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:

RNA polymerase II dynamics shape enhancer–promoter interactions

Abstract

How enhancers control target gene expression over long genomic distances remains an important unsolved problem. Here we investigated enhancer–promoter communication by integrating data from nucleosome-resolution genomic contact maps, nascent transcription and perturbations affecting either RNA polymerase II (Pol II) dynamics or the activity of thousands of candidate enhancers. Integration of new Micro-C experiments with published CRISPRi data demonstrated that enhancers spend more time in close proximity to their target promoters in functional enhancer–promoter pairs compared to nonfunctional pairs, which can be attributed in part to factors unrelated to genomic position. Manipulation of the transcription cycle demonstrated a key role for Pol II in enhancer–promoter interactions. Notably, promoter-proximal paused Pol II itself partially stabilized interactions. We propose an updated model in which elements of transcriptional dynamics shape the duration or frequency of interactions to facilitate enhancer–promoter communication.

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: Micro-C contacts are enriched in functional enhancer–promoter pairs.
Fig. 2: Enhancer–promoter contacts depend on active transcription.
Fig. 3: Changes in Pol II pausing and gene body transcription correlate with enhancer–promoter contacts.
Fig. 4: NELFB depletion and recovery correlate with changes in enhancer–promoter contacts.
Fig. 5: Updated model integrating RNA Pol II dynamics into enhancer–promoter interactions.

Similar content being viewed by others

Data availability

Micro-C data generated in this study were deposited in the Gene Expression Omnibus (GEO) database under accession number GSE206133. H3K27ac and H3K4me2 ChIP–seq data from K562 cells were downloaded from GSE163043. K562 data for ATAC-seq (ENCSR868FGK), CTCF ChIP-–seq (ENCSR447BSF), MNase-seq (ENCSR000CXQ) and NELFE ChIP–seq (ENCSR000DOF) were downloaded from ENCODE. DMSO, TRP and FLV-treated mESCs Micro-C data28 were downloaded from GSE130275. PRO-seq data36,40 for Jurkat T cells were downloaded from GSE66031 and for K562 from GSE60455. PRO-seq data for mECSs harboring a homozygous endogenous NELFB-FKBP12F36V fusion protein, treated and untreated with dTAG-13 (ref. 85), were downloaded from GSE196653. GRO-seq data for mESCs58,99 were downloaded from GSE43390 and GSE48895. Positions for human (hg38) and mouse (mm10) CAGE peaks were downloaded from the FANTOM5 database (https://fantom.gsc.riken.jp/5/).

Code availability

All data normalization and visualization code are available at https://github.com/Danko-Lab/E-P_contacts ref. 105.

References

  1. Levine, M. & Tjian, R. Transcription regulation and animal diversity. Nature 424, 147–151 (2003).

    Article  CAS  PubMed  Google Scholar 

  2. 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 

  3. Gruss, P., Dhar, R. & Khoury, G. Simian virus 40 tandem repeated sequences as an element of the early promoter. Proc. Natl Acad. Sci. USA 78, 943–947 (1981).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  4. Benoist, C. & Chambon, P. In vivo sequence requirements of the SV40 early promoter region. Nature 290, 304–310 (1981).

    Article  CAS  PubMed  Google Scholar 

  5. Choi, O. R. & Engel, J. D. Developmental regulation of β-globin gene switching. Cell 55, 17–26 (1988).

    Article  CAS  PubMed  Google Scholar 

  6. 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 

  7. Robson, M. I., Ringel, A. R. & Mundlos, S. Regulatory landscaping: how enhancer–promoter communication is sculpted in 3D. Mol. Cell 74, 1110–1122 (2019).

    Article  CAS  PubMed  Google Scholar 

  8. Hamamoto, K. & Fukaya, T. Molecular architecture of enhancer–promoter interaction. Curr. Opin. Cell Biol. 74, 62–70 (2022).

    Article  CAS  PubMed  Google Scholar 

  9. 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 

  10. 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 

  11. Horsfield, J. & Marsman, J. Chromatin: Methods and Protocols (Springer, 2022).

  12. 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 

  13. 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 

  14. Fulco, C. P. et al. Activity-by-contact model of enhancer–promoter regulation from thousands of CRISPR perturbations. Nat. Genet. 51, 1664–1669 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

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

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  16. Beagan, J. A. et al. Three-dimensional genome restructuring across timescales of activity-induced neuronal gene expression. Nat. Neurosci. 23, 707–717 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  17. Mateo, L. J. et al. Visualizing DNA folding and RNA in embryos at single-cell resolution. Nature 568, 49–54 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  18. 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 

  19. 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 

  20. 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 

  21. Chen, H. et al. Dynamic interplay between enhancer–promoter topology and gene activity. Nat. Genet. 50, 1296–1303 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  22. Rao, S. S. P. et al. Cohesin loss eliminates all loop domains. Cell 171, 305–320 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  23. El Khattabi, L. et al. A pliable mediator acts as a functional rather than an architectural bridge between promoters and enhancers. Cell 178, 1145–1158 (2019).

    Article  CAS  PubMed  Google Scholar 

  24. 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 

  25. Malik, S. & Roeder, R. G. Mediator: a drawbridge across the enhancer–promoter divide. Mol. cell 64, 433–434 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  26. Ramasamy, S. et al. The mediator complex regulates enhancer–promoter interactions. Preprint at bioRxiv https://doi.org/10.1101/2022.06.15.496245 (2022).

  27. Krietenstein, N. et al. Ultrastructural details of mammalian chromosome architecture. Mol. Cell 78, 554–565 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  28. Hsieh, T.-H. S. et al. Resolving the 3D landscape of transcription-linked mammalian chromatin folding. Mol. Cell 78, 539–553 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  29. Hsieh, T.-H. S., Fudenberg, G., Goloborodko, A. & Rando, O. J. Micro-C XL: assaying chromosome conformation from the nucleosome to the entire genome. Nat. Methods 13, 1009–1011 (2016).

    Article  CAS  PubMed  Google Scholar 

  30. Mahat, D. B. et al. Base-pair-resolution genome-wide mapping of active RNA polymerases using precision nuclear run-on (PRO-seq). Nat. Protoc. 11, 1455–1476 (2016).

    Article  PubMed  PubMed Central  Google Scholar 

  31. Chu, T. et al. Chromatin run-on and sequencing maps the transcriptional regulatory landscape of glioblastoma multiforme. Nat. Genet. 50, 1553–1564 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  32. 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).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  33. Gasperini, M. et al. A genome-wide framework for mapping gene regulation via cellular genetic screens. Cell 176, 377–390 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  34. Fulco, C. P. et al. Systematic mapping of functional enhancer–promoter connections with CRISPR interference. Science 354, 769–773 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  35. Field, A. & Adelman, K. Evaluating enhancer function and transcription. Annu. Rev. Biochem. 89, 213–234 (2020).

    Article  CAS  PubMed  Google Scholar 

  36. Danko, C. G. et al. Identification of active transcriptional regulatory elements from GRO-seq data. Nat. Methods 12, 433–438 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  37. Wang, Z., Chu, T., Choate, L. A. & Danko, C. G. Identification of regulatory elements from nascent transcription using dREG. Genome Res. 29, 293–303 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  38. Zhou, B. et al. Comprehensive, integrated, and phased whole-genome analysis of the primary ENCODE cell line K562. Genome Res. 29, 472–484 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

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

    Article  CAS  PubMed  Google Scholar 

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

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  41. Scruggs, B. S. et al. Bidirectional transcription arises from two distinct hubs of transcription factor binding and active chromatin. Mol. Cell 58, 1101–1112 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  42. Andersson, R., Sandelin, A. & Danko, C. G. A unified architecture of transcriptional regulatory elements. Trends Genet. 31, 426–433 (2015).

    Article  CAS  PubMed  Google Scholar 

  43. Lim, B. & Levine, M. S. enhancer–promoter communication: hubs or loops? Curr. Opin. Genet. Dev. 67, 5–9 (2020).

    Article  PubMed  PubMed Central  Google Scholar 

  44. Cho, W.-K. et al. Mediator and RNA polymerase II clusters associate in transcription-dependent condensates. Science 361, 412–415 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  45. Shrinivas, K. et al. Enhancer features that drive formation of transcriptional condensates. Mol. Cell 75, 549–561 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  46. Lee, J.-H. et al. Enhancer RNA m6A methylation facilitates transcriptional condensate formation and gene activation. Mol. Cell 81, 3368–3385 (2021).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  47. Sabari, B. R. et al. Coactivator condensation at super-enhancers links phase separation and gene control. Science 361, eaar3958 (2018).

    Article  PubMed  PubMed Central  Google Scholar 

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

    Article  CAS  PubMed  Google Scholar 

  49. Di Giammartino, D. C., Polyzos, A. & Apostolou, E. Transcription factors: building hubs in the 3D space. Cell Cycle 19, 2395–2410 (2020).

    Article  PubMed  PubMed Central  Google Scholar 

  50. Guo, Y. E. et al. Pol II phosphorylation regulates a switch between transcriptional and splicing condensates. Nature 572, 543–548 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  51. Sigova, A. A. et al. Transcription factor trapping by RNA in gene regulatory elements. Science 350, 978–981 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  52. Boehning, M. et al. RNA polymerase II clustering through carboxy-terminal domain phase separation. Nat. Struct. Mol. Biol. 25, 833–840 (2018).

    Article  CAS  PubMed  Google Scholar 

  53. Jiang, Y. et al. Genome-wide analyses of chromatin interactions after the loss of Pol I, Pol II, and Pol III. Genome Biol. 21, 158 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  54. Zhang, S., Uebelmesser, N., Barbieri, M. & Papantonis, A. Enhancer–promoter contact formation requires RNAPII and antagonizes loop extrusion. Nat. Genet., 55(5), 832–840 (2023).

  55. Glaser, L. V. et al. Assessing genome-wide dynamic changes in enhancer activity during early mESC differentiation by FAIRE-STARR-seq. Nucleic Acids Res. 49, 12178–12195 (2021).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  56. Hsieh, T.-H. S. et al. Resolving the 3D landscape of transcription-linked mammalian chromatin folding. Preprint at bioRxiv https://doi.org/10.1101/638775 (2019).

  57. Mitchell, J. A. & Fraser, P. Transcription factories are nuclear subcompartments that remain in the absence of transcription. Genes Dev. 22, 20–25 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  58. Jonkers, I., Kwak, H. & Lis, J. T. Genome-wide dynamics of Pol II elongation and its interplay with promoter proximal pausing, chromatin, and exons. eLife 3, e02407 (2014).

    Article  PubMed  PubMed Central  Google Scholar 

  59. Siepel, A. A unified probabilistic modeling framework for eukaryotic transcription based on nascent RNA sequencing data. Preprint at bioRxiv https://doi.org/10.1101/2021.01.12.426408 (2022).

  60. Shao, W. & Zeitlinger, J. Paused RNA polymerase II inhibits new transcriptional initiation. Nat. Genet. 49, 1045–1051 (2017).

    Article  CAS  PubMed  Google Scholar 

  61. Nabet, B. et al. The dTAG system for immediate and target-specific protein degradation. Nat. Chem. Biol. 14, 431–441 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  62. Abuhashem, A. et al. RNA Pol II pausing facilitates phased pluripotency transitions by buffering transcription. Genes Dev. 36, 770–789 (2022).

    Article  PubMed  PubMed Central  Google Scholar 

  63. Aoi, Y. et al. NELF regulates a promoter-proximal step distinct from RNA Pol II pause-release. Mol. Cell 78, 261–274 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  64. Palstra, R.-J. et al. The β-globin nuclear compartment in development and erythroid differentiation. Nat. Genet. 35, 190–194 (2003).

    Article  CAS  PubMed  Google Scholar 

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

    Article  PubMed  Google Scholar 

  66. Chong, S. et al. Imaging dynamic and selective low-complexity domain interactions that control gene transcription. Science 361, eaar2555 (2018).

    Article  PubMed  PubMed Central  Google Scholar 

  67. Fukaya, T., Lim, B. & Levine, M. Enhancer control of transcriptional bursting. Cell 166, 358–368 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  68. Beagrie, R. A. et al. Complex multi-enhancer contacts captured by genome architecture mapping. Nature 543, 519–524 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  69. Hua, P. et al. Defining genome architecture at base-pair resolution. Nature 595, 125–129 (2021).

    Article  CAS  PubMed  Google Scholar 

  70. Wang, Z. et al. Prediction of histone post-translational modification patterns based on nascent transcription data. Nat. Genet. 54, 295–305 (2022).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  71. Martin, B. J. E. et al. Transcription shapes genome-wide histone acetylation patterns. Nat. Commun. 12, 210 (2021).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

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

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  73. Strom, A. R. et al. Phase separation drives heterochromatin domain formation. Biophys. J. 114, 445a (2018).

    Article  Google Scholar 

  74. Nair, S. J. et al. Phase separation of ligand-activated enhancers licenses cooperative chromosomal enhancer assembly. Nat. Struct. Mol. Biol. 26, 193–203 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  75. Barutcu, A. R., Blencowe, B. J. & Rinn, J. L. Differential contribution of steady-state RNA and active transcription in chromatin organization. EMBO Rep. 20, e48068 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  76. Soutourina, J. Transcription regulation by the mediator complex. Nat. Rev. Mol. Cell Biol. 19, 262–274 (2018).

    Article  CAS  PubMed  Google Scholar 

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

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  78. Henriques, T. et al. Stable pausing by RNA polymerase II provides an opportunity to target and integrate regulatory signals. Mol. Cell 52, 517–528 (2013).

    Article  CAS  PubMed  Google Scholar 

  79. Tome, J. M., Tippens, N. D. & Lis, J. T. Single-molecule nascent RNA sequencing identifies regulatory domain architecture at promoters and enhancers. Nat. Genet. 50, 1533–1541 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  80. Pancholi, A. et al. RNA polymerase II clusters form in line with surface condensation on regulatory chromatin. Mol. Syst. Biol. 17, e10272 (2021).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  81. Hajiabadi, H. et al. Deep-learning microscopy image reconstruction with quality control reveals second-scale rearrangements in RNA polymerase II clusters. PNAS Nexus 1, pgac065 (2022).

    Article  PubMed  PubMed Central  Google Scholar 

  82. Henninger, J. E. et al. RNA-mediated feedback control of transcriptional condensates. Cell 184, 207–225 (2021).

    Article  CAS  PubMed  Google Scholar 

  83. Wang, Z. et al. Prediction of histone post-translational modification patterns based on nascent transcription data. Nat Genet. 54, 295–305 (2022).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  84. Luo, Y. et al. New developments on the Encyclopedia of DNA Elements (ENCODE) data portal. Nucleic Acids Res. 48, D882–D889 (2020).

    Article  CAS  PubMed  Google Scholar 

  85. Abuhashem, A., Lee, A. S., Joyner, A. L. & Hadjantonakis, A.-K. Rapid and efficient degradation of endogenous proteins in vivo identifies stage-specific roles of RNA Pol II pausing in mammalian development. Dev. Cell 57, 1068–1080 (2022).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  86. Gu, B. et al. Transcription-coupled changes in nuclear mobility of mammalian cis-regulatory elements. Science 359, 1050–1055 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  87. Imsoonthornruksa, S., Noisa, P., Parnpai, R. & Ketudat-Cairns, M. A simple method for production and purification of soluble and biologically active recombinant human leukemia inhibitory factor (hLIF) fusion protein in Escherichia coli. J. Biotechnol. 151, 295–302 (2011).

    Article  CAS  PubMed  Google Scholar 

  88. 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 

  89. Goloborodko, A. et al. mirnylab/distiller-nf: v0.3.3. Zenodo. https://doi.org/10.5281/zenodo.3350937 (2019).

  90. Li, H. & Durbin, R. Fast and accurate short read alignment with Burrows–Wheeler transform. Bioinformatics 25, 1754–1760 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  91. Ramírez, F. et al. High-resolution TADs reveal DNA sequences underlying genome organization in flies. Nat. Commun. 9, 189 (2018).

    Article  PubMed  PubMed Central  Google Scholar 

  92. Lopez-Delisle, L. et al. pyGenomeTracks: reproducible plots for multivariate genomic datasets. Bioinformatics 37, 422–423 (2021).

    Article  CAS  PubMed  Google Scholar 

  93. Chu, T., Wang, Z., Chou, S.-P. & Danko, C. G. Discovering transcriptional regulatory elements from run-on and sequencing data using the web-based dREG gateway. Curr. Protoc. Bioinformatics 66, e70 (2019).

    Article  PubMed  Google Scholar 

  94. Chou, S.-P., He, A. Y., Wang, Z. & Danko, C. Danko-Lab/proseq2.0: PRO-seq pipeline used by the Danko lab. Zenodo. https://doi.org/10.5281/ZENODO.8015719 (2023).

  95. Schmieder, R. & Edwards, R. Quality control and preprocessing of metagenomic datasets. Bioinformatics 27, 863–864 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  96. Martin, M. Cutadapt removes adapter sequences from high-throughput sequencing reads. EMBnet J. 17, 10 (2011).

    Article  Google Scholar 

  97. Quinlan, A. R. BEDTools: the Swiss-army tool for genome feature analysis. Curr. Protoc. Bioinformatics 47, 1–34 (2014).

    Article  Google Scholar 

  98. Anders, S. & Huber, W. Differential expression analysis for sequence count data. Genome Biol. 11, R106 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  99. Williams, L. H. et al. Pausing of RNA polymerase II regulates mammalian developmental potential through control of signaling networks. Mol. Cell 58, 311–322 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  100. Dorighi, K. M. et al. Mll3 and Mll4 facilitate enhancer RNA synthesis and transcription from promoters independently of H3K4 monomethylation. Mol. Cell 66, 568–576 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  101. Noguchi, S. et al. FANTOM5 CAGE profiles of human and mouse samples. Sci. Data 4, 170112 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  102. Zhao, Y. et al. Deconvolution of Expression for Nascent RNA sequencing data (DENR) highlights pre-RNA isoform diversity in human cells. Bioinformatics 37, 4727–4736 (2021).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  103. 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 

  104. Open2C et al. Cooltools: enabling high-resolution Hi-C analysis in Python. Preprint at bioRxiv https://doi.org/10.1101/2022.10.31.514564 (2022).

  105. Barshad, G. & Wang, Z. Danko-Lab/E-P_contacts: E-P contacts. Zenodo. https://doi.org/10.5281/ZENODO.7948817 (2023).

Download references

Acknowledgements

We thank E. Apostolou and members of her lab for commenting on a manuscript draft as well as members of the Danko, Lis, and Yu Labs for valuable discussions and suggestions throughout the life of this project. Work in this publication was supported by R01-HG010346 and R01-HG009309 (NHGRI) to C.G.D. A.A. is supported by the NIH (T32GM007739 and F30HD103398). Work in AKH Lab is supported by the NIH (R01HD094868, R01DK127821, R01HD086478 and P30CA008748). The content is solely the responsibility of the authors and does not necessarily represent the official views of the US National Institutes of Health. Some of the figures in this manuscript were created using BioRender.

Author information

Authors and Affiliations

Authors

Contributions

G.B., J.J.L. and C.G.D. designed the study. E.J.R. generated PRO-seq data. G.B. and A.G.C. analyzed PRO-seq data. A.A. and A.K.H. contributed NELFB-dTAG mESCs. A.A. performed all treatments and crosslinking for the time course of the NELFB degradation dTAG experiments. G.B. collected all Micro-C data. N.K. and O.J.R. assisted with the collection of Jurkat Micro-C and taught G.B. the protocol. J.J.L., Y.M., Z.W. and G.B. implemented contact normalization by local decay (Contact Caller). G.B., J.J.L., A.G.C., N.K., O.J.R., A.K.H. and C.G.D. analyzed and interpreted data. G.B. and C.G.D. drafted the manuscript. All authors read and approved the final manuscript.

Corresponding author

Correspondence to Charles G. Danko.

Ethics declarations

Competing interests

The authors declare no competing interests.

Peer review

Peer review information

Nature Genetics thanks the anonymous reviewers for their contribution to the peer review of this work. Peer reviewer reports are available.

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 Details for the comparison between functional and nonfunctional enhancer-promoter pairs.

(a) Schematic representation of the LOWESS-based normalization for enhancer-promoter contacts. (b) Box and dot plots, similar to Fig. 1b, comparing the observed contact frequency relative to expected by a local distance-decay function of the validated functional enhancers in the MYC locus (functional pairs) compared to the rest of the dREG-detected TIRs in the TAD (nonfunctional pairs) with the MYC promoter. Here we divided the CRISPRi-tested TIRs to those that fall within the first 0.5 Mb (near, functional: n = 4, nonfunctional: n = 27) or beyond 1.5 Mb (far, functional: n = 3, nonfunctional: n = 12) within the TAD. Two-sided Mann-Whitney p-values are indicated. (c) Violin plots comparing contact levels relative to expected by local distance-decay function of functional versus the nonfunctional enhancer-promoter pairs in the genome, before matching for enhancer-promoter distance, accessibility or target gene expression. On the left column, the results are based on dREG CRISPRi-targeted TIRs33 either before (top) and after (bottom) excluding pairs that do not fall into the same mega-haplotype (MH) or fall within known structural variants (SVs) in K562 cells96. The middle violin plot shows the same as the top-left one, but using data from a different CRISPRi dataset97. The two violin plots on the right show the same as the two on the left, using the same CRISPRi dataset98, but centering on H3K27ac overlapping ATAC-peaks instead of TIRs. Two-sided Mann-Whitney p-values are indicated. (d) Venn diagram showing the overlap between H3K27ac+ ATAC-peaks (H3K27ac)- and dREG TIRs-defined enhancers tested by CRISPRi in33.

Extended Data Fig. 2 Matching possible confounders between CRISPRi functional and nonfunctional pairs.

Histograms demonstrating the distribution of functional and high-confidence nonfunctional enhancer-promoter pairs in terms of enhancer-promoter genomic distance (top), accessibility by mean ATAC-seq signal (middle) and PRO-seq target gene transcription signal in reads per kilobase per million reads (RPKM) (bottom), after matching for these possible confounding factors.

Extended Data Fig. 3 Functional constituent enhancers within super enhancers interact more with the target promoter.

a,b) Violin plots comparing contact levels relative to expected by local distance-decay function of functional versus the nonfunctional enhancer-promoter pairs in the genome, where the enhancers are mapped within (a) or outside (b) K562 defined super enhancers101. Two-sided Mann-Whitney p-values are indicated (c) Dot plot shows the log2 distribution of local genomic distance-normalized contact frequency between CRISPRi-defined functional constituent enhancers within 16 super enhancers compared to other constituent enhancers within these super enhancers. The dashed lines connect data points representing the median values of the same super enhancer. Two-sided Wilcoxon paired-test p-values are shown. (d) Dot plot shows the log2 distribution of local genomic distance-normalized contact frequency between CRISPRi-defined high-confidence nonfunctional constituent enhancers within 15 super enhancers compared to other constituent enhancers within these super enhancers. The dashed lines connect data points representing the median values of the same super enhancer. Two-sided Wilcoxon paired-test p-values is shown. (e) Violin plot shows the distribution of the ratios between the functional constituent enhancers to other constituent enhancers in the same super enhancer (yellow) and between high-confidence nonfunctional constituent enhancers to other constituent enhancers in the same super enhancer. Two-sided Mann-Whitney p-value is indicated.

Extended Data Fig. 4 Micro-C 1D signal near TSSs genome-wide.

One dimensional contact signal for intra-chromosomal contacts with both sides having mapping quality (mapq) ≥ 30. Total median signal was smoothed using a sliding window of 100 bp. Shown are signals around promoter TSSs (orange), enhancer TSSs (purple) and all TSSs genome-wide (black).

Extended Data Fig. 5 Elaborated schematic representation of the APA method used to calculate 1D background-normalized changes in contacts between samples.

(a) To calculate the observed change in contacts, a matrix of contacts between each enhancer-promoter pair within the limited defined genomic distance range was calculated and then all of these matrices were summed to obtain the observed aggregated matrix. To get the obs the sequencing depth-normalized aggregated matrices were divided by the control matrix. Shown are also the depth-normalized aggregated matrices for the DMSO control, TRP- and FLV-treated mESCs, as well as the obs matrices for both treatments. (b) To calculate the 1D signal background matrices we calculated the average of 1D Micro-C signal vectors across cells around enhancers and promoters. Line plot representations of these vectors at 20 kb windows around enhancers and promoters, across cells of 200 bp are shown for both treatment conditions and DMSO control. The 1D background change matrix, B, was calculated by dividing the 1D signal background matrix of each treatment by the control. The 1D background matrices for both treatments and control samples as well as the matrices B for both treatments in mESCs are shown.

Extended Data Fig. 6 Genomic distance has little effect on the shape of enhancer-promoter contacts fold change.

Matrices showing the observed fold-change (a) the 1D background signal fold change (b) and the 1D background-normalized enhancer-promoter fold change (c) following TRP and FLV treatment compared to the DMSO control at 25kb-wide distance ranges starting at 25–50 kb (leftmost column) and ending at 125–150 kb (rightmost column). (d) Matrix showing the 1D background-normalized fold change of contacts between CTCF bound motifs following TRP and FLV treatments, compared with the DMSO control.

Extended Data Fig. 7 Changes in enhancer-promoter contacts at the Pou5f1 locus following transcriptional inhibition.

Virtual 4 C signal showing Micro-C signal associated with Pou5f1 promoter from a ~ 1.3 billion contacts library of untreated mESC, as well as the FLV and TRP treated mESCs (~400 million contacts each). Shown are also GRO-seq and ATAC-seq signals. Two regulatory elements shown to induce Pou5f1 gene expression55 are shown in green and the relative contacts between these regulatory elements and the Pou5f1 promoter, relative to the untreated control, in each treatment are shown in the associated bar plots. The position of the anchor for the virtual 4 C is shown.

Extended Data Fig. 8 Distribution of fold change in gene body transcription for K562 and Jurkat upregulated genes.

(a) Scatterplots where each dot represents a single enhancer-promoter pair where the promoter was associated with higher gene body transcription (top, n = 4,071) and pausing signal (bottom, n = 502) at Jurkat T cells compared to K562. The dots are colored based on the density of dots relative to their coordinates. The associated boxplots show the distribution of enhancer-promoter contacts relative to local background in both cell types, relative to the median ratio in K562. (*** Two-sided Wilcoxon signed-rank test p-value < 1 × 10−100). (b) Boxplots showing the distributions of fold change in gene body signal in genes with no associated paused Pol II change (NPC, K562 > Jurkat: n = 173, Jurkat > K562: n = 64) and associated significant paused Pol II change (PC, K562 > Jurkat: n = 167, Jurkat > K562: n = 43) (‘ns’ - Two-sided Mann-Whitney p-value > 0.5). (c) Boxplot depicting the relative increase of enhancer-promoter contacts associated with promoters of genes with upregulated gene body transcription in Jurkat T-cells with a corresponding significant increase in pausing signal (pause change—PC, n = 43) and without a change in pausing signal (no pause change—NPC, n = 64) (** Two-sided Mann-Whitney p-value < 1 × 10−10).

Extended Data Fig. 9 Changes in enhance-promoter contacts architecture following NELFB depletion.

(a) APA heatmaps of the 1D change-normalized contact change (log2) between enhancer and promoter regions at 20 kb around TSSs. Pixel size is 200 bp square. The APA heatmaps are oriented such that the gene TSS points to the right and the dominant TSS of the enhancer points upwards. (b) Line plot of the median fold changes at the dot (blue), stripes (red) and edges (gray) relative to T = 0 at the different time points of dTAG treatments and following dTAG washout.

Extended Data Fig. 10 Changes in ZRS-Shh contacts following NELFB depletion.

Micro-C contact maps in 10 kb resolution along with the associated virtual 4 C signal and PRO-seq signal in mESCs not treated (top) or treated (bottom) with the dTAG ligand for 30 minutes to degrade NELFB. The positions of the ZRS enhancer and the Shh promoter are indicated in red rectangles.

Supplementary information

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

Barshad, G., Lewis, J.J., Chivu, A.G. et al. RNA polymerase II dynamics shape enhancer–promoter interactions. Nat Genet 55, 1370–1380 (2023). https://doi.org/10.1038/s41588-023-01442-7

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41588-023-01442-7

This article is cited by

Search

Quick links

Nature Briefing: Translational Research

Sign up for the Nature Briefing: Translational Research newsletter — top stories in biotechnology, drug discovery and pharma.

Get what matters in translational research, free to your inbox weekly. Sign up for Nature Briefing: Translational Research