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:

Single-gene imaging links genome topology, promoter–enhancer communication and transcription control

Abstract

Transcription activation by distal enhancers is essential for cell-fate specification and maintenance of cellular identities. How long-range gene regulation is physically achieved, especially within complex regulatory landscapes of non-binary enhancer–promoter configurations, remains elusive. Recent nanoscopy advances have quantitatively linked promoter kinetics and ~100- to 200-nm-sized clusters of enhancer-associated regulatory factors (RFs) at important developmental genes. Here, we further dissect mechanisms of RF clustering and transcription activation in mouse embryonic stem cells. RF recruitment into clusters involves specific molecular recognition of cognate DNA and chromatin-binding sites, suggesting underlying cis-element clustering. Strikingly, imaging of tagged genomic loci, with ≤1 kilobase and ~20-nanometer precision, in live cells, reveals distal enhancer clusters over the extended locus in frequent close proximity to target genes—within RF-clustering distances. These high-interaction-frequency enhancer-cluster ‘superclusters’ create nano-environments wherein clustered RFs activate target genes, providing a structural framework for relating genome organization, focal RF accumulation and transcription activation.

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: Specific interactions with acetylated targets and Sox DNA motifs are needed for Brd4 and Sox2 clustering, respectively, while IDRs are dispensable.
Fig. 2: Specific interactions with acetylated targets and Sox DNA motifs are needed for Brd4 and Sox2 clustering, respectively, at the Pou5f1 locus, while IDRs are dispensable.
Fig. 3: High detection sensitivity enables imaging of tagged genomic loci and measuring of genomic interactions with nanometer precision in live cells.
Fig. 4: Imaging genomic interactions in the extended Pou5f1 locus reveals frequent proximity of distal enhancer clusters and the target gene.
Fig. 5: A shared transcription nanoscale environment created by RF clustering underlies apparent transcription coordination between Nanog sister chromatids.

Similar content being viewed by others

Data availability

Datasets that support the findings in the paper are available in the Zenodo repository, https://doi.org/10.5281/zenodo.3960997. Source data are provided with this paper.

Code availability

Custom-written analysis code is available in the Zenodo repository, https://doi.org/10.5281/zenodo.3960997.

References

  1. Levine, M. Transcriptional enhancers in animal development and evolution. Curr. Biol. 20, R754–R763 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

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

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  3. Ptashne, M. Gene regulation by proteins acting nearby and at a distance. Nature 322, 697–701 (1986).

    Article  CAS  PubMed  Google Scholar 

  4. Blackwood, E. M. & Kadonaga, J. T. Going the distance: a current view of enhancer action. Science 281, 60–63 (1998).

    Article  CAS  PubMed  Google Scholar 

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

    Article  CAS  PubMed  Google Scholar 

  6. 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 

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

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  8. 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 

  9. Sanyal, A., Lajoie, B. R., Jain, G. & Dekker, J. The long-range interaction landscape of gene promoters. Nature 489, 109–113 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  10. Bartman, C. R., Hsu, S. C., Hsiung, C. C., Raj, A. & Blobel, G. A. Enhancer regulation of transcriptional bursting parameters revealed by forced chromatin looping. Mol. Cell 62, 237–247 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  11. Deng, W. et al. Controlling long-range genomic interactions at a native locus by targeted tethering of a looping factor. Cell 149, 1233–1244 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

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

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  13. Morgan, S. L. et al. Manipulation of nuclear architecture through CRISPR-mediated chromosomal looping. Nat. Commun. 8, 15993 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  14. Kim, J. H. et al. LADL: light-activated dynamic looping for endogenous gene expression control. Nat. Methods 16, 633–639 (2019).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

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

    Article  PubMed  PubMed Central  Google Scholar 

  16. Heist, T., Fukaya, T. & Levine, M. Large distances separate coregulated genes in living Drosophila embryos. Proc. Natl Acad. Sci. USA 116, 15062–15067 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

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

    Article  CAS  PubMed  PubMed Central  Google Scholar 

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

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  19. Kim, S. & Shendure, J. Mechanisms of interplay between transcription factors and the 3D genome. Mol. Cell 76, 306–319 (2019).

    Article  CAS  PubMed  Google Scholar 

  20. Misteli, T. Beyond the sequence: cellular organization of genome function. Cell 128, 787–800 (2007).

    Article  CAS  PubMed  Google Scholar 

  21. 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 

  22. 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 

  23. 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 

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

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  25. de Laat, W. & Duboule, D. Topology of mammalian developmental enhancers and their regulatory landscapes. Nature 502, 499–506 (2013).

    Article  PubMed  CAS  Google Scholar 

  26. Amandio, A. R., Lopez-Delisle, L., Bolt, C. C., Mascrez, B. & Duboule, D. A complex regulatory landscape involved in the development of mammalian external genitals. Elife 9, e52962 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  27. Whyte, W. A. et al. Master transcription factors and mediator establish super-enhancers at key cell identity genes. Cell 153, 307–319 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

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

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  29. Lim, B., Heist, T., Levine, M. & Fukaya, T. Visualization of transvection in living Drosophila embryos. Mol. Cell 70, 287–296 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  30. Tan, L., Xing, D., Daley, N. & Xie, X. S. Three-dimensional genome structures of single sensory neurons in mouse visual and olfactory systems. Nat. Struct. Mol. Biol. 26, 297–307 (2019).

    Article  CAS  PubMed  Google Scholar 

  31. Lomvardas, S. et al. Interchromosomal interactions and olfactory receptor choice. Cell 126, 403–413 (2006).

    Article  CAS  PubMed  Google Scholar 

  32. Markenscoff-Papadimitriou, E. et al. Enhancer interaction networks as a means for singular olfactory receptor expression. Cell 159, 543–557 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  33. Iborra, F. J., Pombo, A., Jackson, D. A. & Cook, P. R. Active RNA polymerases are localized within discrete transcription "factories" in human nuclei. J. Cell Sci. 109, 1427–1436 (1996).

    Article  CAS  PubMed  Google Scholar 

  34. Zabidi, M. A. & Stark, A. Regulatory enhancer–core–promoter communication via transcription factors and cofactors. Trends Genet. 32, 801–814 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  35. Pennacchio, L. A., Bickmore, W., Dean, A., Nobrega, M. A. & Bejerano, G. Enhancers: five essential questions. Nat. Rev. Genet. 14, 288–295 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  36. van Steensel, B. et al. Localization of the glucocorticoid receptor in discrete clusters in the cell nucleus. J. Cell Sci. 108, 3003–3011 (1995).

    Article  Google Scholar 

  37. Ghamari, A. et al. In vivo live imaging of RNA polymerase II transcription factories in primary cells. Genes Dev. 27, 767–777 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  38. Mir, M. et al. Dense Bicoid hubs accentuate binding along the morphogen gradient. Genes Dev. 31, 1784–1794 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  39. Li, J. et al. Single-molecule nanoscopy elucidates RNA polymerase II transcription at single genes in live cells. Cell 178, 491–506 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  40. Crocker, J. et al. Low affinity binding site clusters confer Hox specificity and regulatory robustness. Cell 160, 191–203 (2015).

    Article  CAS  PubMed  Google Scholar 

  41. Farley, E. K. et al. Suboptimization of developmental enhancers. Science 350, 325–328 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

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

    Article  CAS  PubMed  Google Scholar 

  43. Allahyar, A. et al. Enhancer hubs and loop collisions identified from single-allele topologies. Nat. Genet. 50, 1151–1160 (2018).

    Article  CAS  PubMed  Google Scholar 

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

  45. Boija, A. et al. Transcription factors activate genes through the phase-separation capacity of their activation domains. Cell 175, 1842–1855 e16 (2018).

    Article  CAS  PubMed  Google Scholar 

  46. Remenyi, A. et al. Crystal structure of a POU/HMG/DNA ternary complex suggests differential assembly of Oct4 and Sox2 on two enhancers. Genes Dev. 17, 2048–2059 (2003).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  47. Nowling, T. K., Johnson, L. R., Wiebe, M. S. & Rizzino, A. Identification of the transactivation domain of the transcription factor Sox-2 and an associated co-activator. J. Biol. Chem. 275, 3810–3818 (2000).

    Article  CAS  PubMed  Google Scholar 

  48. Dey, A. et al. A bromodomain protein, MCAP, associates with mitotic chromosomes and affects G2-to-M transition. Mol. Cell. Biol. 20, 6537–6549 (2000).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  49. Shin, Y. et al. Liquid nuclear condensates mechanically sense and restructure the genome. Cell 175, 1481–1491 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

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

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  51. Li, Y. et al. CRISPR reveals a distal super-enhancer required for Sox2 expression in mouse embryonic stem cells. PLoS ONE 9, e114485 (2014).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  52. Zhou, H. Y. et al. A Sox2 distal enhancer cluster regulates embryonic stem cell differentiation potential. Genes Dev. 28, 2699–2711 (2014).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

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

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  54. Wang, G., Hauver, J., Thomas, Z., Darst, S. A. & Pertsinidis, A. Single-molecule real-time 3D imaging of the transcription cycle by modulation interferometry. Cell 167, 1839–1852 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  55. 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 

  56. 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 

  57. Ochiai, H., Sugawara, T. & Yamamoto, T. Simultaneous live imaging of the transcription and nuclear position of specific genes. Nucleic Acids Res. 43, e127 (2015).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  58. Ochiai, H., Sugawara, T., Sakuma, T. & Yamamoto, T. Stochastic promoter activation affects Nanog expression variability in mouse embryonic stem cells. Sci. Rep. 4, 7125 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  59. Hsu, P. D. et al. DNA targeting specificity of RNA-guided Cas9 nucleases. Nat. Biotechnol. 31, 827–832 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  60. Taniguchi, M. et al. Efficient production of Cre-mediated site-directed recombinants through the utilization of the puromycin resistance gene, pac: a transient gene-integration marker for ES cells. Nucleic Acids Res. 26, 679–680 (1998).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  61. Cao, J. et al. An easy and efficient inducible CRISPR/Cas9 platform with improved specificity for multiple gene targeting. Nucleic Acids Res. 44, e149 (2016).

    PubMed  PubMed Central  Google Scholar 

  62. Grimm, J. B. et al. A general method to improve fluorophores for live-cell and single-molecule microscopy. Nat. Methods 12, 244–250 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  63. Crocker, J. C. & Grier, D. G. Methods of digital video microscopy for colloidal studies. J. Colloid Interface Sci. 179, 298–310 (1996).

    Article  CAS  Google Scholar 

  64. Bonev, B. et al. Multiscale 3D genome rewiring during mouse neural development. Cell 171, 557–572 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

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

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  66. Liu, Z. & Kraus, W. L. Catalytic-independent functions of PARP-1 determine Sox2 pioneer activity at intractable genomic loci. Mol. Cell 65, 589–603.e9 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  67. Robinson, J. T. et al. Integrative genomics viewer. Nat. Biotechnol. 29, 24–26 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  68. Zhou, X. et al. The Human Epigenome Browser at Washington University. Nat. Methods 8, 989–990 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  69. Ahuja, A. K. et al. A short G1 phase imposes constitutive replication stress and fork remodelling in mouse embryonic stem cells. Nat. Commun. 7, 10660 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  70. Fujii-Yamamoto, H., Kim, J. M., Arai, K. & Masai, H. Cell cycle and developmental regulations of replication factors in mouse embryonic stem cells. J. Biol. Chem. 280, 12976–12987 (2005).

    Article  CAS  PubMed  Google Scholar 

  71. Jorgensen, H. F. et al. The impact of chromatin modifiers on the timing of locus replication in mouse embryonic stem cells. Genome Biol. 8, R169 (2007).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

Download references

Acknowledgements

We thank J. Gao, C.-Y. Han, D. Shola and C. Yang (Rockefeller University Gene Targeting Resource Center) for 24 × MS2 targeting at the Sox2 locus. We thank B. Gu and J. Wysocka (Stanford University) for providing reagents and advice regarding dCas9 imaging. We thank L. Lavis (HHMI-Janelia) for dye-labeling reagents. This work is supported by a NYSTEM Postdoctoral Training Award (C32599GG; J.L.), the JST PRESTO program (Japan) (JPMJPR15F2; H.O.) and JSPS KAKENHI (Japan) (JP18H05531 and JP19K06612; H.O.) and partially by JST CREST (Japan) (JPMJCR16G1; H.O.), the Louis V. Gerstner, Jr. Young Investigators Fund (A.P.), a National Cancer Institute grant (P30 CA008748), a National Institutes of Health (NIH) Director’s New Innovator Award (1DP2GM105443-01; A.P.) and the National Institute of General Medical Sciences of NIH (1R01GM135545-01 and 1R21GM134342-01; A.P.).

Author information

Authors and Affiliations

Authors

Contributions

A.P. conceived, designed and supervised the study. J.L. developed and validated the experimental systems. H.O. and T.Y. established Pou5f1 and Nanog 24 × MS2-edited mESCs and materials and methods for live-cell MS2-MCP transcription imaging. J.L., A.H., Y.H., L.C. and A.P. performed experiments and analyzed the data. G.W. developed image-analysis code. A.P. wrote the manuscript.

Corresponding author

Correspondence to Alexandros Pertsinidis.

Ethics declarations

Competing interests

The authors declare no competing interests.

Additional information

Peer review information Peer reviewer reports are available. Anke Sparmann was the primary editor on this article and managed its editorial process and peer review in collaboration with the rest of the editorial team.

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

Extended data

Extended Data Fig. 1 Quantification of Brd4 and Sox2 clustering throughout mESC nuclei.

(a, b) Normalized nuclear level for endogenous as well as ectopically expressed WT and mutant SNAP-tagged proteins. SNAP-SiR fluorescence was quantified over the whole nucleus, and then the data were normalized to the mean nuclear level of the endogenous SNAP-tagged factors. (a) Endogenous, as well as ectopically expressed WT, BDmut and IDRdel Brd4. Data points are from 2 independent experiments with total n= 29, 18, 21, and 16 cells for endogenous, WT, Bdmut, and IDRdel, respectively. (b) Endogenous, as well as ectopically expressed WT, 2M, TAD, and 2D Sox2. Data points are from 2 independent experiments with total n= 43, 15, 21, 18, and 20 cells for endogenous, WT, 2M, TAD, and 2D, respectively. (c, d) Scatter plots of total intensity vs. detection significance (p-value from a Wilcoxon rank-sum test) for each identified nuclear cluster. Representative data from 1 experiment are shown. (c) WT and mutant Brd4 (n=9, 8, 11, and 10 cells, for Endogenous, WT, Bdmut and IDRdel, respectively). (d) WT and mutant Sox2 (n=14, 13, 11, 10, and 11 cells, for Endogenous, WT, HMG point mutant (Sox2M) and HMG, and TAD deletion mutants (Sox2 TAD and Sox2D), respectively). Dashed lines in (c,d): p-value=0.001 cutoff. Box-plots in (a, b): boxes indicate inter-quartile range (IQR: 25%75% intervals) and the median line, whiskers indicate 1.5× the IQR; ‘×’ symbols indicate 1% and 99% percentiles; square symbols indicate the mean. Data for graphs in (a-d) are available as source data online.

Source data

Extended Data Fig. 2 Pol II and RF clustering at the endogenous Sox2 locus in mESCs.

(a) Live-cell imaging (2 μm × 2 μm ROIs, 0.8 s/scan) shows co-localized MCP-mNeonGreen and SiR-Rpb1, -Sox2, -Brd4, and –Med19 foci (yellow arrows). Bottom panels: relative MCP-Rpb1, -Sox2, -Brd4, and -Med19 xy coordinates. (b) 2D distances between MCP and factors (mean ± SD): Rpb1 102 ± 56 nm, Sox2 167 ± 81nm, Brd4 143 ± 80 nm, Med19 115 ± 57 nm (n = 12, 16, 32, and 23 cells, respectively). (c) Relative peak amplitude (A0), background level (B), r.m.s. width sxy = (sxsy)0.5, and integrated signal (πA0sxsy), for the Rpb1, Sox2, Brd4, and Med19 foci, respectively. Peak parameters are obtained by fitting to an elliptical 2D Gaussian function of the form I(x,y) = B + A0·exp[-x2/(2sx)2-y2/(2sy)2]. In the peak amplitude graph, the right-hand y-axis shows the number of molecules, calibrated based on the Pol II counts obtained by target-locking. (d and e) Organization of SNAP-Brd4 at the Sox2 locus. (d) Live-cell imaging reveals co-localized SNAP-tagged Brd4 foci with MCP-mNeonGreen-tagged nascent RNA, for WT and IDRdel, but not for BDmut Brd4. Yellow arrows indicate colocalized Brd4 clusters. (e) Quantification of Brd4 clustering at the Sox2 locus. Each data point corresponds to a single time-point measurement in a single cell (n=23, 8, 13, and 12 total cells for endogenous, WT, BDmut, and IDRdel Brd4, respectively). The left y-axis shows the size of individual clusters, normalized to the average cluster size of endogenous Brd4 at the Sox2 locus (~20 molecules). The right y-axis shows the absolute number of Brd4 molecules per cluster. The x-axis shows the nuclear level, estimated from the fitted local background level in the Sox2 ROI, normalized to the average nuclear level of endogenous Brd4. Solid lines: apparent binding curves by fitting to the function y=Ax/(x+K). Grey bar: endogenous Brd4 expression level. Box-plots in (b, c): boxes indicate inter-quartile range (IQR: 25%75% intervals) and the median line, whiskers indicate 1.5× the IQR; ‘×’ symbols indicate 1% and 99% percentiles; square symbols indicate the mean. Data for graphs in (b, c, e) are available as source data online.

Source data

Extended Data Fig. 3 Precision and reproducibility of dCas9 imaging and 2D distance measurements.

(a) Correlation between dCas9-Halo-JF646 – MCP-mNeongreen 2D distance measurements from 2 independent experiments in the extended Pou5f1 locus. Points: mean distances, error bars: S.E.M.. Solid line: linear fit. Pearson’s r = 0.93, indicates the reproducibility of the measurements. Total n= (29, 34, 23, 13, 32, 21, 17, 20, and 34) and (20, 12, 16, 17, 14, 17, 17, 13, and 16) individual transcription sites were measured in experiments 1 and 2, respectively, for the -80kb, -36kb, -16kb, -5kb, +21kb, +40kb, +90kb, +108kb, and +190kb genomic regions, respectively. (b) Absolute deviation between n=2 independent experiments. The mean absolute deviation is 20.8 nm. (c) xy scatter plots of the positions of the 9 tagged loci relative to the Pou5f1 transcription site. Dashed circles indicate the cutoff 2D distance (200 nm), used in Fig. 4c. Data for graphs in (a-c) are available as source data online.

Source data

Extended Data Fig. 4 Imaging genomic interactions in the extended Sox2 locus reveals frequent proximity of distal enhancer clusters and the target gene.

(a) Organization of the extended Sox2 locus from previously published proximity ligation (Hi-C, 5-C; accession numbers GSM2533818-21, GSM883649) and ChIP-Seq assays. ECs: enhancer clusters, characterized by presence of Sox2, p300 and H3K27ac ChIP signals. Accession numbers of ChIP-seq datasets used: GSM1910642, ENCFF001LJC, ENCFF001LJI, ENCFF001LIR, and ENCFF001KDN for Sox2, p300, Pol II, Ctcf, and H3K27ac, respectively. (b) Average distances and (c) proximity probabilities between the Sox2 transcription site (visualized with the 3′UTR MS2-tagged nascent RNA) and 5 regions between +20 kb and +300 kb from Sox2. (b) Data points: mean, error bars: S.E.M.. (c) Gray area: 95% binomial distribution confidence intervals. (b,c) Pooled data from 2 independent experiments with total n= 26, 24, 21, 38, and 27 individual transcription sites for the +21kb, +55kb, +98kb, +159kb, and +256kb genomic regions, respectively. In (b) Dashed line, indicates the 180 nm cutoff, corresponding to roughly the distance between RF clusters and the Pou5f1 transcription site. The 180nm distance cutoff is also used for the proximity probabilities shown in (c). (d) Cartoon of Sox2, highlighting frequent proximity of distal ECs to the Sox2 transcription site. Data for graphs in (b, c) are available as source data online.

Source data

Extended Data Fig. 5 A shared transcription nano-scale environment created by RF clustering underlies apparent transcription coordination between Sox2 sister chromatids.

(a) Bursting of sister chromatids at Sox2. Images show a resolvable MCP-mNeongreen doublet during the burst. (b) Representative live-cell imaging of the organization of nascent RNA and Brd4 in doublet Sox2 transcription sites. Brd4 is concentrated in the space between the two resolved transcription sites. (c and d) Localization of Brd4 clusters relative to the doublet Sox2 transcription sites (n=11 cells from 2 independent experiments). (c) Scatter plot with the MCP doublet axis aligned along x and the doublet mid-point at the origin. Each doublet is plotted twice, once for each possible orientation. Simulated Brd4 cluster localizations are based on the experimental distribution of Brd4 clusters in singlet Sox2 transcription sites (Extended Data Fig. 2a, b). (d) Distribution of distances along x for the experimental data in (c), as well as 105 simulations of n=11 doublets with independent Brd4 clusters. The experimental Brd4 distribution clearly shows Brd4 density in-between the MCP doublet, contrary to the expected bi-modal distribution if the two transcription sites each have an independent Brd4 cluster. (e) The measured mean absolute deviation of the Brd4 clusters along the axis of the MCP doublet, 157±27 nm (mean±S.E.M., n=11 cells from 2 independent experiments), is significantly smaller than the expected 228±23 nm (mean±S.D., 105 simulations of n=11 doublets with independent Brd4 clusters, p=0.0013 estimated from the numerical simulations). Data for graphs in (c, d, e) are available as source data online.

Source data

Supplementary information

Supplementary Information

Supplementary Figure 1.

Reporting Summary

Peer Review Information

Supplementary Video 1

Bursting of sister chromatids at Nanog. Maximum-intensity projection of MPCP-mNeonGreen showing a transcription burst with apparent coordination between resolvable Nanog sister chromatids. Original data consist of 2.25-μm z stacks (250-nm z steps) obtained at 10 s per stack. Movie is played at 25× speed (4 frames per s).

Supplementary Video 2

Bursting of sister chromatids at Sox2. Maximum-intensity projection of MPCP-mNeonGreen showing a transcription burst with apparent coordination between resolvable Sox2 sister chromatids. Original data consist of 4.75-μm z stacks (250-nm z steps) obtained at 20 s per stack. Movie is played at 12.5× speed (4 frames per s).

Supplementary Table 1

Sequences of gRNAs for imaging the extended Sox2 locus.

Supplementary Table 2

Sequences of gRNAs for imaging the extended Pou5f1 locus.

Source data

Source Data Fig. 1

Statistical source data

Source Data Fig. 2

Statistical source data

Source Data Fig. 3

Statistical source data

Source Data Fig. 4

Statistical source data

Source Data Fig. 5

Statistical source data

Source Data Extended Data Fig. 1

Statistical source data

Source Data Extended Data Fig. 2

Statistical source data

Source Data Extended Data Fig. 3

Statistical source data

Source Data Extended Data Fig. 4

Statistical source data

Source Data Extended Data Fig. 5

Statistical source data

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Li, J., Hsu, A., Hua, Y. et al. Single-gene imaging links genome topology, promoter–enhancer communication and transcription control. Nat Struct Mol Biol 27, 1032–1040 (2020). https://doi.org/10.1038/s41594-020-0493-6

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41594-020-0493-6

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