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:

3D Enhancer–promoter networks provide predictive features for gene expression and coregulation in early embryonic lineages

Abstract

Mammalian embryogenesis commences with two pivotal and binary cell fate decisions that give rise to three essential lineages: the trophectoderm, the epiblast and the primitive endoderm. Although key signaling pathways and transcription factors that control these early embryonic decisions have been identified, the non-coding regulatory elements through which transcriptional regulators enact these fates remain understudied. Here, we characterize, at a genome-wide scale, enhancer activity and 3D connectivity in embryo-derived stem cell lines that represent each of the early developmental fates. We observe extensive enhancer remodeling and fine-scale 3D chromatin rewiring among the three lineages, which strongly associate with transcriptional changes, although distinct groups of genes are irresponsive to topological changes. In each lineage, a high degree of connectivity, or ‘hubness’, positively correlates with levels of gene expression and enriches for cell-type specific and essential genes. Genes within 3D hubs also show a significantly stronger probability of coregulation across lineages compared to genes in linear proximity or within the same contact domains. By incorporating 3D chromatin features, we build a predictive model for transcriptional regulation (3D-HiChAT) that outperforms models using only 1D promoter or proximal variables to predict levels and cell-type specificity of gene expression. Using 3D-HiChAT, we identify, in silico, candidate functional enhancers and hubs in each cell lineage, and with CRISPRi experiments, we validate several enhancers that control gene expression in their respective lineages. Our study identifies 3D regulatory hubs associated with the earliest mammalian lineages and describes their relationship to gene expression and cell identity, providing a framework to comprehensively understand lineage-specific transcriptional behaviors.

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: Transcriptional changes and enhancer remodeling accompany early developmental decisions.
Fig. 2: Hi-C and H3K27ac HiChIP reveal multilayered 3D genomic reorganization and complex networks of putative regulatory interactions in TSCs, ESCs and XEN cells.
Fig. 3: Association of high 3D hubness with levels, cell-type specificity and coregulation of gene expression in early embryonic fates.
Fig. 4: Association of 3D rewiring with cell-type specific gene expression.
Fig. 5: Predictive modeling using 3D chromatin features outperforms promoter-based or 1D-based models for gene expression levels or cell-type specificity.
Fig. 6: Experimental validation of predicted enhancers in ESC and XEN.

Similar content being viewed by others

Data availability

All genomic datasets generated in this study (ChIP-seq, ChIP-exo, ATAC-seq, RNA-seq, 4C-seq, Hi-C and HiChIP) have been uploaded in the Gene Expression Omnibus (GEO) under accession number GSE213645. Source RT–qPCR data (normalized values) are provided along with all statistics in Supplementary Table 8.

Code availability

Custom R scripts used for data analysis in this study have been developed in our lab and are available upon request. The 3D-HiChAT code for calculating predicted gene expression and for scoring impactful enhancers is available on Github at https://github.com/Apostolou-Lab/3DHiChAT.

References

  1. Alberio, R. Regulation of cell fate decisions in early mammalian embryos. Annu. Rev. Anim. Biosci. 8, 377–393, https://doi.org/10.1146/annurev-animal-021419-083841 (2020).

    Article  PubMed  Google Scholar 

  2. Bardot, E. S. & Hadjantonakis, A. K. Mouse gastrulation: coordination of tissue patterning, specification and diversification of cell fate. Mech. Dev. 163, 103617, https://doi.org/10.1016/j.mod.2020.103617 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  3. Rossant, J. Making the mouse blastocyst: past, present, and future. Curr. Top. Dev. Biol. 117, 275–288, https://doi.org/10.1016/bs.ctdb.2015.11.015 (2016).

    Article  PubMed  Google Scholar 

  4. Rossant, J. & Tam, P. P. L. Blastocyst lineage formation, early embryonic asymmetries and axis patterning in the mouse. Development 136, 701–713, https://doi.org/10.1242/dev.017178 (2009).

    Article  CAS  PubMed  Google Scholar 

  5. Grabarek, J. B. et al. Differential plasticity of epiblast and primitive endoderm precursors within the ICM of the early mouse embryo. Development 139, 129–39 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  6. Cui, W. & Mager, J. Transcriptional regulation and genes involved in first lineage specification during preimplantation development. Adv. Anat. Embryol. Cell Biol. 229, 31–46, https://doi.org/10.1007/978-3-319-63187-5_4 (2018).

    Article  PubMed  PubMed Central  Google Scholar 

  7. Frum, T. & Ralston, A. Cell signaling and transcription factors regulating cell fate during formation of the mouse blastocyst. Trends Genet. 31, 402–410, https://doi.org/10.1016/j.tig.2015.04.002 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  8. Muñoz-Descalzo, S., Hadjantonakis, A. K. & Arias, A. M. Wnt/ß-catenin signalling and the dynamics of fate decisions in early mouse embryos and embryonic stem (ES) cells. Semin. Cell Dev. Biol. 47, 101–109 (2015).

    Article  PubMed  Google Scholar 

  9. Lim, B. & Levine, M. S. Enhancer–promoter communication: hubs or loops? Curr. Opin. Genet. Dev. 67, 5–9 (2021).

    Article  CAS  PubMed  Google Scholar 

  10. Schoenfelder, S. & Fraser, P. Long-range enhancer–promoter contacts in gene expression control. Nat. Rev. Genet. 20, 437–455, https://doi.org/10.1038/s41576-019-0128-0 (2019).

    Article  CAS  PubMed  Google Scholar 

  11. Creyghton, M. P. et al. Histone H3K27ac separates active from poised enhancers and predicts developmental state. Proc. Natl Acad. Sci. USA 110, 21931–21936 (2010).

    Article  Google Scholar 

  12. Wu, J. et al. Chromatin analysis in human early development reveals epigenetic transition during ZGA. Nature 557, 256–260 (2018).

    Article  CAS  PubMed  Google Scholar 

  13. Birney, E. et al. Identification and analysis of functional elements in 1% of the human genome by the ENCODE pilot project. Nature 447, 799–816 (2007).

    Article  CAS  PubMed  Google Scholar 

  14. Roadmap Epigenomics Consortium, et al.Integrative analysis of 111 reference human epigenomes. Nature 518, 317–330, https://doi.org/10.1038/nature14248 (2015).

    Article  CAS  PubMed Central  Google Scholar 

  15. Yue, F. et al. A comparative encyclopedia of DNA elements in the mouse genome. Nature 515, 355–364, https://doi.org/10.1038/nature13992 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  16. Arnold, C. D. et al. Genome-wide quantitative enhancer activity maps identified by STARR-seq. Science 339, 1074–1077, https://doi.org/10.1126/science.1232542 (2013).

    Article  CAS  PubMed  Google Scholar 

  17. Lopes, R., Korkmaz, G. & Agami, R. Applying CRISPR-Cas9 tools to identify and characterize transcriptional enhancers. Nat. Rev. Mol. Cell Biol. 17, 597–604, https://doi.org/10.1038/nrm.2016.79 (2016).

    Article  CAS  PubMed  Google Scholar 

  18. Apostolou, E. et al. Genome-wide chromatin interactions of the Nanog locus in pluripotency, differentiation, and reprogramming. Cell Stem Cell 12, 699–712, https://doi.org/10.1016/j.stem.2013.04.013 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  19. Beagan, J. A. et al. Local genome topology can exhibit an incompletely rewired 3D-folding state during somatic cell reprogramming. Cell Stem Cell 18, 611–624, https://doi.org/10.1016/j.stem.2016.04.004 (2016).

  20. Dekker, J. et al. The 4D nucleome project. Nature 549, 219–226, https://doi.org/10.1038/nature23884 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  21. Denholtz, M. et al. Long-range chromatin contacts in embryonic stem cells reveal a role for pluripotency factors and polycomb proteins in genome organization. Cell Stem Cell 13, 602–616, https://doi.org/10.1016/j.stem.2013.08.013 (2013).

    Article  CAS  PubMed  Google Scholar 

  22. Dixon, J. R. et al. Topological domains in mammalian genomes identified by analysis of chromatin interactions. Nature 485, 376–380, https://doi.org/10.1038/nature11082 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  23. Di Giammartino, D. C. & Apostolou, E. The chromatin signature of pluripotency: establishment and maintenance. Curr. Stem Cell Rep. 2, 255–262, https://doi.org/10.1007/s40778-016-0055-3 (2016).

    Article  PubMed  PubMed Central  Google Scholar 

  24. Gorkin, D. U., Leung, D. & Ren, B. The 3D genome in transcriptional regulation and pluripotency. Cell Stem Cell 14, 762–775, https://doi.org/10.1016/j.stem.2014.05.017 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  25. Allahyar, A. et al. Enhancer hubs and loop collisions identified from single-allele topologies. Nat. Genet. 50, 1151–1160, https://doi.org/10.1038/s41588-018-0161-5 (2018).

    Article  CAS  PubMed  Google Scholar 

  26. Beagrie, R. A. et al. Complex multi-enhancer contacts captured by genome architecture mapping. Nature 543, 519–524, https://doi.org/10.1038/nature21411 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  27. Dowen, J. M. et al. Control of cell identity genes occurs in insulated neighborhoods in mammalian chromosomes. Cell 159, 374–387, https://doi.org/10.1016/j.cell.2014.09.030 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  28. Li, G. et al. Extensive promoter-centered chromatin interactions provide a topological basis for transcription regulation. Cell 148, 84–98, https://doi.org/10.1016/j.cell.2011.12.014 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  29. Sun, F. et al. Promoter–enhancer communication occurs primarily within insulated neighborhoods. Mol. Cell 73, 250–263.e5, https://doi.org/10.1016/j.molcel.2018.10.039 (2019).

    Article  CAS  PubMed  Google Scholar 

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

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  31. Downes, D. J. et al. High-resolution targeted 3C interrogation of cis-regulatory element organization at genome-wide scale. Nat. Commun. 12, 531, https://doi.org/10.1038/s41467-020-20809-6 (2021).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  32. 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, https://doi.org/10.1038/ng.2871 (2014).

    Article  CAS  PubMed  Google Scholar 

  33. Hsieh, T. H. S. et al. Resolving the 3D landscape of transcription-linked mammalian chromatin folding. Mol. Cell 78, 539–553.e8, https://doi.org/10.1016/j.molcel.2020.03.002 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  34. Krietenstein, N. et al. Ultrastructural details of mammalian chromosome architecture. Mol. Cell 78, 554–565.e7, https://doi.org/10.1016/j.molcel.2020.03.003 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  35. Crispatzu, G. et al. The chromatin, topological and regulatory properties of pluripotency-associated poised enhancers are conserved in vivo. Nat. Commun. 12, 4344, https://doi.org/10.1038/s41467-021-24641-4 (2021).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  36. Mumbach, M. R. et al. HiChIP: efficient and sensitive analysis of protein-directed genome architecture. Nat. Methods 13, 919–922, https://doi.org/10.1038/nmeth.3999 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  37. Lee, R. et al. CTCF-mediated chromatin looping provides a topological framework for the formation of phase-separated transcriptional condensates. Nucleic Acids Res. 50, 207–226, https://doi.org/10.1093/nar/gkab1242 (2022).

    Article  CAS  PubMed  Google Scholar 

  38. Fulco, C. P. et al. Activity-by-contact model of enhancer–promoter regulation from thousands of CRISPR perturbations. Nat. Genet. 51, 1664–1669, https://doi.org/10.1038/s41588-019-0538-0 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  39. Galouzis, C. C. & Furlong, E. E. M. Regulating specificity in enhancer–promoter communication. Curr. Opin. Cell Biol. 75, 102065, https://doi.org/10.1016/j.ceb.2022.01.010 (2022).

    Article  CAS  PubMed  Google Scholar 

  40. Shlyueva, D., Stampfel, G. & Stark, A. Transcriptional enhancers: from properties to genome-wide predictions. Nat. Rev. Genet. 15, 272–286, https://doi.org/10.1038/nrg3682 (2014).

    Article  CAS  PubMed  Google Scholar 

  41. Collombet, S. et al. Parental-to-embryo switch of chromosome organization in early embryogenesis. Nature 580, 142–146 (2020).

    Article  CAS  PubMed  Google Scholar 

  42. Guo, F. et al. Single-cell multi-omics sequencing of mouse early embryos and embryonic stem cells. Cell Res. 27, 967–988, https://doi.org/10.1038/cr.2017.82 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  43. Mittnenzweig, M. et al. A single-embryo, single-cell time-resolved model for mouse gastrulation. Cell 184, 2825–2842.e22, https://doi.org/10.1016/j.cell.2021.04.004 (2021).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  44. Pijuan-Sala, B. et al. Single-cell chromatin accessibility maps reveal regulatory programs driving early mouse organogenesis. Nat. Cell Biol. 22, 487–497, https://doi.org/10.1038/s41556-020-0489-9 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  45. Nowotschin, S. et al. The emergent landscape of the mouse gut endoderm at single-cell resolution. Nature 569, 361–367, https://doi.org/10.1038/s41586-019-1127-1 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  46. Tanaka, S., Kunath, T., Hadjantonakis, A. K., Nagy, A. & Rossant, J. Promotion to trophoblast stem cell proliferation by FGF4. Science 282, 2072–2075, https://doi.org/10.1126/science.282.5396.2072 (1998).

    Article  CAS  PubMed  Google Scholar 

  47. Tesar, P. J. et al. New cell lines from mouse epiblast share defining features with human embryonic stem cells. Nature 448, 196–199, https://doi.org/10.1038/nature05972 (2007).

    Article  CAS  PubMed  Google Scholar 

  48. Evans, M. J. & Kaufman, M. H. Establishment in culture of pluripotential cells from mouse embryos. Nature 292, 154–156, https://doi.org/10.1038/292154a0 (1981).

    Article  CAS  PubMed  Google Scholar 

  49. Li, Q. V., Rosen, B. P. & Huangfu, D. Decoding pluripotency: genetic screens to interrogate the acquisition, maintenance, and exit of pluripotency. Wiley Interdiscip. Rev. Syst. Biol. Med. 12, e1464, https://doi.org/10.1002/wsbm.1464 (2020).

    Article  PubMed  Google Scholar 

  50. Pelham-Webb, B., Murphy, D. & Apostolou, E. Dynamic 3D chromatin reorganization during establishment and maintenance of pluripotency. Stem Cell Rep. 15, 1176–1195, https://doi.org/10.1016/j.stemcr.2020.10.012 (2020).

    Article  CAS  Google Scholar 

  51. Loof, G. et al. 3D genome topologies distinguish pluripotent epiblast and primitive endoderm cells in the mouse blastocyst. Preprint at bioRxiv https://doi.org/10.1101/2022.10.19.512781 (2022).

  52. Schoenfelder, S. et al. Divergent wiring of repressive and active chromatin interactions between mouse embryonic and trophoblast lineages. Nat. Commun. 9, 4189, https://doi.org/10.1038/s41467-018-06666-4 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  53. Lee, B. K. et al. Super-enhancer-guided mapping of regulatory networks controlling mouse trophoblast stem cells. Nat. Commun. 10, 4749, https://doi.org/10.1038/s41467-019-12720-6 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  54. Thompson, J. J. et al. Extensive co-binding and rapid redistribution of NANOG and GATA6 during emergence of divergent lineages. Nat. Commun. 13, 4257, https://doi.org/10.1038/s41467-022-31938-5 (2022).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  55. Tomikawa, J. et al. Exploring trophoblast-specific Tead4 enhancers through chromatin conformation capture assays followed by functional screening. Nucleic Acids Res. 48, 278–289, https://doi.org/10.1093/nar/gkz1034 (2020).

    Article  CAS  PubMed  Google Scholar 

  56. Wamaitha, S. E. et al. Gata6 potently initiates reprograming of pluripotent and differentiated cells to extraembryonic endoderm stem cells. Genes Dev. 29, 1239–1255, https://doi.org/10.1101/gad.257071.114 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  57. Zhang, Y. et al. Dynamic epigenomic landscapes during early lineage specification in mouse embryos. Nat. Genet. 50, 96–105, https://doi.org/10.1038/s41588-017-0003-x (2018).

    Article  CAS  PubMed  Google Scholar 

  58. Jia, R. et al. Super enhancer profiles identify key cell identity genes during differentiation from embryonic stem cells to trophoblast stem cells super enhencers in trophoblast differentiation. Front. Genet. 12, 762529, https://doi.org/10.3389/fgene.2021.762529 (2021).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  59. Ye, S., Li, P., Tong, C. & Ying, Q. L. Embryonic stem cell self-renewal pathways converge on the transcription factor Tfcp2l1. EMBO J. 32, 2548–2560, https://doi.org/10.1038/emboj.2013.175 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  60. Sun, H. et al. Tfcp2l1 safeguards the maintenance of human embryonic stem cell self-renewal. J. Cell. Physiol. 233, 6944–6951, https://doi.org/10.1002/jcp.26483 (2018).

    Article  CAS  PubMed  Google Scholar 

  61. Yeo, J. C. et al. Klf2 is an essential factor that sustains ground state pluripotency. Cell Stem Cell 14, 864–872 (2014).

    Article  CAS  PubMed  Google Scholar 

  62. Chappell, J. & Dalton, S. Roles for MYC in the establishment and maintenance of pluripotency. Cold Spring Harb. Perspect. Med. 3, a014381, https://doi.org/10.1101/cshperspect.a014381 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  63. Kim, Y., Zheng, X. & Zheng, Y. Proliferation and differentiation of mouse embryonic stem cells lacking all lamins. Cell Res. 23, 1420–1423, https://doi.org/10.1038/cr.2013.118 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  64. Sehgal, P., Chaturvedi, P., Kumaran, R. I., Kumar, S. & Parnaik, V. K. Lamin A/C haploinsufficiency modulates the differentiation potential of mouse embryonic stem cells. PLoS One 8, e57891, https://doi.org/10.1371/journal.pone.0057891 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  65. Rideout, W. M. et al. Generation of mice from wild-type and targeted ES cells by nuclear cloning. Nat. Genet. 24, 109–110, https://doi.org/10.1038/72753 (2000).

    Article  CAS  PubMed  Google Scholar 

  66. Kunath, T. et al. Imprinted X-inactivation in extra-embryonic endoderm cell lines from mouse blastocysts. Development 132, 1649–1661, https://doi.org/10.1242/dev.01715 (2005).

    Article  CAS  PubMed  Google Scholar 

  67. McLean, C. Y. et al. GREAT improves functional interpretation of cis-regulatory regions. Nat. Biotechnol. 28, 495–501, https://doi.org/10.1038/nbt.1630 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  68. Whyte, W. A. et al. Master transcription factors and mediator establish super-enhancers at key cell identity genes. Cell 153, 307–319, https://doi.org/10.1016/j.cell.2013.03.035 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  69. Zhou, H. Y. et al. A Sox2 distal enhancer cluster regulates embryonic stem cell differentiation potential. Genes Dev. 28, 2699–2711, https://doi.org/10.1101/gad.248526.114 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

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

    Article  CAS  PubMed  Google Scholar 

  71. Artus, J., Piliszek, A. & Hadjantonakis, A. K. The primitive endoderm lineage of the mouse blastocyst: sequential transcription factor activation and regulation of differentiation by Sox17. Dev. Biol. 350, 393–404, https://doi.org/10.1016/j.ydbio.2010.12.007 (2011).

    Article  CAS  PubMed  Google Scholar 

  72. Ling, K. W. et al. GATA-2 plays two functionally distinct roles during the ontogeny of hematopoietic stem cells. J. Exp. Med. 200, 871–882, https://doi.org/10.1084/jem.20031556 (2004).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  73. Takahashi, K. & Yamanaka, S. Induction of pluripotent stem cells from mouse embryonic and adult fibroblast cultures by defined factors. Cell 126, 663–676, https://doi.org/10.1016/j.cell.2006.07.024 (2006).

    Article  CAS  PubMed  Google Scholar 

  74. Renaud, S. J., Kubota, K., Rumi, M. A. K. & Soares, M. J. The FOS transcription factor family differentially controls trophoblast migration and invasion. J. Biol. Chem. 289, 5025–5039, https://doi.org/10.1074/jbc.M113.523746 (2014).

    Article  CAS  PubMed  Google Scholar 

  75. Knöfler, M., Vasicek, R. & Schreiber, M. Key regulatory transcription factors involved in placental trophoblast development—a review. Placenta 22, S83–S92, https://doi.org/10.1053/plac.2001.0648 (2001).

    Article  PubMed  Google Scholar 

  76. Benchetrit, H. et al. Direct induction of the three pre-implantation blastocyst cell types from fibroblasts. Cell Stem Cell 24, 983–994.e7, https://doi.org/10.1016/j.stem.2019.03.018 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  77. Fujikura, J. et al. Differentiation of embryonic stem cells is induced by GATA factors. Genes Dev. 16, 784–789, https://doi.org/10.1101/gad.968802 (2002).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  78. Fraser, J. et al. Hierarchical folding and reorganization of chromosomes are linked to transcriptional changes in cellular differentiation. Mol. Syst. Biol. 11, 852 (2015).

    Article  PubMed  PubMed Central  Google Scholar 

  79. Dixon, J. R. et al. Chromatin architecture reorganization during stem cell differentiation. Nature 518, 331–336, https://doi.org/10.1038/nature14222 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  80. Hu, G. et al. Transformation of accessible chromatin and 3D nucleome underlies lineage commitment of early T cells. Immunity 48, 227–242.e8, https://doi.org/10.1016/j.immuni.2018.01.013 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  81. Bhattacharyya, S., Chandra, V., Vijayanand, P. & Ay, F. Identification of significant chromatin contacts from HiChIP data by FitHiChIP. Nat. Commun. 10, 4221 (2019).

    Article  PubMed  PubMed Central  Google Scholar 

  82. Tang, L., Hill, M. C., Ellinor, P. T. & Li, M. Bacon: a comprehensive computational benchmarking framework for evaluating targeted chromatin conformation capture-specific methodologies. Genome Biol. 23, 30 (2022).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  83. Shohat, S. & Shifman, S. Genes essential for embryonic stem cells are associated with neurodevelopmental disorders. Genome Res. 29, 1910–1918 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  84. Tzelepis, K. et al. A CRISPR dropout screen identifies genetic vulnerabilities and therapeutic targets in acute myeloid leukemia. Cell Rep. 17, 1193–1205 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  85. Di Giammartino, D. C. et al. KLF4 is involved in the organization and regulation of pluripotency-associated three-dimensional enhancer networks. Nat. Cell Biol. 21, 1179–1190 (2019).

    Article  PubMed  PubMed Central  Google Scholar 

  86. Krijger, P. H. L. & De Laat, W. Regulation of disease-associated gene expression in the 3D genome. Nat. Rev. Mol. Cell Biol. 17, 771–782 (2016).

    Article  CAS  PubMed  Google Scholar 

  87. Miguel-Escalada, I. et al. Human pancreatic islet three-dimensional chromatin architecture provides insights into the genetics of type 2 diabetes. Nat. Genet. 51, 1137–1148 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  88. 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 

  89. Dejosez, M. et al. Regulatory architecture of housekeeping genes is driven by promoter assemblies. Cell Rep. 42, 112505 (2023).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  90. Sheffield, N. C. & Bock, C. LOLA: enrichment analysis for genomic region sets and regulatory elements in R and bioconductor. Bioinformatics 32, 587–589 (2016).

    Article  CAS  PubMed  Google Scholar 

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

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  92. Luo, R. et al. Dynamic network-guided CRISPRi screen identifies CTCF-loop-constrained nonlinear enhancer gene regulatory activity during cell state transitions. Nat. Genet. 55, 1336–1346 (2023).

    Article  CAS  PubMed  Google Scholar 

  93. Wang, X. et al. The transcription factor TFCP2L1 induces expression of distinct target genes and promotes self-renewal of mouse and human embryonic stem cells. J. Biol. Chem. 294, 6007–6016 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  94. Qiu, D. et al. Klf2 and Tfcp2l1, two Wnt/β-catenin targets, act synergistically to induce and maintain naive pluripotency. Stem Cell Rep. 5, 314–322 (2015).

    Article  CAS  Google Scholar 

  95. Papathanasiou, M. et al. Identification of a dynamic gene regulatory network required for pluripotency factor‐induced reprogramming of mouse fibroblasts and hepatocytes. EMBO J. 40, 102236 (2021).

    Article  Google Scholar 

  96. Li, Y. et al. Gene expression profiling reveals the heterogeneous transcriptional activity of Oct3/4 and its possible interaction with Gli2 in mouse embryonic stem cells. Genomics 102, 456–467 (2013).

    Article  CAS  PubMed  Google Scholar 

  97. Higgs, D. R. Enhancer–promoter interactions and transcription. Nat. Genet. 52, 470–471 (2020).

    Article  CAS  PubMed  Google Scholar 

  98. Spitz, F. & Furlong, E. E. M. Transcription factors: from enhancer binding to developmental control. Nat. Rev. Genet. 13, 613–626 (2012).

    Article  CAS  PubMed  Google Scholar 

  99. Li, J. & Pertsinidis, A. New insights into promoter–enhancer communication mechanisms revealed by dynamic single-molecule imaging. Biochem. Soc. Trans. 49, 1299–1309 (2021).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  100. Schmitt, A. D. et al. A compendium of chromatin contact maps reveals spatially active regions in the human genome. Cell Rep. 17, 2042–2059 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  101. 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 

  102. 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 

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

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  104. 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 

  105. Beer, M. A., Shigaki, D. & Huangfu, D. Enhancer predictions and genome-wide regulatory circuits. Annu. Rev. Genomics Hum. Genet. 21, 37–54 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  106. Tobias, I. C. et al. Transcriptional enhancers: from prediction to functional assessment on a genome-wide scale. Genome 64, 426–448 (2021).

    Article  CAS  PubMed  Google Scholar 

  107. Ernst, J. & Kellis, M. ChromHMM: automating chromatin-state discovery and characterization. Nat. Methods 9, 215–216 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  108. Tippens, N. D. et al. Transcription imparts architecture, function and logic to enhancer units. Nat. Genet. 52, 1067–1075 (2020).

    Article  PubMed  PubMed Central  Google Scholar 

  109. Cao, Q. et al. Reconstruction of enhancer-target networks in 935 samples of human primary cells, tissues and cell lines. Nat. Genet. 49, 1428–1436 (2017).

    Article  CAS  PubMed  Google Scholar 

  110. Whalen, S., Truty, R. M. & Pollard, K. S. Enhancer–promoter interactions are encoded by complex genomic signatures on looping chromatin. Nat. Genet. 48, 488–496 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  111. Karbalayghareh, A., Sahin, M. & Leslie, C. S. Chromatin interaction-aware gene regulatory modeling with graph attention networks. Genome Res. 32, 930–944 (2022).

    PubMed  PubMed Central  Google Scholar 

  112. Bigness, J., Loinaz, X., Patel, S., Larschan, E. & Singh, R. Integrating long-range regulatory interactions to predict gene expression using graph convolutional networks. J. Comput. Biol. 29, 409–424 (2022).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  113. Uyehara, C. M. & Apostolou, E. 3D Enhancer–promoter interactions and multi-connected hubs: organizational principles and functional roles. Cell Rep. 42, 112068 (2023).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  114. Garg, V. et al. Single-cell analysis of bidirectional reprogramming between early embryonic states reveals mechanisms of differential lineage plasticities. Preprint at bioRxiv https://doi.org/10.1101/2023.03.28.534648 (2023).

  115. Niakan, K. K. et al. Novel role for the orphan nuclear receptor Dax1 in embryogenesis, different from steroidogenesis. Mol. Genet. Metab. 88, 261–271, https://doi.org/10.1016/j.ymgme.2005.12.010 (2006).

    Article  CAS  PubMed  Google Scholar 

  116. Concordet, J. P. & Haeussler, M. CRISPOR: intuitive guide selection for CRISPR/Cas9 genome editing experiments and screens. Nucleic Acids Res. 46, W242–W245, https://doi.org/10.1093/nar/gky354 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  117. Buenrostro, J. D., Wu, B., Chang, H. Y. & Greenleaf, W. J. ATAC-seq: a method for assaying chromatin accessibility genome-wide. Curr. Protoc. Mol. Biol. 109, 21.29.1–21.29.9, https://doi.org/10.1002/0471142727.mb2129s109 (2015).

    Article  PubMed  Google Scholar 

  118. Krijger, P. H. L., Geeven, G., Bianchi, V., Hilvering, C. R. E. & de Laat, W. 4C-seq from beginning to end: a detailed protocol for sample preparation and data analysis. Methods 170, 17–32, https://doi.org/10.1016/j.ymeth.2019.07.014 (2020).

    Article  CAS  PubMed  Google Scholar 

  119. Anders, S., Pyl, P. T. & Huber, W. HTSeq-A Python framework to work with high-throughput sequencing data. Bioinformatics 31, 166–169, https://doi.org/10.1093/bioinformatics/btu638 (2015).

    Article  CAS  PubMed  Google Scholar 

  120. Hounkpe, B. W., Chenou, F., de Lima, F. & de Paula, E. V. HRT Atlas v1.0 database: redefining human and mouse housekeeping genes and candidate reference transcripts by mining massive RNA-seq datasets. Nucleic Acids Res. 49, D947–D955, https://doi.org/10.1093/nar/gkaa609 (2021).

    Article  CAS  PubMed  Google Scholar 

  121. Langmead, B. & Salzberg, S. L. Fast gapped-read alignment with Bowtie 2. Nat. Methods 9, 357–359, https://doi.org/10.1038/nmeth.1923 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  122. Li, H. et al. The sequence alignment/map format and SAMtools. Bioinformatics 25, 2078–2079, https://doi.org/10.1093/bioinformatics/btp352 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  123. Quinlan, A. R. & Hall, I. M. BEDTools: a flexible suite of utilities for comparing genomic features. Bioinformatics 26, 841–842, https://doi.org/10.1093/bioinformatics/btq033 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  124. Lazaris, C., Kelly, S., Ntziachristos, P., Aifantis, I. & Tsirigos, A. HiC-bench: comprehensive and reproducible Hi-C data analysis designed for parameter exploration and benchmarking. BMC Genomics 18, 22, https://doi.org/10.1186/s12864-016-3387-6 (2017).

    Article  PubMed  PubMed Central  Google Scholar 

  125. Durand, N. C. et al. Juicer provides a one-click system for analyzing loop-resolution Hi-C experiments. Cell Syst. 3, 95–98, https://doi.org/10.1016/j.cels.2016.07.002 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  126. Zheng, X. & Zheng, Y. CscoreTool: fast Hi-C compartment analysis at high resolution. Bioinformatics 34, 1568–1570, https://doi.org/10.1093/bioinformatics/btx802 (2018).

    Article  CAS  PubMed  Google Scholar 

  127. Kent, W. J., Zweig, A. S., Barber, G., Hinrichs, A. S. & Karolchik, D. BigWig and BigBed: enabling browsing of large distributed datasets. Bioinformatics 26, 224–7 (2010).

    Article  Google Scholar 

  128. Mumbach, M. R. et al. Enhancer connectome in primary human cells identifies target genes of disease-associated DNA elements. Nat. Genet. 49, 1602–1612 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  129. Rubin, A. J. et al. Coupled single-cell CRISPR screening and epigenomic profiling reveals causal gene regulatory networks. Cell 176, 361–376.e17 (2019).

    Article  CAS  PubMed  Google Scholar 

Download references

Acknowledgements

We are grateful to all members of the Apostolou and Stadtfeld groups for critical reading of the manuscript and input. We also thank J. Pulecio and D. Huangfu for advice on the functional experiments and C. Leslie for advice on the modeling. Additionally, we are thankful to the reviewers for their insightful and constructive criticism to improve the manuscript. This work was partly supported by a HiChIP research grant from Arima Genomics. D.M. was supported by the T32 HD060600. E.A. is a recipient of the Mark Foundation Emerging Leader Award and supported by the National Institutes of Health (1R01GM138635, 1U01DK128852, RM1GM139738) and the Tri-Institutional Stem Cell Initiative of the Starr Foundation.

Author information

Authors and Affiliations

Authors

Contributions

E.A. and A.P. conceived and designed the study and analyses with input from D.M., E.S., M.S., A.K.H. and A.T. All genomic and functional experiments were performed by D.M., E.S. and D.C.G. V.G. provided help with TSCs and XEN cell lines and L.E. provided material for the EpiSC genomics experiments. C.M.U. assisted with HiChIP visualization. U.L. assisted with CTCF ChIP-exo in ESCs. A.P. performed all computational analyses with help from J.R.H., A.K. and guidance from A.T. and E.A. E.C. performed all gene ontology analyses. E.A. wrote the manuscript together with D.M., E.S. and A.P. and input from all authors.

Corresponding authors

Correspondence to Aristotelis Tsirigos, Alexander Polyzos or Effie Apostolou.

Ethics declarations

Competing interests

The authors declare no competing interests.

Peer review

Peer review information

Nature Structural & Molecular Biology thanks Alvaro Rada-Iglesias, Pedro Rocha and the other, anonymous, reviewer(s) for their contribution to the peer review of this work. Tiago Faial, Carolina Perdigoto and Dimitris Typas were the primary editors on this article and managed its editorial process and peer review in collaboration with the rest of the editorial team.

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 Related to Figure 1.

a. Representative single (xy) stack epifluorescence images of immunofluorescence experiments showing expression of key lineage markers (greyscale) in TSC, ESC and XEN cells. Cells were counterstained with DAPI (blue) for DNA content. n = 3 independent experiments. Scale bar 100μm. b. Principal component analysis (PCA) of all TSC, ESC and XEN replicates based on their RNA-seq, ATAC-seq and H3K27ac ChIP-seq profiles. PCA plots were designed based on the top10% of most variable genes or peaks in all three cell lines. In each plot, circles indicate the experimental data presented in this study, while squares and triangles correspond to publicly available RNA-seq data (Supplementary Table 7) or independent -unpublished- studies from our lab, respectively. c. Stacked barplot showing the distribution of H3K27 occupancy among intergenic regions, gene bodies or TSS (promoter +/- 1.5 kb) for each K-Mean cluster as identified in Fig. 1c. Note: all statistics are provided in Supplementary Table 8.

Extended Data Fig. 2 Related to Figure 2.

a. Principal Component Analysis (PCA) plot of all lineages and replicates based on their compartment scores at 100 kb resolution (top) and on their TAD insulation levels at 40 kb resolution (bottom). b. Boxplots showing median expression changes between ESC and TSC (n = 4,327 genes), and TSC and XEN (n = 4156 genes) cells of genes in unaltered compartments (grey box and dashed line) or compartments with shifts as described in Fig. 2b. Asterisks indicate significance (p < 0.001) by two-sided Wilcoxon rank test. c. Volcano plot showing differential Hi-C interactivity at 40 kb resolution between ESC – TSC and TSC - XEN. X-axis shows delta interactivity while y-axis shows -log10(p-value) calculated by two-sided Student’s t-test. Significant changes (p-value < 0.05 and Diff>0.1 or <-0.1) are noted with blue and red color. d. Boxplots showing gene expression (n = 2352 genes) and enhancer strength (n = 15,982 peaks) changes between ESC-TSC regions with connectivity changes as described in Fig. 2c. Asterisks indicate significance (p < 0.001) by two-sided Student’s t-test. e. Boxplot comparing the sizes of HiChIP-detected loops in the three cell lineages across n = 2 independent Hi-C samples. f. Aggregate peak analysis (APA) showing the aggregate signal of MicroC data in ESC33 centered around ESC HiChIP interacting regions identified by FitHiChIP v9.0 at 5 kb resolution. (See Supplementary Table 4 and Methods). g. IGV tracks aligning H3K27ac HiChIP results (arcs on top and virtual 4C representation in the middle) with 4C-seq normalized signals around PDGFRA promoter in XEN along with corresponding H3K27ac ChIP-seq occupancy. h. Boxplot showing the median expression levels of a curated list of skipped and looped genes in ESC, XEN and TSC across n = 2 independent HiChIP and RNA-seq samples. Selected genes have similar ranges of H3K27ac signal at promoters. Asterisks indicate significance (p-value < 0.05), as calculated by two-sided Wilcoxon rank sum test. (See Supplementary Table 4). Note: all statistics are provided in Supplementary Table 8.

Extended Data Fig. 3 Related to Figure 3.

a. HiGlass visualization of H3K27ac HiChIP results around a TSC related hub (Cdx2) and a XEN-related hub (Gata6) in TSC, ESC, and XEN along with the corresponding H3K27ac HiChIP derived arcs and H3K27ac ChIP-seq signals. Interacting scores are presented in 5 kb resolution. b. Barplot showing the percentages of essential genes -as identified in two recent studies83,84- within the least (Q1) versus most (Q10) connected hubs. The preferential enrichment of essential genes in Q10 is significant (p-value < 0.001, two-sided Fisher’s exact test). c. Stacked barplots showing the proportions of different HiChIP loop subtypes in TSC, ESC and XEN cells. Loops were separated into 5 chromatin interaction categories based on the presence of regulatory elements, such as promoter/TSS (P) or putative enhancer (E, H3K27ac peak). X- anchors were defined as anchors that do not contain any TSS nor an H3K27ac peak. d. Boxplot showing the size distribution of X loops (X-E and X-P) compared to E-E, E-P and P-P loops in all cell lines. (n = 60,909 (TSC), 81,679 (ESC), 77,124 (TSC) loops across n = 2 independent HiChIP experiments). e. Boxplots showing expression changes between any two cell types around multiconnected genes (n > =5 in both cell types of interest), when at least one of their conserved anchors switches chromatin states: either from X-to-E (enhancer gain) or from E-to-X (enhancer loss) Asterisks indicate significance < 0.05 by two-sided Wilcoxon rank-sum test. (See also Supplementary Table 8). Note: all statistics are provided in Supplementary Table 8.

Extended Data Fig. 4 Related to Figure 4.

a. Correlation between differential HiChIP connectivity/hubness and differential gene expression in connectivity and differential gene expression in ESC and TSC cells (top) and TSC and XEN cells (bottom). R represents Spearman correlation identifies distinct groups of genes. We focus on the two most prominent groups: 3D-insensitive genes, defined as genes with differential connectivity >3 but no transcriptional changes (log2FC < 1 or >-1) and 3D-concordant genes for which connectivity and expression changes (log2FC > 1 or <-1) positively correlate (Supplementary Table 5). b-e. Gene ontology analysis depicting the most significant biological processes enriched in the 3D concordant and 3D-insensitive genes in each pairwise comparison (ESC vs TSC and TSC vs XEN) as defined in (a). All genes in A compartments were used as background. For further details see also Supplementary Table 3. f. Comparison of connectivity, gene expression levels as well as H3K27ac and ATAC CPM levels between ESC and TSC cells (left) and TSC and XEN cells (right) at promoters of 3D-concordant (n = 1818 for ESC/TSC and n = 1108 for TSC/XEN) and 3D-insensitive genes (n = 2637 for ESC/TSC, n = 2230 for TSC/XEN) as defined in (a). Insensitive genes show higher levels of connectivity, H3K27ac, ATAC and expression in both cell types. X-axis indicates cell type (T = TSC, E = ESC, X = XEN). Two-sided Wilcoxon rank sum test was used for all comparisons with p > 0.05 indicating significance. (Supplementary Table 8). Note: all statistics are provided in Supplementary Table 8.

Extended Data Fig. 5 Related to Figure 5.

a. Barplot of feature importance showing 10 1D (pink) and 3D (blue) features, ranked from high to low. Light blue indicates features not selected. (See Supplementary Table 6). b. Spearman correlation values for each variable considered for our 3D model with gene expression (left) and differential expression (right). Dots represent minimum, mean and maximum correlation scores. (See Supplementary Table 6). c. Area Under Curve (AUC) scores and Spearman Correlation for classifying gene expression (top 10% high vs low, left graph) and predicting levels (right graph) in ESC or TSC cells using 3D-HiChAT, Promoter and Linear models. Dots represent average scores from LOCO training approach (n = 20). Error bars show standard deviation. (See Extended Data Fig. 5, Supplementary Table 6). d. Plots showing AUC and Spearman correlation for classifying gene expression (top 10% high vs low, left graph) and predicting levels (right graph) using 3D-HiChAT in various lineages including mouse lineages and published human data: Naïve T cells, T-Helper 17 Cells (Th17), and T regulatory cells (Tregs)128,129. e. AUC scores and Spearman Correlation generated for classifying differential expression (top 10% up or downregulated, left) and predicting expression changes (right) between XEN and ESC using 3D-HiChAT, Promoter and Linear models. Dots represent average scores from LOCO training approach (n = 20). Error bars show standard deviation. (See Supplementary Table 6). f. Ranked perturbation scores (%) predicted by in silico perturbations of ~46 K E-P pairs in ESC, ~46.7 K in TSC and ~53.1 K in XEN using 3D-HiChAT. Dotted horizontal lines indicate selected cut-offs for impactful perturbations. g. Scatterplot comparing predicted perturbation scores from 3D-HiChAT with respective ABC scores. R Spearman correlation values are shown on the top. h. Boxplots showing that enhancers with high 3D-HiChAT-predicted perturbation scores and low ABC scores (red) are more distal to their target genes (loop size) than those with high scores in both models (blue) (left plot n = 3,428 enhancers). Enhancers/anchors with high 3D-HiChAT scores are more distal to the ones with high ABC scores (>0.7) (right plot n = 8,445 enhancers). Asterisks indicates significance calculated by two-sided Wilcoxon rank-sum test, p-val<0.001. Note: all statistics are provided in Supplementary Table 8.

Extended Data Fig. 6 Related to Figure 6.

a. Visualization of the Tfcp2l1 Locus showing H3K27ac HiChIP arcs, H3K27ac ChIP and Compartment c-scores called by Hi-C for TSC, ESC, and XEN. Notably, a group of putative enhancers upstream of Gli2 are uniquely expressed and only in an A compartment in ESC. b. IGV tracks of the Tfcp2l1-Gli2 locus showing the two enhancers chosen for functional validation, Enh3 and Enh14. H3K27ac HiChIP derived arcs originating from both enhancers are shown as well. RT-qPCR showing relative expression levels of Tfcp2l and Gli2 upon CRISPRi perturbation of Enh3 compared to control cells infected with empty vector (EV). Dots indicate independent experiments (n = 3). Error bars represent mean ± SD. Asterisks indicate significance, with p-value < 0.05, as calculated using unpaired one-tailed t-test. c. Schematic showing experimental strategy for generating a stable XEN line expressing dCas-BFP-KRAB (CRISPRi). Representative FACs plot from n > 12 independent experiments. d. AUC curve (red) showing a value of 0.71 when comparing our precited perturbation scores to our experimental validations presented in Fig. 6i for n = 40 different E-P pairs. e. Scatter plot comparing the predicted perturbation scores and the ABC scores for each of the 40 experimentally tested E-P pairs. Spearman Correlation value of -0.49. Different colors indicate different groups reflecting the concordance or discordance between predictions and experimental validations as shown in Fig. 6i. TP: true positive, TN: true negative, FP: false positive, FN: false negative. Note: all statistics are provided in Supplementary Table 8.

Supplementary information

Supplementary Information

Supplementary methods with the corresponding references

Reporting Summary

Supplementary Table 1

Genomic QCs

Supplementary Table 2

List of H3K27ac k-means peaks and super enhancers in each cell type

Supplementary Table 3

Summary of gene ontology and motif enrichment analyses

Supplementary Table 4

Summary of H3K27ac HiChIP contacts and hubs

Supplementary Table 5

List of 3D-concordant, discordant, and 3D-insensitive genes in each pairwise comparison and their corresponding gene ontology analysis.

Supplementary Table 6

Information about our 2D and 3D models: list of variables and models, performance and predictions from in silico perturbations.

Supplementary Table 7

List of oligonucleotides, reagents and resources

Supplementary Table 8

Statistical analyses for all figures

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

Murphy, D., Salataj, E., Di Giammartino, D.C. et al. 3D Enhancer–promoter networks provide predictive features for gene expression and coregulation in early embryonic lineages. Nat Struct Mol Biol 31, 125–140 (2024). https://doi.org/10.1038/s41594-023-01130-4

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41594-023-01130-4

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