mol. Biol. 29, 1148–1158 (2022).

    Article  CAS  PubMed  Google Scholar 

  27. Visel, A., Minovitsky, S., Dubchak, I. & Pennacchio, L. A. VISTA Enhancer Browser—a database of tissue-specific human enhancers. Nucleic Acids Res. 35, D88–D92 (2006).

    Article  PubMed  PubMed Central  Google Scholar 

  28. Attanasio, C. et al. Fine tuning of craniofacial morphology by distant-acting enhancers. Science 342, 1241006 (2013).

    Article  PubMed  PubMed Central  Google Scholar 

  29. Dickel, D. E. et al. Genome-wide compendium and functional assessment of in vivo heart enhancers. Nat. Commun. 7, 12923 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  30. Dickel, D. E. et al. Ultraconserved enhancers are required for normal development. Cell 172, 491–499 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  31. Kvon, E. Z. et al. Progressive loss of function in a limb enhancer during snake evolution. Cell 167, 633–642 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  32. Osterwalder, M. et al. Enhancer redundancy provides phenotypic robustness in mammalian development. Nature 554, 239–243 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  33. Gorkin, D. U. et al. An atlas of dynamic chromatin landscapes in mouse fetal development. Nature 583, 744–751 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  34. Smith, C. M. et al. The mouse Gene Expression Database (GXD): 2014 update. Nucleic Acids Res. 42, D818–D824 (2014).

    Article  CAS  PubMed  Google Scholar 

  35. Kvon, E. Z., Waymack, R., Gad, M. & Wunderlich, Z. Enhancer redundancy in development and disease. Nat. Rev. Genet. 22, 324–336 (2021).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

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

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  37. Fudenberg, G. & Imakaev, M. FISH-ing for captured contacts: towards reconciling FISH and 3C. Nat. Methods 14, 673–678 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  38. Giorgetti, L. & Heard, E. Closing the loop: 3C versus DNA FISH. Genome Biol. 17, 215 (2016).

    Article  PubMed  PubMed Central  Google Scholar 

  39. Kubo, N. et al. Promoter-proximal CTCF binding promotes distal enhancer-dependent gene activation. Nat. Struct. Mol. Biol. 28, 152–161 (2021).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  40. Rao, S. S. P. et al. A 3D map of the human genome at kilobase resolution reveals principles of chromatin looping. Cell 159, 1665–1680 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  41. Paliou, C. et al. Preformed chromatin topology assists transcriptional robustness of Shh during limb development. Proc. Natl Acad. Sci. USA 116, 12390–12399 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  42. Sagai, T., Hosoya, M., Mizushina, Y., Tamura, M. & Shiroishi, T. Elimination of a long-range cis-regulatory module causes complete loss of limb-specific Shh expression and truncation of the mouse limb. Development 132, 797–803 (2005).

    Article  CAS  PubMed  Google Scholar 

  43. Hsieh, T.-H. S. et al. Enhancer–promoter interactions and transcription are largely maintained upon acute loss of CTCF, cohesin, WAPL or YY1. Nat. Genet. 54, 1919–1932 (2022).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  44. Zuin, J. et al. Nonlinear control of transcription through enhancer–promoter interactions. Nature 604, 571–577 (2022).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  45. Xiao, J. Y., Hafner, A. & Boettiger, A. N. How subtle changes in 3D structure can create large changes in transcription. eLife 10, e64320 (2021).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  46. Williamson, I. et al. Spatial genome organization: contrasting views from chromosome conformation capture and fluorescence in situ hybridization. Genes Dev. 28, 2778–2791 (2014).

    Article  PubMed  PubMed Central  Google Scholar 

  47. Brandão, H. B., Gabriele, M. & Hansen, A. S. Tracking and interpreting long-range chromatin interactions with super-resolution live-cell imaging. Curr. Opin. Cell Biol. 70, 18–26 (2021).

    Article  PubMed  Google Scholar 

  48. Chen, L.-F., Lee, J. & Boettiger, A. Recent progress and challenges in single-cell imaging of enhancer–promoter interaction. Curr. Opin. Genet. Dev. 79, 102023 (2023).

    Article  CAS  PubMed  Google Scholar 

  49. Jerkovic, I. & Cavalli, G. Understanding 3D genome organization by multidisciplinary methods. Nat. Rev. Mol. Cell Biol. 22, 511–528 (2021).

    Article  CAS  PubMed  Google Scholar 

  50. Pownall, M. E. et al. Chromatin expansion microscopy reveals nanoscale organization of transcription and chromatin. Science 381, 92–100 (2023).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  51. Espinola, S. M. et al. Cis-regulatory chromatin loops arise before TADs and gene activation, and are independent of cell fate during early Drosophila development. Nat. Genet. 53, 477–486 (2021).

    Article  CAS  PubMed  Google Scholar 

  52. Simmons, E. et al. Independence of chromatin conformation and gene regulation during Drosophila dorsoventral patterning. Nat. Genet. 53, 487–499 (2021).

    Article  Google Scholar 

  53. Platt, J. L. et al. Capture‐C reveals preformed chromatin interactions between HIF‐binding sites and distant promoters. EMBO Rep. 17, 1410–1421 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  54. Ray, J. et al. Chromatin conformation remains stable upon extensive transcriptional changes driven by heat shock. Proc. Natl Acad. Sci. USA 116, 19431–19439 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  55. Pollex, T. et al. New enhancer–promoter interactions are gained during tissue differentiation and reflect changes in E/P activity. Preprint at bioRxiv https://doi.org/10.1101/2022.12.07.519443 (2022).

  56. Smemo, S. et al. Obesity-associated variants within FTO form long-range functional connections with IRX3. Nature 507, 371–375 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  57. Birnbaum, R. Y. et al. Functional characterization of tissue-specific enhancers in the DLX5/6 locus. Hum. Mol. Genet. 21, 4930–4938 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  58. 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 

  59. Li, G. et al. Extensive promoter-centered chromatin interactions provide a topological basis for transcription regulation. Cell 148, 84–98 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  60. Kvon, E. Z. et al. Genome-scale functional characterization of Drosophila developmental enhancers in vivo. Nature 512, 91–95 (2014).

    Article  CAS  PubMed  Google Scholar 

  61. Calhoun, V. C., Stathopoulos, A. & Levine, M. Promoter-proximal tethering elements regulate enhancer–promoter specificity in the Drosophila Antennapedia complex. Proc. Natl Acad. Sci. USA 99, 9243–9247 (2002).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  62. Ringel, A. R. et al. Repression and 3D-restructuring resolves regulatory conflicts in evolutionarily rearranged genomes. Cell 185, 3689–3704 (2022).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  63. Juven-Gershon, T. & Kadonaga, J. T. Regulation of gene expression via the core promoter and the basal transcriptional machinery. Dev. Biol. 339, 225–229 (2010).

    Article  CAS  PubMed  Google Scholar 

  64. Zabidi, M. A. et al. Enhancer–core-promoter specificity separates developmental and housekeeping gene regulation. Nature 518, 556–559 (2014).

    Article  PubMed  PubMed Central  Google Scholar 

  65. Bergman, D. T. et al. Compatibility rules of human enhancer and promoter sequences. Nature 607, 176–184 (2022).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  66. Martinez-Ara, M., Comoglio, F., van Arensbergen, J. & van Steensel, B. Systematic analysis of intrinsic enhancer–promoter compatibility in the mouse genome. Mol. Cell 82, 2519–2531 (2022).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  67. Batut, P. J. et al. Genome organization controls transcriptional dynamics during development. Science 375, 566–570 (2022).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  68. Pachano, T. et al. Orphan CpG islands amplify poised enhancer regulatory activity and determine target gene responsiveness. Nat. Genet. 53, 1036–1049 (2021).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  69. Galouzis, C. C. & Furlong, E. E. M. Regulating specificity in enhancer–promoter communication. Curr. Opin. Cell Biol. 75, 102065 (2022).

    Article  CAS  PubMed  Google Scholar 

  70. Chakraborty, S. et al. Enhancer–promoter interactions can bypass CTCF-mediated boundaries and contribute to phenotypic robustness. Nat. Genet. 55, 280–290 (2023).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  71. Hung, T.-C., Kingsley, D. M. & Boettiger, A. N. Boundary stacking interactions enable cross-TAD enhancer–promoter communication during limb development. Nat. Genet. 56, 306–314 (2024).

  72. Jiang, Y. et al. The methyltransferase SETDB1 regulates a large neuron-specific topological chromatin domain. Nat. Genet. 49, 1239–1250 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  73. Wang, Y. et al. The 3D Genome Browser: a web-based browser for visualizing 3D genome organization and long-range chromatin interactions. Genome Biol. 19, 151 (2018).

    Article  PubMed  PubMed Central  Google Scholar 

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

    Article  Google Scholar 

  75. Amândio, A. R. et al. Sequential in cis mutagenesis in vivo reveals various functions for CTCF sites at the mouse HoxD cluster. Genes Dev. 35, 1490–1509 (2021).

    Article  PubMed  PubMed Central  Google Scholar 

  76. Flöttmann, R. et al. Noncoding copy-number variations are associated with congenital limb malformation. Genet. Med. 20, 599–607 (2018).

    Article  PubMed  Google Scholar 

  77. Short, P. J. et al. De novo mutations in regulatory elements in neurodevelopmental disorders. Nature 555, 611–616 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  78. Firth, H. V. et al. DECIPHER: database of chromosomal imbalance and phenotype in humans using Ensembl resources. Am. J. Hum. Genet. 84, 524–533 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  79. ENCODE Project Consortium et al. Expanded encyclopaedias of DNA elements in the human and mouse genomes. Nature 583, 699–710 (2020).

    Article  Google Scholar 

  80. Dryden, N. H. et al. Unbiased analysis of potential targets of breast cancer susceptibility loci by capture Hi-C. Genome Res. 24, 1854–1868 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  81. Hughes, J. R. et al. Analysis of hundreds of cis-regulatory landscapes at high resolution in a single, high-throughput experiment. Nat. Genet. 46, 205–212 (2014).

    Article  CAS  PubMed  Google Scholar 

  82. Orlando, G., Kinnersley, B. & Houlston, R. S. Capture Hi-C library generation and analysis to detect chromatin interactions. Curr. Protoc. Hum. Genet. 98, e63 (2018).

    Article  PubMed  Google Scholar 

  83. 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 

  84. Wingett, S. et al. HiCUP: pipeline for mapping and processing Hi-C data. F1000Res. 4, 1310 (2015).

    Article  PubMed  PubMed Central  Google Scholar 

  85. Cairns, J. et al. CHiCAGO: robust detection of DNA looping interactions in capture Hi-C data. Genome Biol. 17, 127 (2016).

    Article  PubMed  PubMed Central  Google Scholar 

  86. Freire-Pritchett, P. et al. Detecting chromosomal interactions in capture Hi-C data with CHiCAGO and companion tools. Nat. Protoc. 16, 4144–4176 (2021).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  87. Schoenfelder, S. et al. The pluripotent regulatory circuitry connecting promoters to their long-range interacting elements. Genome Res. 25, 582–597 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  88. Cairns, J., Orchard, W. R., Malysheva, V. & Spivakov, M. Chicdiff: a computational pipeline for detecting differential chromosomal interactions in capture Hi-C data. Bioinformatics 35, 4764–4766 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  89. Zerbino, D. R., Wilder, S. P., Johnson, N., Juettemann, T. & Flicek, P. R. The Ensembl Regulatory Build. Genome Biol. 16, 56 (2015).

    Article  PubMed  PubMed Central  Google Scholar 

  90. Yu, G., Wang, L.-G., Han, Y. & He, Q.-Y. clusterProfiler: an R package for comparing biological themes among gene clusters. OMICS 16, 284–287 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  91. Wu, T. et al. clusterProfiler 4.0: a universal enrichment tool for interpreting omics data. Innovation 2, 100141 (2021).

    CAS  PubMed  PubMed Central  Google Scholar 

  92. Osterwalder, M. et al. Characterization of mammalian in vivo enhancers using mouse transgenesis and CRISPR genome editing. Methods Mol. Biol. 2403, 147–186 (2022).

    Article  CAS  PubMed  Google Scholar 

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

    Article  CAS  PubMed  Google Scholar 

  94. Putri, G. H., Anders, S., Pyl, P. T., Pimanda, J. E. & Zanini, F. Analysing high-throughput sequencing data in Python with HTSeq 2.0. Bioinformatics 38, 2943–2945 (2022).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  95. Love, M. I., Huber, W. & Anders, S. Moderated estimation of fold change and dispersion for RNA-seq data with DESeq2. Genome Biol. 15, 550 (2014).

    Article  PubMed  PubMed Central  Google Scholar 

  96. Chaumeil, J., Micsinai, M. & Skok, J. A. Combined immunofluorescence and DNA FISH on 3D-preserved interphase nuclei to study changes in 3D nuclear organization. J. Vis. Exp. 3, e50087 (2013).

    Google Scholar 

  97. Naciri, I. et al. Linking chromosomal silencing with Xist expression from autosomal integrated transgenes. Front. Cell Dev. Biol. 9, 693154 (2021).

    Article  PubMed  PubMed Central  Google Scholar 

  98. Chaumeil, J., Augui, S., Chow, J. C. & Heard, E. Combined immunofluorescence, RNA fluorescent in situ hybridization, and DNA fluorescent in situ hybridization to study chromatin changes, transcriptional activity, nuclear organization, and X-chromosome inactivation. Methods Mol. Biol. 463, 297–308 (2008).

    Article  CAS  PubMed  Google Scholar 

  99. Bolland, D. J., King, M. R., Reik, W., Corcoran, A. E. & Krueger, C. Robust 3D DNA FISH using directly labeled probes. J. Vis. Exp. https://doi.org/10.3791/50587 (2013).

    Article  PubMed  PubMed Central  Google Scholar 

  100. Chen, Z. & Kvon, E. Z. Analysis pipeline for Chen_et_al_2024 paper. Zenodo https://doi.org/10.5281/zenodo.10594800 (2024).

    Article  Google Scholar 

  101. Kvon, E. Z. et al. Comprehensive in vivo interrogation reveals phenotypic impact of human enhancer variants. Cell 180, 1262–1271 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  102. Uslu, V. V. et al. Long-range enhancers regulating Myc expression are required for normal facial morphogenesis. Nat. Genet. 46, 753–758 (2014).

    Article  CAS  PubMed  Google Scholar 

  103. Padhi, E. M. et al. Coding and noncoding variants in EBF3 are involved in HADDS and simplex autism. Hum. Genomics 15, 44 (2021).

    CAS  Google Scholar 

  104. Turner, T. N. et al. Genomic patterns of de novo mutation in simplex autism. Cell 171, 710–722 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  105. Shibata, M., Kurokawa, D., Nakao, H., Ohmura, T. & Aizawa, S. MicroRNA-9 modulates Cajal–Retzius cell differentiation by suppressing Foxg1 expression in mouse medial pallium. J. Neurosci. 28, 10415–10421 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  106. Shimogori, T. et al. A genomic atlas of mouse hypothalamic development. Nat. Neurosci. 13, 767–775 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  107. Rodríguez-Carballo, E. et al. The HoxD cluster is a dynamic and resilient TAD boundary controlling the segregation of antagonistic regulatory landscapes. Genes Dev. 31, 2264–2281 (2017).

    Article  PubMed  PubMed Central  Google Scholar 

Download references

Acknowledgements

We acknowledge the UCI Transgenic Mouse Facility for help with generation of enhancer-knockout mice and the UCI Genomics Research and Technology Hub for help with sequencing as well as L. Scipioni and S. Sun for help with DNA FISH. This work was supported by National Institutes of Health grants R00HG009682 and DP2GM149555 (to E.Z.K.), R01HG003988 (to L.A.P.) and F31HD112201 (to G.B.). Z.C. was supported by NSF grant DMS1763272 (to Qing Nie) and Simons Foundation grant 594598 (to Qing Nie). J.L.-R. is funded by the Spanish Ministerio de Ciencia e Innovación (grant PID2020-113497GB-I00 and institutional María de Maeztu grant CEX2020-001088-M). Research conducted at the E.O. Lawrence Berkeley National Laboratory was performed under Department of Energy Contract DE-AC02-05CH11231, University of California. The funders had no role in study design, data collection and analysis, decision to publish or preparation of the manuscript.

Author information

Authors and Affiliations

Authors

Contributions

E.Z.K. conceived the project with input from Z.C., V.S., D.E.D., A.V. and L.A.P. Z.C., V.S., I.B. and E.Z.K. designed experiments. Z.C., V.S., G.B., S.J., B.C. and E.Z.K. performed capture Hi-C experiments, and Z.C. analyzed the data with input from E.Z.K. and B.J.M. Z.C., S.J., A.D. and E.Z.K. performed enhancer-knockout studies, and Z.C. analyzed the data. Z.C. and G.B. performed 3D-FISH experiments and analyzed the data. A.A.-C. and J.L.-R. performed ISH experiments. E.Z.K and Z.C. wrote the paper with input from the remaining authors.

Corresponding author

Correspondence to Evgeny Z. Kvon.

Ethics declarations

Competing interests

The authors declare no competing interests.

Peer review

Peer review information

Nature Genetics thanks Daniel Ibrahim, Douglas Higgs and the other, anonymous, reviewer(s) for their contribution to the peer review of this work.

Additional information

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

Extended data

Extended Data Fig. 1 Enhancer capture Hi-C identifies enhancer-centric chromatin interactions in mouse embryonic tissues.

a, Unique on-target read counts for each library. The percentages above indicate the capture rates for each library. b, c, Principal component analysis and hierarchical clustering of all replicates based on the presence of peaks called by CHiCAGO in each replicate (considering peaks with valid di-tags on neighboring fragments). d, Significant enhancer-centric chromatin interactions identified in this study. The number on each link represents the number of fragments falling into different annotation categories and the width of links is proportional to the percentage (in the parentheses) of different kinds of interactions. Only interactions within 2 Mb are included. CTCF sites with ‘B’: CTCF sites at TAD boundary; Pc: polycomb; Enh: enhancers; Bait-Enh: baited enhancers; Pr: promoters. e, An average number of interactions detected per bait for different kinds of baits (promoter (n = 176), enhancer (n = 935) and negative control elements (n = 87)). Data are represented as mean ± s.e.m. f, Distribution of genomic distances between enhancers and the TSSs of interacting genes (black, frequencies; red, cumulative). g, Violin plots showing read counts on promoters of active genes that interact with enhancer baits (n = 541), promoter baits (n = 126) and control element baits (n = 25). The central horizontal lines are the median, with the boxes extending from the 25th to the 75th percentiles. The whiskers further extend by ±1.5 times the interquartile range from the limits of each box. h, Histogram showing the proportion of bait regions that interact with proximal genes and distal genes. i, Venn diagram showing the overlap between significant interactions called from enhancer baits and corresponding promoter baits. All P values were calculated by a two-sided Wilcox test and adjusted for multiple testing. j, Zoom-in view on Zic1/Zic4 locus for hs654 interaction profiles across 10 tissues. The average size for each pooled fragment is ~3 kb. FB, forebrain. MB, midbrain. HB, hindbrain. CF, face. HR, heart. FL, forelimb. HL, hindlimb. TK, trunk. TL, tail. NT, neural tube.

Extended Data Fig. 2 Examples of enhancer—promoter interactions linked to congenital disorders.

a, Hs1507 limb enhancer (green) located in the non-coding region which is duplicated in patients with polydactyly (pink bar indicates the homologous region in the mouse genome)8. Hs1507 forms significant chromatin interactions with the promoter of the Epha4 (blue line) located ~1.5 Mb away. Shown is the Epha4 genomic region (chr1:74,788,119-77,634,678; mm10). b, Many de novo rare variants identified in patients with preaxial polydactyly101 are located in the ZRS limb enhancer (green bar) which forms significant interactions with the promoter of Shh located ~850 kb away. Shown is the Shh (blue line) genomic region (chr5:28,320,000-29,400,000; mm10). c, Hs1877 face enhancer (green) located in the non-coding region containing 146 SNPs found in patients with cleft lip risk (purple bar indicates the homologous region in the mouse genome)102. Hs1877 forms significant chromatin interactions with the promoter of the Myc (blue line) located ~900 kb away in the face. The Myc genomic region (chr15:61,880,003-63,506,895; mm10). d, Three de novo rare variants identified in patients with autism are located in the hs737 midbrain/hindbrain enhancer (green bar)103,104, which forms strong significant interactions with the promoter of Ebf3 (blue line) located ~1,000 kb away in the midbrain. Shown is the Ebf3 genomic region (chr7:136,018,204-137,420,338; mm10). Red arrowheads indicate capture Hi-C viewpoints.

Extended Data Fig. 3 Examples of enhancer—enhancer chromatin interactions.

a, The Mir9-2 genomic region (chr13:83,558,457-84,861,438; mm10) is shown with chromatin interaction heatmaps centered on hs268 (blue), hs267 (green), hs266 (yellow) and hs853 (red) enhancers in the forebrain (FB) and forelimb (FL). Shown on the top are hs268, hs267, hs266 and hs853 enhancer activities in a transgenic mid-gestation (E11.5) mouse embryo, which match with the expression profiles of Mir9 in the brain and neural tube at E11.5105,106. Red arrowheads indicate capture Hi-C viewpoints. Arches indicate significant interactions in the forebrain. Shown on the bottom are H3K27ac (yellow) and H3K4me3 (green) ChIP-seq tracks in forebrain and limb buds (LB) at E11.5, CTCF (light blue) ChIP-seq tracks in the whole brain (WB) and forelimb at E12.533,74,75,107. b, Schematic depicting 3D chromatin interactions between enhancers and Mir9-2 gene in the forebrain and forelimb. c, Coronal sections of forebrain for hs268, hs267, hs266 and hs853 enhancer activity from VISTA enhancer database27, which reproducibly label the same subregions in E11.5 forebrain as C130071C03Rik (Mir9-2 precursor) expression106. d, e, Chromatin interaction heatmaps centered on mm1165, hs746, mm428 and mm427 enhancers in the face (CF) and forebrain (FB) for Msx1 genomic region (chr5: 37,554,764-38,206,723; mm10) (d) and hs1315 and mm1403 enhancers in the neural tube (NT) and forelimb (FL) for Tfap2a genomic region (chr13: 39,098,000-41,000,000; mm10) (e). Shown on the top are mm1165, hs746, mm428, mm427, hs1315 and mm1403 enhancer activities in a transgenic mid-gestation (E11.5) mouse embryos. Arches indicate significant interactions. Red arrowheads indicate capture Hi-C viewpoints.

Extended Data Fig. 4 Properties of enhancer-interacting and skipped promoters.

ac, The CpG methylation (a), mRNA expression levels (b) and DNase signal (c) of enhancer-interacting and skipped promoters in tissues where enhancers are active. High me, high methylation skipped promoters (>50% CpG methylation within ± 1 kb from TSS). Low me, low methylation skipped promoters (<50% CpG methylation within ± 1 kb from TSS). d, e, H3K27me3 (d), H3K9me3 (e) signal at ± 2.5 kb of enhancer-interacting and skipped promoters in tissues where enhancers are active. The pie charts below show the fraction of promoters marked with H3K27me3 or H3K9me3. f, Pie charts showing the fraction of skipped promoters marked by CpG methylation, H3K27me3, H3K9me3 or the combination of marks. g-i, Violin plot showing CpG length (g), or CpG methylation level at transcription start sites for enhancer-interacting and skipped genes with different window sizes ± 250 bp (h) and ± 2 kb (i)). The number of high and low methylated skipped as well as interacting promoters in CpG analysis are n = 58, n = 86 and n = 71 (CF), n = 138, n = 126 and n = 90 (FB), n = 64, n = 116 and n = 96 (FL) and n = 100, n = 162 and n = 102 (HB), n = 55, n = 92 and n = 91 (HL), n = 213, n = 169 and n = 125 (MB) and, n = 87, n = 86 and n = 87 (NT). FB, forebrain. MB, midbrain. HB, hindbrain. CF, face. FL, forelimb. HL, hindlimb. NT, neural tube. HR, heart. P values are calculated by two-sided Wilcoxon rank test after adjusted for multiple testing (a-c, f-i) or by one-sided chi-squared test (d, e). A statistical test was not performed for H3K9me3 since most of the values are zero. The same DNA methylation, mRNA expression, DNaseI hypersensitivity, H3K27ac and H3K9me3 dataset (a mixture of fore- and hindlimb buds) were used for both fore- and hindlimb interaction analyses. For the boxplots in panels a-e and g-i, the central horizontal lines are the median, with the boxes extending from the 25th to the 75th percentiles. The whiskers further extend by ±1.5 times the interquartile range from the limits of each box.

Extended Data Fig. 5 Zic1/Zic4 and Mir9-2 brain enhancer knock-outs.

a, Map of the deleted region encompassing hs654 midbrain enhancer of Zic1/Zic4 together with H3K27ac, DNase-seq, ATAC-seq from midbrain and conservation track across 60 species. b, Sanger sequencing of the PCR product from hs654 knock-out mice (n = 4 biological replicates). c, representative PCR genotyping results of the hs654 enhancer knockout mice. Lanes in the gel were rearranged so that results for wild-type and heterozygous mice are adjacent to each other. d, Map of the deleted region encompassing hs267, hs266 and hs853 forebrain enhancers of Mir9-2 together with H3K27ac, DNase-seq, ATAC-seq from midbrain and conservation track across 60 species. e, Sanger sequencing of the PCR product from hs267-853 knock-out mice (n = 3 biological replicates). f, representative PCR genotyping results of the hs267-853 enhancer knockout mice. g, Genotype frequency data for enhancer knockout lines. Mice homozygous for either deletion were born at normal Mendelian ratios, and no gross phenotypes or impairments were observed. P-values were calculated using the one-sided chi-square test.

Source data

Extended Data Fig. 6 Enhancer–promoter interaction frequency across tissues.

a, The ratio of E–P interaction frequency between active and inactive tissues. Red dashed line indicates no difference between active and inactive tissues. b, Univariate logistic regression for relative interaction frequencies and enhancer activity across all tissues. c, The ratio of E–P interaction frequency between active and inactive tissues for interactions within or across TADs. d, The distribution of distances between the closest TAD boundary and enhancer for enhancers acting within or across TADs. e, The ratio of interaction frequency between active and inactive tissues on interacting promoters or intervening regions before and after removing ENCODE annotated elements ( ± 20 kb). Red dashed line indicates no difference between active and inactive tissues. f, The ratio of E–P interaction frequency between active and inactive tissues for enhancers with different ranks. Only tissues with ≥10 interactions in each rank category are shown. g, The fold-change of gene expression levels between active state (baited enhancers interact with active promoters) and inactive state (baited enhancers don’t interact with promoters or in inactive tissues). Data are represented as mean ± s.e.m. h, The ratio of E–P interaction frequency between active and inactive tissues for expressed genes (TPM > = 0.5) and lowly expressed or inactive genes (TPM < 0.5). i, Chromatin interaction profiles in forebrain, midbrain and hindbrain centered on the enhancer hs1172 at Nr2f1 locus (chr13:78,057,768-78,705,499). j, The ratio of E–P interaction frequency between active and inactive brain regions for enhancers active in one of the brain domains. k-m, Cumulative frequency plots of interprobe distances for the indicated loci and tissues. n-p, Frequency distribution of FISH interprobe distances in 250 nm bins between Zic1/4 and hs654 (n), Mir9-2 and hs266 (o), Snai2 and hs1431 (p) in indicated tissues. P values are calculated by paired-sample (a, c, e, g, h, j) or unpaired-sample (d, f) two-sided Wilcoxon rank test and adjusted for multiple testing or by one-sided chi-squared test (b, n-p). For the boxplots in panels a, c-f, h and j, the central horizontal lines are the median, with the boxes extending from the 25th to the 75th percentiles. The whiskers further extend by ±1.5 times the interquartile range from the limits of each box.

Extended Data Fig. 7 Properties of invariant enhancer–promoter interactions.

a, Metaplot showing average ratio of enhancer interaction frequency between active and inactive tissues for invariant (interactions present in all 7 main tissues: brain, face, limb, heart, neural tube, trunk and tail, n = 171) and tissue-specific ( ≤ 6 main tissues, n = 775) interactions. Light blue/orange shading indicates 95% confidence intervals estimated by non-parametric bootstrapping. 10 kb upstream and downstream of enhancer bait and target promoter are also indicated (grey bar). b, The average ratio of invariant enhancer-promoter interaction frequency between active and inactive tissues for enhancers active in the brain, face, limb, heart and neural tube E–P. Data is shown only for tissues with at least 20 active enhancers that form invariant E–P interactions. P values were calculated by paired-sample two-sided Wilcox test and adjusted for multiple testing. c, The number of tissues in which enhancers forming invariant (10 tissues, n = 98) or tissue-specific ( ≤ 4 tissues, n = 196) E–P interactions are active in vivo. d, The average phyloP scores of enhancers forming invariant (10 tissues, n = 98) or tissue-specific ( ≤ 4 tissues, n = 196) E–P interactions. P values in panels c and d were calculated by two-sided Wilcox test. e, Gene Ontology enrichment for genes that form invariant (10 tissues) E–P interactions (Biological process and Molecular function). Q values were calculated by over-representation test and adjusted for multiple testing. For the boxplots in panels b-d, the central horizontal lines are the median, with the boxes extending from the 25th to the 75th percentiles. The whiskers further extend by ±1.5 times the interquartile range from the limits of each box.

Extended Data Fig. 8 Tissue specificity of enhancer-enhancer chromatin interactions.

a, Pie chart showing the fraction of E–E interactions present in different numbers of tissues. b, The average ratio of E–E interaction frequency between active and inactive tissues for enhancers active in neural tube, heart, tail and trunk. The number of E–E interactions for each tissue is indicated at the top. P values were calculated by paired-sample two-sided Wilcox test and adjusted for multiple testing. c, The average ratio of enhancer–enhancer interaction frequency between active and inactive tissues for enhancers of different ranks. The E–E interaction number for rank 3 to 5 are n = 217, n = 122 and n = 69 (brain), n = 53, n = 59 and n = 18 (cf), n = 100, n = 84 and n = 45 (limb), n = 80, n = 51 and n = 32 (nt), respectively. Cf: face. Nt: neural tube. P values were calculated by unpaired-sample two-sided Wilcox test with multiple testing. For the boxplots in panels b and c, the central horizontal lines are the median, with the boxes extending from the 25th to the 75th percentiles. The whiskers further extend by ±1.5 times the interquartile range from the limits of each box.

Extended Data Fig. 9 CRISPR/Cas9-mediated ZRS limb enhancer replacement with a fragment of the lacZ gene.

a, Schematic overview of the strategy for ZRS enhancer replacement. A 4.5 kb mouse genomic region containing the ZRS enhancer (red) is shown together with the vertebrate conservation track (dark blue). The donor vector contained two homology arms (gray) and an inactive fragment of the lacZ coding sequence (blue). The sgRNA recognition site is indicated in purple. PCR primers used for genotyping are shown as arrows. b, PCR genotyping analysis of heterozygous and wildtype mice using primer pairs LacZ-F1 and LacZ-R1 or LacZ-F2 and LacZ-R2. See Methods for details. c, Shh whole-mount in situ hybridization in E10.5 wild type (left) and ZRSlacZ/lacZ knock-in embryos (n ≥ 3 biological replicates for each genotype). Shh expression is not detectable in limb buds but is present elsewhere in the embryo.

Source data

Supplementary information

Reporting Summary

Supplementary Tables 1–5

Supplementary Table 1. Enhancer bait regions. Supplementary Table 2. Enhancer-centric significant interactions. Supplementary Table 3. Interactions for regions linked to human congenital disorders. Supplementary Table 4. Probe information for 3D DNA FISH. Supplementary Table 5. Primer sequences used in this study.

Source data

Source Data Fig. 5

Statistical source data for 3D DNA FISH.

Source Data Extended Data Fig. 5

Uncropped gels for Extended Data Fig. 5.

Source Data Extended Data Fig. 9

Uncropped gels for Extended Data Fig. 9b and all WISH images for Extended Data Fig. 9c.

Rights and permissions

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

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Chen, Z., Snetkova, V., Bower, G. et al. Increased enhancer–promoter interactions during developmental enhancer activation in mammals. Nat Genet 56, 675–685 (2024). https://doi.org/10.1038/s41588-024-01681-2

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41588-024-01681-2

This article is cited by

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