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:

The structural basis for cohesin–CTCF-anchored loops

Abstract

Cohesin catalyses the folding of the genome into loops that are anchored by CTCF1. The molecular mechanism of how cohesin and CTCF structure the 3D genome has remained unclear. Here we show that a segment within the CTCF N terminus interacts with the SA2–SCC1 subunits of human cohesin. We report a crystal structure of SA2–SCC1 in complex with CTCF at a resolution of 2.7 Å, which reveals the molecular basis of the interaction. We demonstrate that this interaction is specifically required for CTCF-anchored loops and contributes to the positioning of cohesin at CTCF binding sites. A similar motif is present in a number of established and newly identified cohesin ligands, including the cohesin release factor WAPL2,3. Our data suggest that CTCF enables the formation of chromatin loops by protecting cohesin against loop release. These results provide fundamental insights into the molecular mechanism that enables the dynamic regulation of chromatin folding by cohesin and CTCF.

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: Structure of the SA2–SCC1–CTCF complex.
Fig. 2: CTCF interaction stabilizes cohesin on DNA.
Fig. 3: CTCF–CES interaction is required for CTCF-anchored loops.
Fig. 4: CTCF–CES interaction promotes localization of cohesin to CTCF sites.

Similar content being viewed by others

Data availability

Coordinates are available from the PDB under accession number 6QNX for the SA2–SCC1–CTCF complex. The generated Hi-C, RNA sequencing and ChIP–seq data have been deposited in GEO, accession number GSE126637. Any other relevant data are available from the corresponding authors upon reasonable request.

References

  1. Dekker, J. & Mirny, L. The 3D genome as moderator of chromosomal communication. Cell 164, 1110–1121 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  2. Hara, K. et al. Structure of cohesin subcomplex pinpoints direct shugoshin–Wapl antagonism in centromeric cohesion. Nat. Struct. Mol. Biol. 21, 864–870 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  3. Shintomi, K. & Hirano, T. Releasing cohesin from chromosome arms in early mitosis: opposing actions of Wapl–Pds5 and Sgo1. Genes Dev. 23, 2224–2236 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  4. Merkenschlager, M. & Nora, E. P. CTCF and cohesin in genome folding and transcriptional gene regulation. Annu. Rev. Genomics Hum. Genet. 17, 17–43 (2016).

    Article  CAS  PubMed  Google Scholar 

  5. Rowley, M. J. & Corces, V. G. Organizational principles of 3D genome architecture. Nat. Rev. Genet. 19, 789–800 (2018).

    Article  CAS  PubMed  Google Scholar 

  6. Yatskevich, S., Rhodes, J. & Nasmyth, K. Organization of chromosomal DNA by SMC complexes. Annu. Rev. Genet. 53, 445–482 (2019).

    Article  CAS  PubMed  Google Scholar 

  7. Alipour, E. & Marko, J. F. Self-organization of domain structures by DNA-loop-extruding enzymes. Nucleic Acids Res. 40, 11202–11212 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  8. Fudenberg, G. et al. Formation of chromosomal domains by loop extrusion. Cell Rep. 15, 2038–2049 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  9. Sanborn, A. L. et al. Chromatin extrusion explains key features of loop and domain formation in wild-type and engineered genomes. Proc. Natl Acad. Sci. USA 112, E6456–E6465 (2015).

    CAS  PubMed  PubMed Central  Google Scholar 

  10. Haarhuis, J. H. I. et al. The cohesin release factor WAPL restricts chromatin loop extension. Cell 169, 693–707 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  11. Nora, E. P. et al. Targeted degradation of CTCF decouples local insulation of chromosome domains from genomic compartmentalization. Cell 169, 930–944 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  12. Wutz, G. et al. Topologically associating domains and chromatin loops depend on cohesin and are regulated by CTCF, WAPL, and PDS5 proteins. EMBO J. 36, 3573–3599 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  13. Guo, Y. et al. CRISPR inversion of CTCF sites alters genome topology and enhancer/promoter function. Cell 162, 900–910 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  14. de Wit, E. et al. CTCF binding polarity determines chromatin looping. Mol. Cell 60, 676–684 (2015).

    Article  PubMed  CAS  Google Scholar 

  15. Rao, S. S. 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 

  16. Vietri Rudan, M. et al. Comparative Hi-C reveals that CTCF underlies evolution of chromosomal domain architecture. Cell Rep. 10, 1297–1309 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

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

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  18. Gassler, J. et al. A mechanism of cohesin-dependent loop extrusion organizes zygotic genome architecture. EMBO J. 36, 3600–3618 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  19. Rubio, E. D. et al. CTCF physically links cohesin to chromatin. Proc. Natl Acad. Sci. USA 105, 8309–8314 (2008).

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  20. Xiao, T., Wallace, J. & Felsenfeld, G. Specific sites in the C terminus of CTCF interact with the SA2 subunit of the cohesin complex and are required for cohesin-dependent insulation activity. Mol. Cell. Biol. 31, 2174–2183 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  21. Pezzi, N. et al. STAG3, a novel gene encoding a protein involved in meiotic chromosome pairing and location of STAG3-related genes flanking the Williams-Beuren syndrome deletion. FASEB J. 14, 581–592 (2000).

    Article  CAS  PubMed  Google Scholar 

  22. Orgil, O. et al. A conserved domain in the Scc3 subunit of cohesin mediates the interaction with both Mcd1 and the cohesin loader complex. PLoS Genet. 11, e1005036 (2015).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  23. Roig, M. B. et al. Structure and function of cohesin’s Scc3/SA regulatory subunit. FEBS Lett. 588, 3692–3702 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  24. Forbes, S. A. et al. COSMIC: somatic cancer genetics at high-resolution. Nucleic Acids Res. 45, D777–D783 (2017).

    Article  CAS  PubMed  Google Scholar 

  25. Beckouët, F. et al. Releasing activity disengages cohesin’s Smc3/Scc1 interface in a process blocked by acetylation. Mol. Cell 61, 563–574 (2016).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  26. Gandhi, R., Gillespie, P. J. & Hirano, T. Human Wapl is a cohesin-binding protein that promotes sister-chromatid resolution in mitotic prophase. Curr. Biol. 16, 2406–2417 (2006).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  27. Kueng, S. et al. Wapl controls the dynamic association of cohesin with chromatin. Cell 127, 955–967 (2006).

    Article  CAS  PubMed  Google Scholar 

  28. Liu, H., Rankin, S. & Yu, H. Phosphorylation-enabled binding of SGO1–PP2A to cohesin protects sororin and centromeric cohesion during mitosis. Nat. Cell Biol. 15, 40–49 (2013).

    Article  CAS  PubMed  Google Scholar 

  29. Ouyang, Z. et al. Structure of the human cohesin inhibitor Wapl. Proc. Natl Acad. Sci. USA 110, 11355–11360 (2013).

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  30. Krystkowiak, I. & Davey, N. E. SLiMSearch: a framework for proteome-wide discovery and annotation of functional modules in intrinsically disordered regions. Nucleic Acids Res. 45, W464–W469 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

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

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

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

    Article  ADS  CAS  PubMed  Google Scholar 

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

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  34. Ouyang, Z., Zheng, G., Tomchick, D. R., Luo, X. & Yu, H. Structural basis and IP6 requirement for Pds5-dependent cohesin dynamics. Mol. Cell 62, 248–259 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  35. Chan, K. L. et al. Cohesin’s DNA exit gate is distinct from its entrance gate and is regulated by acetylation. Cell 150, 961–974 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  36. Buheitel, J. & Stemmann, O. Prophase pathway-dependent removal of cohesin from human chromosomes requires opening of the Smc3–Scc1 gate. EMBO J. 32, 666–676 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  37. Eichinger, C. S., Kurze, A., Oliveira, R. A. & Nasmyth, K. Disengaging the Smc3/kleisin interface releases cohesin from Drosophila chromosomes during interphase and mitosis. EMBO J. 32, 656–665 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  38. Sedeño Cacciatore, Á. & Rowland, B. D. Loop formation by SMC complexes: turning heads, bending elbows, and fixed anchors. Curr. Opin. Genet. Dev. 55, 11–18 (2019).

    Article  PubMed  CAS  Google Scholar 

  39. Tang, Z. et al. CTCF-mediated human 3D genome architecture reveals chromatin topology for transcription. Cell 163, 1611–1627 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  40. Nagy, G. et al. Motif oriented high-resolution analysis of ChIP-seq data reveals the topological order of CTCF and cohesin proteins on DNA. BMC Genomics 17, 637 (2016).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  41. Kschonsak, M. et al. Structural basis for a safety-belt mechanism that anchors condensin to chromosomes. Cell 171, 588–600 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  42. Ganji, M. et al. Real-time imaging of DNA loop extrusion by condensin. Science 360, 102–105 (2018).

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  43. Li, Y. et al. Structural basis for Scc3-dependent cohesin recruitment to chromatin. eLife 7, e38356 (2018).

    Article  PubMed  PubMed Central  Google Scholar 

  44. Studier, F. W. Protein production by auto-induction in high density shaking cultures. Protein Expr. Purif. 41, 207–234 (2005).

    Article  CAS  PubMed  Google Scholar 

  45. Bowler, M. W. et al. MASSIF-1: a beamline dedicated to the fully automatic characterization and data collection from crystals of biological macromolecules. J. Synchrotron Radiat. 22, 1540–1547 (2015).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  46. Svensson, O., Malbet-Monaco, S., Popov, A., Nurizzo, D. & Bowler, M. W. Fully automatic characterization and data collection from crystals of biological macromolecules. Acta Crystallogr. D 71, 1757–1767 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  47. Svensson, O., Gilski, M., Nurizzo, D. & Bowler, M. W. Multi-position data collection and dynamic beam sizing: recent improvements to the automatic data-collection algorithms on MASSIF-1. Acta Crystallogr. D 74, 433–440 (2018).

    Article  CAS  Google Scholar 

  48. Kabsch, W. Integration, scaling, space-group assignment and post-refinement. Acta Crystallogr. D 66, 133–144 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  49. Winn, M. D. et al. Overview of the CCP4 suite and current developments. Acta Crystallogr. D 67, 235–242 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  50. McCoy, A. J. et al. Phaser crystallographic software. J. Appl. Crystallogr. 40, 658–674 (2007).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  51. Emsley, P., Lohkamp, B., Scott, W. G. & Cowtan, K. Features and development of Coot. Acta Crystallogr. D 66, 486–501 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  52. Adams, P. D. et al. PHENIX: a comprehensive Python-based system for macromolecular structure solution. Acta Crystallogr. D 66, 213–221 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  53. Chen, V. B. et al. MolProbity: all-atom structure validation for macromolecular crystallography. Acta Crystallogr. D 66, 12–21 (2010).

    Article  CAS  PubMed  Google Scholar 

  54. Yin, M. et al. Molecular mechanism of directional CTCF recognition of a diverse range of genomic sites. Cell Res. 27, 1365–1377 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  55. Rhodes, J. D. P. et al. Cohesin can remain associated with chromosomes during DNA replication. Cell Rep. 20, 2749–2755 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  56. Servant, N. et al. HiC-Pro: an optimized and flexible pipeline for Hi-C data processing. Genome Biol. 16, 259 (2015).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  57. Yang, T. et al. HiCRep: assessing the reproducibility of Hi-C data using a stratum-adjusted correlation coefficient. Genome Res. 27, 1939–1949 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  58. Durand, N. C. et al. Juicer provides a one-click system for analyzing loop-resolution Hi-C experiments. Cell Syst. 3, 95–98 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  59. Imakaev, M. et al. Iterative correction of Hi-C data reveals hallmarks of chromosome organization. Nat. Methods 9, 999–1003 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  60. Lévy-Leduc, C., Delattre, M., Mary-Huard, T. & Robin, S. Two-dimensional segmentation for analyzing Hi-C data. Bioinformatics 30, i386–i392 (2014).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  61. Crane, E. et al. Condensin-driven remodelling of X chromosome topology during dosage compensation. Nature 523, 240–244 (2015).

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  62. Flyamer, I. M. et al. Single-nucleus Hi-C reveals unique chromatin reorganization at oocyte-to-zygote transition. Nature 544, 110–114 (2017).

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

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

    Article  Google Scholar 

  64. Langmead, B. & Salzberg, S. L. Fast gapped-read alignment with Bowtie 2. Nat. Methods 9, 357–359 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  65. Ramírez, F., Dündar, F., Diehl, S., Grüning, B. A. & Manke, T. deepTools: a flexible platform for exploring deep-sequencing data. Nucleic Acids Res. 42, W187–W191 (2014).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  66. Feng, J., Liu, T., Qin, B., Zhang, Y. & Liu, X. S. Identifying ChIP–seq enrichment using MACS. Nat. Protoc. 7, 1728–1740 (2012).

    Article  CAS  PubMed  Google Scholar 

  67. Quinlan, A. R. & Hall, I. M. BEDTools: a flexible suite of utilities for comparing genomic features. Bioinformatics 26, 841–842 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  68. Amemiya, H.M., Kundaje, A., & Boyle, A.P. The ENCODE blacklist: identification of problematic regions of the genome. Sci Rep. 9, 9354 (2019).

    Article  ADS  PubMed  PubMed Central  CAS  Google Scholar 

  69. Mathelier, A. et al. JASPAR 2014: an extensively expanded and updated open-access database of transcription factor binding profiles. Nucleic Acids Res. 42, D142–D147 (2014).

    Article  CAS  PubMed  Google Scholar 

  70. Grant, C. E., Bailey, T. L. & Noble, W. S. FIMO: scanning for occurrences of a given motif. Bioinformatics 27, 1017–1018 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  71. Kim, D. et al. TopHat2: accurate alignment of transcriptomes in the presence of insertions, deletions and gene fusions. Genome Biol. 14, R36 (2013).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  72. Anders, S., Pyl, P. T. & Huber, W. HTSeq—a Python framework to work with high-throughput sequencing data. Bioinformatics 31, 166–169 (2015).

    Article  CAS  PubMed  Google Scholar 

  73. 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  CAS  Google Scholar 

  74. Landgraf, C. et al. Protein interaction networks by proteome peptide scanning. PLoS Biol. 2, e14 (2004).

    Article  PubMed  PubMed Central  Google Scholar 

Download references

Acknowledgements

This work was funded by EMBL. J.H.I.H., Á.S.C. and B.D.R. were supported by an ERC CoG (772471 ‘CohesinLooping’), M.S.v.R. was supported by the Boehringer Ingelheim Fonds and H.T. and E.d.W. were supported by an ERC StG (637587 ‘HAP-PHEN’). H.T. and E.d.W. are part of the Oncode Institute, which is partly financed by the Dutch Cancer Society. We thank the staff at the ESRF beamline Massif-1; T. Gibson for advice concerning short linear motifs; J. Rhodes and K. Nasmyth for reagents and advice on Halo tagging; R. van der Weide for advice and bioinformatic analyses; and R. Kerkhoven and the NKI Genomics Core Facility for sequencing.

Author information

Authors and Affiliations

Authors

Contributions

Y.L. and K.W.M. initiated the project and proposed the CES motif. Y.L. performed biochemical studies and structural analyses with support from K.W.M. J.H.I.H., R.O., M.S.v.R., L.W. and H.T. performed wet-laboratory cell-based experiments and Á.S.C. performed bioinformatic analyses. K.W.M., E.d.W., B.D.R. and D.P. provided supervision. Y.L., K.W.M., B.D.R. and D.P. were involved in conceptualization, project administration and wrote the original and revised draft with input from all authors.

Corresponding authors

Correspondence to Kyle W. Muir, Elzo de Wit, Benjamin D. Rowland or Daniel Panne.

Ethics declarations

Competing interests

The authors declare no competing interests.

Additional information

Peer review information Nature thanks Victor Corces, Karl-Peter Hopfner and the other, anonymous, reviewer(s) for their contribution to the peer review of this work.

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

Extended data figures and tables

Extended Data Fig. 1 Biochemical analysis of CTCF binding to SA2–SCC1.

a, Domain architecture of CTCF. CTCF fragments tested for SA2–SCC1 binding by GST pulldown analysis are indicated. The region that retains SA2–SCC1 is highlighted in magenta. b, Summary data showing results of GST pulldowns. The input and the bound fractions were analysed by SDS–PAGE. CTCF fragments that bind SA2–SCC1 are shown in magenta. The experiment was repeated once. c, ITC curves. The binding stoichiometry (N) and dissociation constants (Kd) are indicated. The experiment was repeated three times, with consistent results. d, Fo − Fc omit electron-density Fourier map contoured at 3σ. e, LIGPLOT representation of the interaction between the CTCF peptide and SA2–SCC1. The CTCF peptide is shown in magenta, SA2 in blue and SCC1 in green bonds.

Extended Data Fig. 2 Analysis of the SA2–SCC1–CTCF structure.

a, Multiple sequence alignment of SA2 (here denoted STAG2) orthologues and paralogues. *Key amino-acid residues that engage CTCF. b, Missense mutation frequencies plotted onto the SA2 structure. R370 (a hotspot in SA2) is indicated. The inset shows an overview of the mutation hotspots R370 of SA2), Y226 and F228 of CTCF, and S334, K335, R338 and L341 of SCC). c, ITC progress curves of binding between WAPL(423–463) and SA2–SCC1. d, Competition between SGO1 and CTCF for SA2–SCC1 binding. SA2–SCC1 was incubated with GST–CTCF(86–267). Increasing amounts (lanes 4–8) (molar ratios are indicated) of the SGO1 phosphorylated at T346 peptide (spanning residues 331–349) were added and the input and the bound fraction analysed by SDS–PAGE. The experiment was repeated twice. One representative example is shown. e, Domain architecture and sequence alignments of cohesin regulators that contain F/YXF motifs. Putative CES-interacting residues are highlighted in red. f, Regular expression motif used to query the human and yeast proteomes for factors containing F/YXF motifs. Regular expression syntax: letters denote a specific amino acid; square brackets denote a subset of allowed amino acids; curly brackets denote length variability.

Extended Data Fig. 3 Generation of CTCFY226A/F228A cells.

a, Schematic of CRISPR–Cas9-based generation of CTCFY226A/F228A cells. The guide targets cleavage of exon 1 of the CTCF gene. The repair oligonucleotide renders the gene noncleavable by Cas9, and simultaneously introduces mutations in the codons that encode Y226 and F228. b, The CTCFY226A/F228A mutation was confirmed by Sanger sequencing, including a silent mutation at position 229. c, Western blot depicting Halo-tagged SCC1 in wild-type and CTCFY226A/F228A cells. The parental wild-type cells are included as a control. This experiment was performed once. d, Representative images of cells in G1 and G2, as indicated by their nuclear and cytoplasmic localization of DHB–iRFP, respectively. e, Chromatin-bound levels of CTCF and SMC1 analysed by western blot. Histone H4 is used as a control for the chromatin fraction. The CTCFY226A/F228A mutation does not evidently affect overall CTCF and cohesin levels on chromatin. WCE, whole-cell extract; CB, chromatin-bound fraction. This experiment was performed twice with similar results. f, Relative SCC1–Halo fluorescence intensity quantified in the unbleached area directly after photobleaching, as a proxy for the chromatin-bound fraction of SCC1. This nondiffusive fraction is not evidently affected by the CTCFY226A/F228A mutation. Individual cells of three independent experiments are plotted as dots and their mean is indicated (21 wild-type cells and 17 CTCFY226A/F228A cells were scored).

Extended Data Fig. 4 TAD analyses and Hi-C replicates.

a, Schematic of a Hi-C matrix displaying DNA–DNA contacts across a genomic region that includes two TADs. TADs in general are flanked by inwards-pointing CTCF sites (magenta arrows). Signal close to the diagonal line reflects short-range contacts, and contacts that span longer distances are found further away from the diagonal. The contacts within a TAD are formed by cohesin complexes (blue circles). Cohesin builds loops that it can enlarge until it encounters CTCF. Some TADs are enriched for contacts between the two CTCF sites that lie at their boundaries. These contacts are referred to as CTCF-anchored loops. b, Aggregate TAD analysis depicting the average contact frequency across TADs defined in wild-type cells. c, Heat map of the insulation score61 at TAD borders, as defined for wild-type cells. d, Aggregate peak analysis as in Fig. 3c, using two independent library preparations per genotype. e, Aggregate TAD analysis for wild-type and CTCFY226A/F228A cells as in b. f, Heat map of insulation scores at TAD borders for wild-type and CTCFY226A/F228A cells as in c.

Extended Data Fig. 5 CTCFY226A/F228A mutation has little effect on CTCF levels at CTCF sites.

a, Hi-C contact matrix of region chromosome 16: 77000000–78300000 at 10-kb resolution for the wild-type cell line (bottom triangles) and the CTCFY226A/F228A cell line (top triangles). CTCF sites are depicted below; those selected for qPCR are shown in colour. Red triangles indicate sites with a forward motif and blue triangles indicate sites with a reverse motif. The numbers underneath indicate the qPCR primer pairs shown in b. Primer pair 11 (indicated with *) is at a locus devoid of SCC1 and CTCF. b, ChIP–qPCR analysis of SCC1 (cohesin) enrichment at the aforementioned CTCF sites and control locus (*) in wild-type and CTCFY226A/F228A cells. The mean of three independent ChIP experiments is shown with the s.d. c, ChIP–seq tracks for SCC1 and CTCF at region chromosome 16: 77000000–78300000 in wild-type and CTCFY226A/F228A cells. The loci used for ChIP–qPCR analysis are indicated below the SCC1 ChIP–seq tracks. RPKM, reads per kilobase per million reads. d, ChIP–qPCR analysis of CTCF abundance at loci 1–7, as described in Fig. 3d. Analysis includes IgG as a control. The mean of two independent ChIP experiments is shown. Details of replicates are given in the Methods. e, ChIP–qPCR analysis of CTCF abundance at loci 8–12, as described in Extended Data Fig. 4a. Analysis includes IgG as a control. The mean of two independent ChIP experiments is shown. Details of replicates are given in the extended methods.

Extended Data Fig. 6 Compartmentalization is largely maintained in cells that contain the CTCFY226A/F228A mutation.

a, Hi-C contact matrices of the q-arm of chromosome 2 at 500-kb resolution. The corresponding compartment scores are plotted above. b, Genome-wide comparison of compartment scores for wild-type and CTCFY226A/F228A cells. Pearson correlation = 0.97. c, Saddle plots representing the interaction between A and B compartments. d, A region of chromosome 1 (55500000–59500000) at 10-kb resolution that contains no obvious CTCF-anchored loops. e, Relative contact probability profiles for wild-type and CTCFY226A/F228A mutant cells (left), compared to previously published12 contact profiles upon degradation of CTCF (middle) or SCC1 (right). The contact probability profile is affected only slightly in the CTCFY226A/F228A mutants, similar to the effects of CTCF depletion.

Extended Data Fig. 7 Identification of CES ligands.

a, Plot depicting the log2-transformed fold change in gene expression in relation to the mean of the normalized counts for each gene. Differentially expressed genes (adjusted P value < 0.05, two-tailed Wald test adjusted for multiple testing using the Benjamini–Hochberg procedure) are shown in red. Gene names are included for the 40 genes with the highest fold change. b, Western blot assessing knockdown of CTCF and the cohesin subunit SMC1 upon transfection with a control siRNA targeting luciferase (luc) or siRNAs targeting CTCF or SMC1A. This experiment was performed twice with similar results. c, Colony-formation assay of wild-type and CTCFY226A/F228A cells upon transfection with a control siRNA targeting luciferase or siRNAs targeting CTCF or SMC1A. CTCF remains essential for viability in CTCFY226A/F228A cells. This experiment was performed four times with similar results. d, Peptide array annotation (top left), binding of SA2–SCC1 (top right) or SA2(F371A)–SCC1 mutant (bottom left) and antibody control (bottom right). Three independent experiments were done, with consistent results. One representative example is shown. e, Amino acid sequences of the peptides. Predicted lead-anchoring residues are coloured red.

Extended Data Table 1 Summary of ITC data, and X-ray data collection and refinement statistics
Extended Data Table 2 Quantification of peptide arrays
Extended Data Table 3 Primers and Hi-C statistics

Supplementary information

Supplementary Figure 1

Uncropped images of western blots shown in Extended Data Fig. 3 and Extended Data Fig. 7. The relevant sections used for generating the figures are boxed

Reporting Summary

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Li, Y., Haarhuis, J.H.I., Sedeño Cacciatore, Á. et al. The structural basis for cohesin–CTCF-anchored loops. Nature 578, 472–476 (2020). https://doi.org/10.1038/s41586-019-1910-z

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41586-019-1910-z

This article is cited by

Comments

By submitting a comment you agree to abide by our Terms and Community Guidelines. If you find something abusive or that does not comply with our terms or guidelines please flag it as inappropriate.

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