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:

Alteration of genome folding via contact domain boundary insertion

Abstract

Animal chromosomes are partitioned into contact domains. Pathogenic domain disruptions can result from chromosomal rearrangements or perturbation of architectural factors. However, such broad-scale alterations are insufficient to define the minimal requirements for domain formation. Moreover, to what extent domains can be engineered is just beginning to be explored. In an attempt to create contact domains, we inserted a 2-kb DNA sequence underlying a tissue-invariant domain boundary—containing a CTCF-binding site (CBS) and a transcription start site (TSS)—into 16 ectopic loci across 11 chromosomes, and characterized its architectural impact. Depending on local constraints, this fragment variably formed new domains, partitioned existing ones, altered compartmentalization and initiated contacts reflecting chromatin loop extrusion. Deletions of the CBS or the TSS individually or in combination within inserts revealed its distinct contributions to genome folding. Altogether, short DNA insertions can suffice to shape the spatial genome in a manner influenced by chromatin context.

This is a preview of subscription content, access via your institution

Access options

Rent or buy this article

Prices vary by article type

from$1.95

to$39.95

Prices may be subject to local taxes which are calculated during checkout

Fig. 1: Domain boundary insertions create de novo contact domains.
Fig. 2: Domain boundary insertions can strengthen pre-established boundaries.
Fig. 3: An insertion into a complex genomic region modestly changes short-range interactions, without domain-level impact.
Fig. 4: TSS can influence domain formation by switching its compartment signature.
Fig. 5: TSSs and CTCF cooperatively contribute to new domain formation by driving proximal and distal genome folding, respectively.
Fig. 6: Possible context dependency in how SINE B2 elements shape mouse genome architecture in recent evolution, and a graphic summary of the present study.

Similar content being viewed by others

Data availability

All main, extended data and supplementary figures include publicly available data. All Hi-C, Capture-C, RNA-seq, ChIP–seq, and other applicable next-generation sequencing raw data and processed data generated from the present study are available under accession no. GSE137376 (GEO database). Mouse CTCF ChIP–seq and mouse Hi-C domain boundaries (both asynchronous) shown in Fig. 6a–c are derived from Zhang et al.19 (https://doi.org/10.1038/s41586-019-1778-y), accession no. GSE129997 (GEO database). In Supplementary Fig. 1: Hi-C heatmaps from all cell lines, except for HAP1, are from GEO, accession no. GSE63525 by Rao et al.4 (https://doi.org/10.1016/j.cell.2014.11.021); K562 ChIP–seq data are from ENCODE, CTCF (DCC accession no. ENCSR000AKO), SMC3 (DCC accession no. ENCSR000EGW), RAD21 (DCC accession no. ENCSR000FAD) and Pol2 (DCC accession no. ENCSR000FAY). Source data are provided with this paper.

Code availability

Code used in the present study is available upon request as well as on GitHub (https://github.com/dizhmp/boundary-insertion).

References

  1. 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 

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

    Article  CAS  PubMed  PubMed Central  Google Scholar 

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

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  4. 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 

  5. Phillips-Cremins, J. E. et al. Architectural protein subclasses shape 3D organization of genomes during lineage commitment. Cell 153, 1281–1295 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  6. Schwarzer, W. et al. Two independent modes of chromatin organization revealed by cohesin removal. Nature 551, 51–56 (2017).

    Article  PubMed  PubMed Central  Google Scholar 

  7. Rao, S. S. P. et al. Cohesin loss eliminates all loop domains. Cell 171, 305–320.e24 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  8. Rowley, M. J. et al. Evolutionarily conserved principles predict 3D chromatin organization. Mol. Cell 67, 837–852.e7 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  9. Nora, E. P. et al. Targeted degradation of CTCF decouples local insulation of chromosome domains from genomic compartmentalization. Cell 169, 930–944.e22 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  10. Hug, C. B., Grimaldi, A. G., Kruse, K. & Vaquerizas, J. M. Chromatin architecture emerges during zygotic genome activation independent of transcription. Cell 169, 216–228.e19 (2017).

    Article  CAS  PubMed  Google Scholar 

  11. Franke, M. et al. Formation of new chromatin domains determines pathogenicity of genomic duplications. Nature 538, 265–269 (2016).

    Article  CAS  PubMed  Google Scholar 

  12. 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 

  13. Fudenberg, G. & Pollard, K. S. Chromatin features constrain structural variation across evolutionary timescales. Proc. Natl Acad. Sci. USA 116, 2175–2180 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  14. Symmons, O. et al. The shh topological domain facilitates the action of remote enhancers by reducing the effects of genomic distances. Dev. Cell 39, 529–543 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

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

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  16. Narendra, V. et al. CTCF establishes discrete functional chromatin domains at the Hox clusters during differentiation. Science 347, 1017–1021 (2015).

    CAS  PubMed  PubMed Central  Google Scholar 

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

    Article  CAS  PubMed  Google Scholar 

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

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  19. Zhang, Y. et al. Transcriptionally active HERV-H retrotransposons demarcate topologically associating domains in human pluripotent stem cells. Nat. Genet. 51, 1380–1388 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  20. Barutcu, A. R., Maass, P. G., Lewandowski, J. P., Weiner, C. L. & Rinn, J. L. A TAD boundary is preserved upon deletion of the CTCF-rich Firre locus. Nat. Commun. 9, 1444 (2018).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  21. Mátés, L. et al. Molecular evolution of a novel hyperactive Sleeping Beauty transposase enables robust stable gene transfer in vertebrates. Nat. Genet. 41, 753–761 (2009).

    Article  CAS  PubMed  Google Scholar 

  22. Carette, J. E. et al. Ebola virus entry requires the cholesterol transporter Niemann–Pick C1. Nature 477, 340–343 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  23. Haarhuis, J. H. I. et al. The cohesin release factor WAPL restricts chromatin loop extension. Cell 169, 693–707.e14 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  24. Van Bortle, K. et al. Insulator function and topological domain border strength scale with architectural protein occupancy. Genome Biol. 15, R82 (2014).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  25. Mayer, A. et al. Native elongating transcript sequencing reveals human transcriptional activity at nucleotide resolution. Cell 161, 541–554 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  26. Vian, L. et al. The energetics and physiological impact of cohesin extrusion. Cell 173, 1165–1178.e20 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  27. Redolfi, J. et al. DamC reveals principles of chromatin folding in vivo without crosslinking and ligation. Nat. Struct. Mol. Biol. 26, 471–480 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  28. 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, 6456 (2015).

    Google Scholar 

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

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  30. Dixon, J. R. et al. Chromatin architecture reorganization during stem cell differentiation. Nature 518, 331–336 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  31. Krijger, P. H. L. et al. Cell-of-origin-specific 3D genome structure acquired during somatic cell reprogramming. Cell Stem Cell 18, 597–610 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  32. Ke, Y. et al. 3D chromatin structures of mature gametes and structural reprogramming during mammalian embryogenesis. Cell 170, 367–381.e20 (2017).

    Article  CAS  PubMed  Google Scholar 

  33. Du, Z. et al. Allelic reprogramming of 3D chromatin architecture during early mammalian development. Nature 547, 232–235 (2017).

    Article  CAS  PubMed  Google Scholar 

  34. Heinz, S. et al. Transcription elongation can affect genome 3D structure. Cell 174, 1522–1536.e22 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  35. Gong, Y. et al. Stratification of TAD boundaries reveals preferential insulation of super-enhancers by strong boundaries. Nat. Commun. 9, 542 (2018).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

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

    Article  CAS  PubMed  Google Scholar 

  37. Nuebler, J., Fudenberg, G., Imakaev, M., Abdennur, N. & Mirny, L. A. Chromatin organization by an interplay of loop extrusion and compartmental segregation. Proc. Natl Acad. Sci. USA 115, E6697–E6706 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  38. Sun, L. et al. Mixed lineage kinase domain-like protein mediates necrosis signaling downstream of RIP3 kinase. Cell 148, 213–227 (2012).

    Article  CAS  PubMed  Google Scholar 

  39. Zhao, J. et al. Mixed lineage kinase domain-like is a key receptor interacting protein 3 downstream component of TNF-induced necrosis. Proc. Natl Acad. Sci. USA 109, 5322–5327 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  40. Galluzzi, L., Buqué, A., Kepp, O., Zitvogel, L. & Kroemer, G. Immunogenic cell death in cancer and infectious disease. Nat. Rev. Immunol. 17, 97–111 (2017).

    Article  CAS  PubMed  Google Scholar 

  41. Shan, B., Pan, H., Najafov, A. & Yuan, J. Necroptosis in development and diseases. Genes Dev. 32, 327–340 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  42. Yuan, J., Amin, P. & Ofengeim, D. Necroptosis and RIPK1-mediated neuroinflammation in CNS diseases. Nat. Rev. Neurosci. 20, 19–33 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  43. Chung, C. C. et al. Meta-analysis identifies four new loci associated with testicular germ cell tumor. Nat. Genet. 45, 680–685 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  44. Astle, W. J. et al. The allelic landscape of human blood cell trait variation and links to common complex disease. Cell 167, 1415–1429.e19 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  45. Mitchell, J. S. et al. Genome-wide association study identifies multiple susceptibility loci for multiple myeloma. Nat. Commun. 7, 12050 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  46. Hou, C., Zhao, H., Tanimoto, K. & Dean, A. CTCF-dependent enhancer-blocking by alternative chromatin loop formation. Proc. Natl Acad. Sci. USA 105, 20398–20403 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  47. Rawat, P., Jalan, M., Sadhu, A., Kanaujia, A. & Srivastava, M. Chromatin domain organization of the TCRb locus and its perturbation by ectopic CTCF binding. Mol. Cell Biol. 37, e00557–16 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  48. Cong, L. et al. Multiplex genome engineering using CRISPR/Cas systems. Science 339, 819–823 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  49. Mali, P. et al. RNA-guided human genome engineering via Cas9. Science 339, 823–826 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  50. Busslinger, G. A. et al. Cohesin is positioned in mammalian genomes by transcription, CTCF and Wapl. Nature 544, 503–507 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  51. Despang, A. et al. Functional dissection of the Sox9—Kcnj2 locus identifies nonessential and instructive roles of TAD architecture. Nat. Genet 51, 1263–1271 (2019).

    Article  CAS  PubMed  Google Scholar 

  52. Choudhary, M. N. et al. Co-opted transposons help perpetuate conserved higher-order chromosomal structures. Genome Biol. 21, 16 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  53. Karijolich, J., Zhao, Y., Alla, R. & Glaunsinger, B. Genome-wide mapping of infection-induced SINE RNAs reveals a role in selective mRNA export. Nucleic Acids Res. 45, 6194–6208 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  54. Zhang, H. et al. Chromatin structure dynamics during the mitosis-to-G1 phase transition. Nature 576, 158–162 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  55. Sundaram, V. et al. Widespread contribution of transposable elements to the innovation of gene regulatory networks. Genome Res. 24, 1963–1976 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  56. Schmidt, D. et al. Waves of retrotransposon expansion remodel genome organization and CTCF binding in multiple mammalian lineages. Cell 148, 335–348 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  57. Bourque, G. et al. Evolution of the mammalian transcription factor binding repertoire via transposable elements. Genome Res. 18, 1752–1762 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  58. Thybert, D. et al. Repeat associated mechanisms of genome evolution and function revealed by the Mus caroli and Mus pahari genomes. Genome Res. 28, 448–459 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

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

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  60. Zhang, Y. et al. Chromatin connectivity maps reveal dynamic promoter-enhancer long-range associations. Nature 504, 306–310 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  61. Kentepozidou, E. et al. Clustered CTCF binding is an evolutionary mechanism to maintain topologically associating domains. Genome Biol. 21, 5 (2020).

    Article  PubMed  PubMed Central  Google Scholar 

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

    Article  CAS  PubMed  Google Scholar 

  63. Zhan, Y. et al. Reciprocal insulation analysis of Hi-C data shows that TADs represent a functionally but not structurally privileged scale in the hierarchical folding of chromosomes. Genome Res. 27, 479–490 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  64. Hsieh, T. S. et al. Resolving the 3D landscape of transcription-linked mammalian chromatin folding. Mol. Cell 78, 539–553.e8 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  65. Krietenstein, N. et al. Ultrastructural details of mammalian chromosome architecture. Mol. Cell 78, 554–565.e7 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  66. Kurita, R. et al. Establishment of immortalized human erythroid progenitor cell lines able to produce enucleated red blood cells. PLoS ONE 8, e59890 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  67. Zayed, H., Izsvák, Z., Walisko, O. & Ivics, Z. Development of hyperactive sleeping beauty transposon vectors by mutational analysis. Mol. Ther. 9, 292–304 (2004).

    Article  CAS  PubMed  Google Scholar 

  68. Huang, P. et al. Comparative analysis of three-dimensional chromosomal architecture identifies a novel fetal hemoglobin regulatory element. Genes Dev. 31, 1704–1713 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  69. Davies, J. O. J. et al. Multiplexed analysis of chromosome conformation at vastly improved sensitivity. Nat. Methods 13, 74–80 (2016).

    Article  CAS  PubMed  Google Scholar 

  70. Hsiung, C. C.- et al. A hyperactive transcriptional state marks genome reactivation at the mitosis-G1 transition. Genes Dev. 30, 1423–1439 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  71. Hsiau, T. et al. Inference of CRISPR edits from Sanger trace data. Preprint at bioRxiv https://doi.org/10.1101/251082 (2019).

  72. Kim, S., Kim, D., Cho, S. W., Kim, J. & Kim, J. Highly efficient RNA-guided genome editing in human cells via delivery of purified Cas9 ribonucleoproteins. Genome Res. 24, 1012–1019 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

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

    CAS  PubMed  PubMed Central  Google Scholar 

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

    Article  CAS  Google Scholar 

  75. Sloan, C. A. et al. ENCODE data at the ENCODE portal. Nucleic Acids Res. 44, 726–732 (2016).

    Article  CAS  Google Scholar 

  76. Kerpedjiev, P. et al. HiGlass: web-based visual exploration and analysis of genome interaction maps. Genome Biol. 19, 125 (2018).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  77. Forcato, M. et al. Comparison of computational methods for Hi-C data analysis. Nat. Methods 14, 679–685 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

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

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  79. Filippova, D., Patro, R., Duggal, G. & Kingsford, C. Identification of alternative topological domains in chromatin. Algorithms Mol. Biol. 9, 14 (2014).

    Article  PubMed  PubMed Central  Google Scholar 

  80. Eisenberg, E. & Levanon, E. Y. Human housekeeping genes, revisited. Trends Genet. 29, 569–574 (2013).

    Article  CAS  PubMed  Google Scholar 

  81. 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 

  82. Li, H. et al. The sequence Alignment/Map format and SAMtools. Bioinformatics 25, 2078–2079 (2009).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  83. 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 

  84. 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 

  85. Gilgenast, T. G. & Phillips-Cremins, J. E. Systematic evaluation of statistical methods for identifying looping interactions in 5C data. Cell Syst. 8, 197–211.e13 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  86. Hunter, J. D. Matplotlib: a 2D graphics environment. Comput. Sci. Eng. 9, 90–95 (2007).

    Article  Google Scholar 

  87. Ambrosini, G., Groux, R. & Bucher, P. PWMScan: a fast tool for scanning entire genomes with a position-specific weight matrix. Bioinformatics 34, 2483–2484 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  88. Khan, A. et al. JASPAR 2018: update of the open-access database of transcription factor binding profiles and its web framework. Nucleic Acids Res. 46, D260–D266 (2018).

    Article  CAS  PubMed  Google Scholar 

  89. 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 

  90. Martin, M. Cutadapt removes adapter sequences from high-throughput sequencing reads. EMBnet. J. 17, 10–12 (2011).

    Article  Google Scholar 

  91. Magoč, T. & Salzberg, S. L. FLASH: fast length adjustment of short reads to improve genome assemblies. Bioinformatics 27, 2957–2963 (2011).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  92. Langmead, B. Aligning short sequencing reads with Bowtie. Curr. Protoc. Bioinform. Chapter 11, Unit 11.7 (2010).

    Google Scholar 

  93. Zhang, Y. et al. Model-based analysis of ChIP-Seq (MACS). Genome Biol. 9, R137 (2008).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  94. Xu, S., Grullon, S., Ge, K. & Peng, W. Spatial clustering for identification of ChIP-enriched regions (SICER) to map regions of histone methylation patterns in embryonic stem cells. Methods Mol. Biol. 1150, 97–111 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  95. Ramírez, F. et al. deepTools2: a next generation web server for deep-sequencing data analysis. Nucleic Acids Res. 44, W160–W165 (2016).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  96. Ross-Innes, C. S. et al. Differential oestrogen receptor binding is associated with clinical outcome in breast cancer. Nature 481, 389–393 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  97. Patro, R., Duggal, G., Love, M. I., Irizarry, R. A. & Kingsford, C. Salmon: fast and bias-aware quantification of transcript expression using dual-phase inference. Nat. Methods 14, 417–419 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  98. Soneson, C., Love, M. I. & Robinson, M. D. Differential analyses for RNA-seq: transcript-level estimates improve gene-level inferences. F1000Research 4, 1521 (2015).

    Article  PubMed  CAS  Google Scholar 

  99. 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 

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

    Article  CAS  PubMed  Google Scholar 

  101. Weiss, M. J., Yu, C. & Orkin, S. H. Erythroid-cell-specific properties of transcription factor GATA-1 revealed by phenotypic rescue of a gene-targeted cell line. Mol. Cell. Biol. 17, 1642–1651 (1997).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  102. Norton, H. K. et al. Detecting hierarchical genome folding with network modularity. Nat. Methods 15, 119–122 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

Download references

Acknowledgements

We thank B. van Steensel (Netherlands Cancer Institute) for providing HAP1 cells; Z. Izsvák (Max Delbrück Center) and Z. Ivics (The Paul Ehrlich Institute) for providing the Sleeping Beauty transposon constructs; A. Raj, O. Symmons and F. Yue for helpful comments on the manuscript. We thank the Flow Cytometry Core at the Children’s Hospital of Philadelphia; J. Yano and P. Evans for assistance; and members of the Blobel laboratory for helpful discussions. This work was supported by grants (nos. R01DK054937 and U01HL129998A to G.A.B and R24DK106766 to R.C.H. and G.A.B.). This work was also supported by the Spatial and Functional Genomics program at the Children’s Hospital of Philadelphia.

Author information

Authors and Affiliations

Authors

Contributions

D.Z. and G.A.B. conceived the study and designed the experiments. D.Z. performed a large majority of the experiments, analyzed all datasets and interpreted the results. P.H. conducted Hi-C and Capture-C for half the replicates for transposon-edited and control cell lines, and helped with Hi-C and Capture-C analysis and interpretation. M.S. helped generate and characterize cell lines derived from CRISPR targeting the TSSs and the 2-kb elements. C.A.K., B.G. and R.C.H. prepared ChIP–seq and RNA-seq libraries, performed all sequencing, uploaded sequencing data, and conducted RNA-seq alignment and ChIP–seq peak calling. H.Z. generated mouse ChIP and Hi-C datasets used for recent mouse genome evolution analysis. T.G.G. and J.E.P.-C. helped with Hi-C data visualization and interpretation. D.Z. and G.A.B wrote the paper with input from all authors.

Corresponding authors

Correspondence to Di Zhang or Gerd A. Blobel.

Ethics declarations

Competing interests

The authors declare no competing interests.

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 Generation and characterization of transposon genome-edited clones with multiple insertions.

a, Estimated insertion copy numbers using qPCR (see Methods) after transposon insertion in pooled cells and in single-cell-derived clones (numbered). N = 1 qPCR measurement. b, Insertion site mapping: fragmented gDNA containing insertions are captured by biotinylated oligos capturing the inversed repeats (green rectangles), which flank the 2 kb element (orange rectangles). Junction reads are mapped to identify insertion sites. c, Junction read coverage for Clone 21: horizontal axis denotes genomic coordinates (single nucleotide resolution) with > 25X coverage; vertical axis shows read coverage. The spike in the middle of each peak consists of two neighboring nucleotides between which an insertion is located. Data from N = 1 experiment. d, The locations and orientations of Clone 21 insertion sites. The CBS and TSS are in cis (Fig. 1a). “” denotes that the CBS is on the plus strand and that the TSS transcribes from left to right, and vice versa for “”. Each insertion site orientation was confirmed in (g). e, Junction read coverage for Clone 25, similar to (c). Data is from N = 1 experiment. f, The locations and orientations of Clone 25 insertion sites, similar to (d). g, Insertion-driven transcription in both directions/strands measured by quantitative PCR with reverse transcription (RT-qPCR). Transcript levels were normalized relative to the geometric mean of the Ct values of 11 housekeeping genes. N = 2 independent experiments for each genotype.

Extended Data Fig. 2 Insertion-driven new domains: detailed comparisons (an extension to Fig. 1).

Throughout, red arrow: insertion site; green arrow: upstream or downstream CBSs; blue/purple arrow: nearby boundaries; orange arrowhead in the browser tracks: site and orientation of the insertion. Green lines demarcate new domains. Yellow/green rectangles (squares) indicate regions with overall depleted (enriched) contacts upon insertion. (a, b): related to Fig. 1b, c. a, An extension to Fig. 1b showing Hi-C maps for both no-insertion controls (left and middle) and the insertion clone (right) at C21S4, each accompanied by corresponding data tracks. b, Log2 fold changes in interaction frequencies between two no-insertion controls (left), and between the insertion clone and no-insertion controls (middle and right) for the region in (a). Yellow/green rectangles: depleted interactions upon insertion; yellow/green squares: increased interactions between two B-compartment domains partitioned by the new domain with A compartment signature. (c, d): related to Fig. 1d, e. c, An extension to Fig. 1d showing both no-insertion controls at C21S2. d, Log2 fold changes in interaction frequencies between no-insertion controls and between insertion and no-insertion controls for the region in (c). (e, f): related to Fig. 1f, g. e, An extension to Fig. 1f showing both no-insertion controls at C21S5. f, Log2 fold changes between no-insertion controls and between insertion and no-insertion controls for the region in (e). Each Hi-C heatmap presents merged data from 2 independent experiments for each genotype. 2 CTCF & RAD21 ChIP-seq and 2 RNA-seq experiments were performed for each genotype, with 1 of each displayed.

Extended Data Fig. 3 Additional insertion loci with possible domain-level changes.

Throughout, red arrow: insertion site; green or blue arrow: nearby boundaries; orange/blue arrowhead in the browser tracks: site and orientation of insertion. Green lines demarcate (possible) new domains. Yellow/green rectangles indicate regions with overall depleted contacts upon insertion. a, De novo domain upon insertion at C21S1: Hi-C maps for both no-insertion controls (left and middle) and the insertion clone (right) at C21S1, each accompanied by corresponding data tracks. b, Insulation scores for the region in (a). c, Log2 fold changes in interaction frequencies between the two no-insertion controls (left) and between the insertion clone and no-insertion controls (middle and right) for the region in (a). d, A small subtle domain forms upon insertion at C25S3 locus. e, Insulation scores for the region in (d). f, Log2 fold changes in interaction frequencies for the region in (d). g, Modest strengthening of an existing boundary upon insertion at C25S4. h, Insulation scores for the region in (g). i, Log2 fold changes for the region in (g). j, Subtle strengthening of an existing boundary upon insertion at C25S1. The black arrowheads point at insertion-associated changes. k, Insulation scores for the region in (j). l, Log2 fold changes for the region in (j). Each Hi-C heatmap presents merged data from 2 independent experiments for each genotype. 2 CTCF & RAD21 ChIP-seq and 2 RNA-seq experiments were performed for each genotype, with 1 of each displayed.

Extended Data Fig. 4 An ectopic insertion can redirect its local chromatin from B to A compartment.

Throughout, left: compartment eigenvectors (cyan denotes B compartment; red denotes A compartment) for the ~14 Mb region marked by the green rectangle on the chromosome diagram; middle: Hi-C heatmaps for this ~14 Mb region surrounding C21S4; right: distal interactions between this ~14 Mb region and a ~40 Mb region downstream marked by the purple rectangle. Black arrows: compartment switch; orange arrowhead: location of the insertion; black arrowhead: corresponding locations in no-insertion controls. a, No-insertion control 1 (WT) at C21S4. b, No-insertion control 2 (Clone 25) at C21S4. c, Insertion clone (Clone 21) at C21S4: compartment eigenvectors demonstrate the insertion locus trending from a strong B compartment towards A as the largest change in the region. The Hi-C heatmap for the ~14 Mb with the insertion at the center shows a plaid like pattern, with gained interactions between the insertion locus and its nearby A compartment regions. Distal interactions (right) shows the insertion locus forming distal interactions with other A-compartment regions (black arrows), which are absent in (a, b). Each Hi-C result depicts merged data from 2 independent Hi-C experiments for each genotype.

Extended Data Fig. 5 Boundary-associated DNA insertions can strengthen pre-established boundaries: additional controls (an extension to Fig. 2).

Throughout, red arrow: insertion site; green or blue arrow: nearby boundaries; Blue/orange arrowhead in the browser tracks: site and orientation of the insertion. Yellow/green rectangles indicate regions with overall depleted contacts upon insertion. (a, b) are related to Fig. 2a–c. a, An extension to Fig. 2a showing both no-insertion controls (left and middle) and the insertion clone (right) at C25S5, each accompanied by corresponding data tracks. b, An extension to Fig. 2c: log2 fold changes in interaction frequencies between two no-insertion controls (left) and between the insertion clone and no-insertion controls (middle and right) for the region in (a). (c, d) are related to Fig. 2d–f. c, An extension to Fig. 2d showing both no-insertion controls at C21S7. d, An extension to Fig. 2f: log2 fold changes in interaction frequencies between two no-insertion controls and between the insertion clone and no-insertion controls for the region shown in (d). Each Hi-C heatmap represents merged data from 2 independent experiments for each genotype. 2 CTCF & RAD21 ChIP-seq and 2 RNA-seq experiments were conducted for each genotype, with 1 of each exhibited.

Extended Data Fig. 6 Insertion loci without apparent detectable domain-level changes.

Throughout, red arrow: insertion site; orange/blue arrowhead in the browser tracks: locus/orientation of the insertion. a, An insertion at C21S6: Hi-C maps for both no-insertion controls (left and middle) and the insertion clone (right) at C21S6, each accompanied by corresponding data tracks. b, Insulation scores for the region in (a). c, Log2 fold changes in interaction frequencies between two no-insertion controls (left) and between the insertion clone and no-insertion controls (middle and right) for the region in (a). d, Hi-C contact maps at C21S10. e, Insulation score profiles for the region in (d). f, Log2 fold changes in interaction frequencies between two no-insertion controls and between the insertion clone and no-insertion controls for the region in (d). g, Hi-C contact maps at C25S6. h, Insulation score profiles for the region in (g). i, Log2 fold changes in interaction frequencies for the region shown in (g). Each Hi-C heatmap presents merged data from 2 independent experiments performed for each genotype. 2 CTCF & RAD21 ChIP-seq and 2 RNA-seq experiments were performed for each genotype, with 1 of each displayed.

Extended Data Fig. 7 Transcription of insertion-proximal genes remains mostly stable, with MLKL as an exception.

a, An MA plot showing Clone 21 vs. non-Clone 21 transcriptomes. Each dot: a gene; red dots: differentially expressed (DE) genes at an FDR < 0.01; color-coded circles: insertion-proximal genes by distance ranges; red line: no-change line; two orange lines: +/− 1 log2 fold change. b, Clone 25 vs. non-Clone 25 transcriptomes. c, Clone 21 has ~95 DE genes transcriptome-wide (related to (a)). d, Clone 25 has ~160 DE genes transcriptome-wide (related to (b)). e, DE status of all insertion-proximal genes. The DE gene between 50 kb and 500 kb to an insertion, MLKL, is characterized in (f) and (h). In (ae), 2 RNA-seq experiments were performed for each genotype. DE analysis was conducted with Clone 21 vs. non-Clone 21 (WT and Clone 25) and Clone 25 vs. non-Clone 25 (WT and Clone 21). f, RT-qPCR of MLKL and GLG1/RFWD3, two genes flanking the insertion (see (h)). N = 2 independent experiments for each genotype. g, GWAS significant variants near GLG1/RFWD3/MLKL insertion locus43,44,45. h, GLG1/RFWD3/MLKL locus (blue arrowhead: location/orientation of the insertion) using ChIP-seq/RNA-seq/Capture-C. The insertion coincides with reduced RAD21 binding at a peak immediately downstream. The insertion contacts the promoter of GLG1 (Capture-C: Probing the insert). MLKL promoter also interacts with GLG1 promoter (Capture-C: Probing MLKL promoter), albeit no apparent changes in interactions of MLKL promoter upon insertion. Capture-C presents merged data from 2 independent experiments for each genotype. 2 CTCF & RAD21 ChIP-seq, 1 H3K27ac ChIP-seq and 2 RNA-seq experiments were conducted for each genotype, with 1 of each shown.

Source data

Extended Data Fig. 8 CRISPR dissections of insertion, and CTCF/RAD21 at C21S4.

a, Left: sgRNAs within the insertion element (red lines: Pol2/CTCF peak centers). Right: TSS_sgRNA_2&4 and TSS_sgRNA_3&4 reduce transcription more effectively at C21S4. N = 1. b, CRISPR deletion of the inserted CBS spares transcription. c, Clone 21 ΔTSS: TSS_sgRNA_2&4-edited Clone 21 abrogates transcription, with the CBS intact. d, e, Clone 21 ΔCTCF/ΔTSS #1&#2: Clone 21 with its CBS already disrupted (b) further edited with TSS_sgRNA_2&4 and TSS_sgRNA_3&4, respectively. In (be), N = 2 experiments for each genotype. In (f, g and i), red arrow: insertion site; green arrow: downstream CBSs; blue/purple arrow: strong boundary nearby; orange arrowhead: insertion location/orientation. f, Hi-C of Clone 21 ΔCTCF/ΔTSS #1 (d) at C21S4: a ~27 Mb heterozygous deletion (h) influences heatmap interpretation. g, Hi-C of Clone 21 ΔCTCF/ΔTSS #2 (e) at C21S4: domain configuration restored close to pre-insertion level (Fig. 4a). h, Virtual 4 C (black arrow: viewpoint; red star: C21S4; GRIK2: C21S5): Clone 21 ΔCTCF/ΔTSS #1 has both short-range contacts and strong >25-Mb distal contacts, suggesting a heterozygous deletion between C21S4 and C21S5 (grey bars: chromosomes; dashed line: deletion). i, ΔCBS/ΔTSS restores nearby chromatin folding pattern to pre-insertion levels. Differentially bound CTCF (C2, C4) and RAD21 peaks (R1-R5) upon insertion highlighted. Directionality Index of Clone 21 ΔCTCF/ΔTSS #1 Capture-C: Fig. 4j. In (fi), each Hi-C/Capture-C describes merged data from at least 2 independent experiments for each genotype. 2 CTCF/RAD21 ChIP-seq and 1 H3K27ac ChIP-seq for each genotype, with 1 of each shown. j, Pairwise comparisons between genotypes of CTCF binding (C2 and C4: (i) and Fig. 4f–i). k, Pairwise comparisons between genotypes of RAD21 binding (R1-R5: (i) and Fig. 4f–i). In (j, k), non-Clone 21: 3 genotypes without Clone21 insertions, each with 2 ChIP-seq replicates. Clone21 CTCF/TSS and derived CRISPR clones: 1 genotype, each with 2 ChIP-seq replicates. P-values (not adjusted for multiple comparisons): from a two-sided Wald test.

Source data

Extended Data Fig. 9 CRISPR dissections of insertion, and RAD21 distribution at C21S2.

a, TSS_sgRNA_2&4 and TSS_sgRNA_3&4 (as in Extended Data Fig. 8a) reduce transcription more effectively at C21S2 in CRISPR-Cas9 RNP-transfected cells. N = 1 experiment. b, Deletion of the inserted CBS reduces but does not abolish transcription at C21S2. c, Clone 21 ΔTSS derived from TSS_sgRNA_2&4-edited Clone 21 abrogates transcription, with the CBS intact. d, Clone 21 ΔCTCF/ΔTSS #1: derived from CBS-disrupted Clone 21 (b) further edited with TSS_sgRNA_2&4. e, Clone 21 ΔCTCF/ΔTSS #2: derived from CBS-disrupted Clone 21 (b) further edited with TSS_sgRNA_3&4. In (be), N = 2 independent experiments for each genotype. In (fh), red arrow: insertion site; green or blue arrow: downstream CBSs; orange arrowhead in the browser tracks: locus/orientation of the insertion. f, g, Hi-C maps of Clone 21 ΔCTCF/ΔTSS #1 (d) and of Clone 21 ΔCTCF/ΔTSS #2 (e), respectively, at C21S2: deletions of both the CBS and the TSS restore the domain configuration close to pre-insertion level (Fig. 5a). h, Capture-C and corresponding data tracks showing that ΔCTCF/ΔTSS rescues local chromatin contact pattern close to that of WT. Differentially bound RAD21 peaks (R6, R7) upon CBS-TSS insertion highlighted. Directionality Index of Clone 21 ΔCTCF/ΔTSS #1 Capture-C: Fig. 5j. In (fh), each Hi-C/Capture-C depicts merged data from at least 2 independent experiments for each genotype. 2 CTCF/RAD21 ChIP-seq and 1 H3K27ac ChIP-seq for each genotype, with 1 of each shown. i, Pairwise comparisons between genotypes of RAD21 binding at two RAD21 peaks (R6 and R7, as in (h) and Fig. 5f–i). Non-Clone 21: 3 genotypes without Clone 21 insertions, each with 2 ChIP-seq replicates. All others: 1 genotype, each with 2 ChIP-seq replicates. P-values (not adjusted for multiple comparisons) are derived from a two-sided Wald test through DiffBind.

Source data

Extended Data Fig. 10 Deletion of the endogenous 2 kb element leads to a boundary shift, while local domain organization is stable.

a, Hi-C of no-deletion control showing the endogenous boundary where the 2 kb element (blue arrowhead) is derived, accompanied by corresponding data tracks. b, Deletion of the 2 kb (crossed-out blue arrowhead) leaves the overall domain configuration largely intact. The highlighted ~400 kb region is further examined in (c) and (f). c, Insulation scores show overall concordance, with a possible shift in boundary by ~60 kb to the left upon deletion. d, Genotyping confirms the desired deletion between sgRNAs flanking the 2 kb. e, ChIP-seq further verifies the deletion, as reflected in lack of signal (black arrows) within the 2 kb element (highlighted). f, Upon 2 kb deletion (highlighted in red), the point of local maximal insulation shifts ~60 kb to the left (c), coinciding with the distance between the TSSs of PARL and its nearest transcribed gene: MAP6D1 (highlighted in yellow). This shift (red line) also corresponds to the distance between the deleted CBS and its nearest CTCF peak to the left, which now has reduced CTCF/RAD21 binding. Each Hi-C result presents merged data from 2 independent experiments for each genotype. 2 CTCF & RAD21 ChIP-seq experiments for each genotype, with 1 of each shown.

Source data

Supplementary information

Supplementary Information

Supplementary Figs. 1 and 2

Reporting Summary

Supplementary Tables

Supplementary Tables 1–4

Supplementary Data 1

Sequence map of the Sleeping Beauty transposon with the 2-kb insert.

Source data

Source Data Fig. 6

Statistical source data for Fig. 6c.

Source Data Extended Data Fig. 7

Statistical source data for Extended Data Fig. 7e.

Source Data Extended Data Fig. 8

Statistical source data for Extended Data Fig. 8j and k.

Source Data Extended Data Fig. 9

Statistical source data for Extended Data 9i.

Source Data Extended Data Fig. 10

Uncropped gel for Extended Data Fig. 10d.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Zhang, D., Huang, P., Sharma, M. et al. Alteration of genome folding via contact domain boundary insertion. Nat Genet 52, 1076–1087 (2020). https://doi.org/10.1038/s41588-020-0680-8

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41588-020-0680-8

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