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:

Transcription imparts architecture, function and logic to enhancer units

Abstract

Distal enhancers play pivotal roles in development and disease yet remain one of the least understood regulatory elements. We used massively parallel reporter assays to perform functional comparisons of two leading enhancer models and find that gene-distal transcription start sites are robust predictors of active enhancers with higher resolution than histone modifications. We show that active enhancer units are precisely delineated by active transcription start sites, validate that these boundaries are sufficient for capturing enhancer function, and confirm that core promoter sequences are necessary for this activity. We assay adjacent enhancers and find that their joint activity is often driven by the stronger unit within the cluster. Finally, we validate these results through functional dissection of a distal enhancer cluster using CRISPR–Cas9 deletions. In summary, definition of high-resolution enhancer boundaries enables deconvolution of complex regulatory loci into modular units.

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: Divergent transcription identifies enhancer boundaries in high resolution.
Fig. 2: Transcription marks active eSTARR-seq enhancers.
Fig. 3: Enhancer unit boundaries reveal sequence architecture.
Fig. 4: Function and features of enhancer TSS.
Fig. 5: Functional dissection of adjacent enhancers.
Fig. 6: Dissection of the NMU enhancer.

Similar content being viewed by others

Data availability

The eSTARR-seq data are available through the ENCODE data portal (www.encodeproject.org) under accession nos. ENCSR514FNW, ENCSR729EGU and ENCSR585AGE. Processed GRO-cap data were obtained from the Gene Expression Omnibus (accession no. GSE60456). Raw sequencing files for the HiDRA study were obtained from the Sequence Read Archive (accession no. SRP118092). All candidate regulatory element clones generated in this study and used for the eSTARR-seq and luciferase assays are available upon request. Please address requests to haiyuan.yu@cornell.edu. Source data are provided with this paper.

Code availability

All analysis scripts are available as R Jupyter Notebooks on Github (https://github.com/hyulab/eSTARR).

References

  1. Serfling, E., Jasin, M. & Schaffner, W. Enhancers and eukaryotic gene transcription. Trends Genet. 1, 224–230 (1985).

    CAS  Google Scholar 

  2. Arnold, C. D. et al. Genome-wide quantitative enhancer activity maps identified by STARR-seq. Science 339, 1074–1077 (2013).

    CAS  PubMed  Google Scholar 

  3. Canver, M. C. et al. BCL11A enhancer dissection by Cas9-mediated in situ saturating mutagenesis. Nature 527, 192–197 (2015).

    CAS  PubMed  PubMed Central  Google Scholar 

  4. Tuan, D., Solomon, W., Li, Q. & London, I. M. The ‘beta-like-globin’ gene domain in human erythroid cells. Proc. Natl Acad. Sci. USA 82, 6384–6388 (1985).

    CAS  PubMed  PubMed Central  Google Scholar 

  5. Orkin, S. H. Regulation of globin gene expression in erythroid cells. Eur. J. Biochem. 231, 271–281 (1995).

    CAS  PubMed  Google Scholar 

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

    CAS  PubMed  PubMed Central  Google Scholar 

  7. Creyghton, M. P. et al. Histone H3K27ac separates active from poised enhancers and predicts developmental state. Proc. Natl Acad. Sci. USA 107, 21931–21936 (2010).

    CAS  PubMed  PubMed Central  Google Scholar 

  8. Heintzman, N. D. et al. Distinct and predictive chromatin signatures of transcriptional promoters and enhancers in the human genome. Nat. Genet. 39, 311–318 (2007).

    CAS  PubMed  Google Scholar 

  9. Dorighi, K. M. et al. Mll3 and Mll4 facilitate enhancer RNA synthesis and transcription from promoters independently of H3K4 monomethylation. Mol. Cell 66, 568–576.e4 (2017).

    CAS  PubMed  PubMed Central  Google Scholar 

  10. Henriques, T. et al. Widespread transcriptional pausing and elongation control at enhancers. Genes Dev. 32, 26–41 (2018).

    CAS  PubMed  PubMed Central  Google Scholar 

  11. Kellis, M. et al. Defining functional DNA elements in the human genome. Proc. Natl Acad. Sci. USA 111, 6131–6138 (2014).

    CAS  PubMed  PubMed Central  Google Scholar 

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

    CAS  PubMed  PubMed Central  Google Scholar 

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

    CAS  PubMed  PubMed Central  Google Scholar 

  14. Kim, T.-K. et al. Widespread transcription at neuronal activity-regulated enhancers. Nature 465, 182–187 (2010).

    CAS  PubMed  PubMed Central  Google Scholar 

  15. Engreitz, J. M. et al. Local regulation of gene expression by lncRNA promoters, transcription and splicing. Nature 539, 452–455 (2016).

    CAS  PubMed  PubMed Central  Google Scholar 

  16. Joung, J. et al. Genome-scale activation screen identifies a lncRNA locus regulating a gene neighbourhood. Nature 548, 343–346 (2017).

    CAS  PubMed  PubMed Central  Google Scholar 

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

    CAS  PubMed  PubMed Central  Google Scholar 

  18. Tippens, N. D., Vihervaara, A. & Lis, J. T. Enhancer transcription: what, where, when, and why? Genes Dev. 32, 1–3 (2018).

    CAS  PubMed  PubMed Central  Google Scholar 

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

    CAS  PubMed  PubMed Central  Google Scholar 

  20. Lu, F., Portz, B. & Gilmour, D. S. The C-terminal domain of RNA polymerase II is a multivalent targeting sequence that supports Drosophila development with only consensus heptads. Mol. Cell 73, 1232–1242.e4 (2019).

    CAS  PubMed  PubMed Central  Google Scholar 

  21. Lu, H. et al. Phase-separation mechanism for C-terminal hyperphosphorylation of RNA polymerase II. Nature 558, 318–323 (2018).

    CAS  PubMed  PubMed Central  Google Scholar 

  22. Andersson, R. et al. An atlas of active enhancers across human cell types and tissues. Nature 507, 455–461 (2014).

    CAS  PubMed  PubMed Central  Google Scholar 

  23. Andersson, R. Promoter or enhancer, what’s the difference? Deconstruction of established distinctions and presentation of a unifying model. Bioessays 37, 314–323 (2015).

    PubMed  Google Scholar 

  24. Vo Ngoc, L., Wang, Y. L., Kassavetis, G. A. & Kadonaga, J. T. The punctilious RNA polymerase II core promoter. Genes Dev. 31, 1289–1301 (2017).

    PubMed  PubMed Central  Google Scholar 

  25. Inoue, F. et al. A systematic comparison reveals substantial differences in chromosomal versus episomal encoding of enhancer activity. Genome Res. 27, 38–52 (2017).

    CAS  PubMed  PubMed Central  Google Scholar 

  26. Muerdter, F. et al. Resolving systematic errors in widely used enhancer activity assays in human cells. Nat. Methods 15, 141–149 (2017).

    PubMed  PubMed Central  Google Scholar 

  27. Klein, J. et al. A systematic evaluation of the design, orientation, and sequence context dependencies of massively parallel reporter assays. Preprint at bioRxiv https://www.biorxiv.org/content/10.1101/576405v1 (2019).

  28. Kristjándsóttir, K. et al. Population-scale study of eRNA transcription reveals bipartite functional enhancer architecture. Preprint at bioRxiv https://www.biorxiv.org/content/10.1101/426908v2 (2018).

  29. Mikhaylichenko, O. et al. The degree of enhancer or promoter activity is reflected by the levels and directionality of eRNA transcription. Genes Dev. 32, 42–57 (2018).

    CAS  PubMed  PubMed Central  Google Scholar 

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

    CAS  PubMed  Google Scholar 

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

    CAS  PubMed  PubMed Central  Google Scholar 

  32. Paulson, M., Press, C., Smith, E., Tanese, N. & Levy, D. E. IFN-stimulated transcription through a TBP-free acetyltransferase complex escapes viral shutoff. Nat. Cell Biol. 4, 140–147 (2002).

    CAS  PubMed  Google Scholar 

  33. Zabidi, M. A. et al. Enhancer–core-promoter specificity separates developmental and housekeeping gene regulation. Nature 518, 556–559 (2015).

    CAS  PubMed  Google Scholar 

  34. Haberle, V. et al. Transcriptional cofactors display specificity for distinct types of core promoters. Nature 570, 122–126 (2019).

    CAS  PubMed  Google Scholar 

  35. Grossman, S. R. et al. Positional specificity of different transcription factor classes within enhancers. Proc. Natl Acad. Sci. USA 115, E7222–E7230 (2018).

    CAS  PubMed  PubMed Central  Google Scholar 

  36. Yang, X. & Vingron, M. Classifying human promoters by occupancy patterns identifies recurring sequence elements, combinatorial binding, and spatial interactions. BMC Biol. 16, 138 (2018).

    CAS  PubMed  PubMed Central  Google Scholar 

  37. Wang, X. et al. High-resolution genome-wide functional dissection of transcriptional regulatory regions and nucleotides in human. Nat. Commun. 9, 5380 (2018).

    CAS  PubMed  PubMed Central  Google Scholar 

  38. Gasperini, M. et al. A genome-wide framework for mapping gene regulation via cellular genetic screens. Cell 176, 377–390.e19 (2019).

    CAS  PubMed  PubMed Central  Google Scholar 

  39. Dukler, N., Gulko, B., Huang, Y.-F. & Siepel, A. Is a super-enhancer greater than the sum of its parts? Nat. Genet. 49, 2–3 (2016).

    PubMed  PubMed Central  Google Scholar 

  40. Shin, H. Y. et al. Hierarchy within the mammary STAT5-driven Wap super-enhancer. Nat. Genet. 48, 904–911 (2016).

    CAS  PubMed  PubMed Central  Google Scholar 

  41. Kwak, H., Fuda, N. J., Core, L. J. & Lis, J. T. Precise maps of RNA polymerase reveal how promoters direct initiation and pausing. Science 339, 950–953 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  42. Smith, R. P. et al. Massively parallel decoding of mammalian regulatory sequences supports a flexible organizational model. Nat. Genet. 45, 1021–1028 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  43. Vierstra, J. et al. Mouse regulatory DNA landscapes reveal global principles of cis-regulatory evolution. Science 346, 1007–1012 (2014).

    CAS  PubMed  PubMed Central  Google Scholar 

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

    CAS  PubMed  Google Scholar 

  45. Shao, W., Alcantara, S. G. & Zeitlinger, J. Reporter-ChIP-nexus reveals strong contribution of the Drosophila initiator sequence to RNA polymerase pausing. Elife 8, e41461 (2019).

    PubMed  PubMed Central  Google Scholar 

  46. Larsson, A. J. M. et al. Genomic encoding of transcriptional burst kinetics. Nature 565, 251–254 (2019).

    CAS  PubMed  PubMed Central  Google Scholar 

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

    CAS  PubMed  PubMed Central  Google Scholar 

  48. Hay, D. et al. Genetic dissection of the α-globin super-enhancer in vivo. Nat. Genet. 48, 895–903 (2016).

    CAS  PubMed  PubMed Central  Google Scholar 

  49. Huang, J. et al. Dynamic control of enhancer repertoires drives lineage and stage-specific transcription during hematopoiesis. Dev. Cell 36, 9–23 (2016).

    PubMed  PubMed Central  Google Scholar 

  50. Kim, H. S. et al. Pluripotency factors functionally premark cell-type-restricted enhancers in ES cells. Nature 556, 510–514 (2018).

    CAS  PubMed  PubMed Central  Google Scholar 

  51. Wei, X. et al. A massively parallel pipeline to clone DNA variants and examine molecular phenotypes of human disease mutations. PLoS Genet. 10, e1004819 (2014).

    PubMed  PubMed Central  Google Scholar 

  52. Arad, U. Modified Hirt procedure for rapid purification of extrachromosomal DNA from mammalian cells. Biotechniques 24, 760–762 (1998).

    CAS  PubMed  Google Scholar 

  53. Picelli, S. et al. Tn5 transposase and tagmentation procedures for massively scaled sequencing projects. Genome Res. 24, 2033–2040 (2014).

    CAS  PubMed  PubMed Central  Google Scholar 

  54. Wang, Z., Martins, A. L. & Danko, C. G. RTFBSDB: an integrated framework for transcription factor binding site analysis. Bioinformatics 32, 3024–3026 (2016).

    CAS  PubMed  PubMed Central  Google Scholar 

  55. Chow, R. D. et al. In vivo profiling of metastatic double knockouts through CRISPR–Cpf1 screens. Nat. Methods 16, 405–408 (2019).

    CAS  PubMed  PubMed Central  Google Scholar 

  56. Ran, F. A. et al. Genome engineering using the CRISPR–Cas9 system. Nat. Protoc. 8, 2281–2308 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  57. Stringer, B. W. et al. A reference collection of patient-derived cell line and xenograft models of proneural, classical and mesenchymal glioblastoma. Sci. Rep. 9, 4902 (2019).

    PubMed  PubMed Central  Google Scholar 

Download references

Acknowledgements

The human pSTARR-seq plasmid was a gift from A. Stark (plasmid 71509; Addgene). We thank M. Gasperini, J. Tome and J. Shendure for sharing the clonal ∆eNMU K562 cells and helpful advice. We thank C. Fulco and J. Engreitz for helpful discussions and guidance. This work was supported by grants from the National Institutes of Health (HG009393 to J.T.L. and H.Y.; GM25232 to J.T.L.; DK115398 and HG008126 to H.Y.). N.D.T. was supported by a Cornell University Center for Vertebrate Genomics Scholarship and National Institutes of Health training grant T32HD057854.

Author information

Authors and Affiliations

Authors

Contributions

N.D.T., J.L., A.O., J.T.L. and H.Y. conceived the project and designed the enhancer comparison screen. N.D.T. conceived the dissecting enhancer cooperativity and mechanisms. J.L. performed cloning, primer design, Cas9 deletions and all eSTARR- and Clone-seq assays. N.D.T. optimized and prepared the enhancer fusions with guidance from A.O., H.Y. and J.T.L. N.D.T. and A.K-Y.L. performed the analysis with feedback from J.G.B., J.L., A.O., J.T.L. and H.Y. S.D.W. designed the single guide RNAs. N.D.T. wrote the manuscript with feedback from all authors.

Corresponding authors

Correspondence to John T. Lis or Haiyuan Yu.

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 Design and validation of eSTARR-seq and selected candidates.

a, Size distribution of candidates is shown by ChromHMM class. b, Correlation between luciferase, STARR-seq, and eSTARR-seq reporter activity in HeLa cells. Luciferase and STARR-seq data are from (Arnold et al., 2013). c, eSTARR-seq activity is shown relative to each elements’ size for both candidate elements (blue) and negative controls (gray). Line indicates a fitted loess curve estimate of size bias for eSTARR-seq and 95% confidence interval in gray.

Extended Data Fig. 2 Comparison with the SCP1 promoter.

a, Correlation between replicates using SCP1. b, eSTARR-seq activity vs element length using SCP1, averaged from n = 3 transfection replicates. c, eSTARR-seq activity in forward vs reverse cloning orientations using SCP1 (averaged from n = 3). d, Percent of elements from each ChromHMM class with significant enhancer activity for SCP1. Error bars indicate standard error calculated for a sample of binary trials, centered on the observed success rate. e, SCP1 eSTARR-seq activity of elements cloned using TSS + 60 bp boundaries (x) or TSS + 200 boundaries (y). Gray area shows 95% confidence interval of linear regression from n = 93 elements. f, eSTARR-seq activity of MYC (x) vs SCP1 (y) as the promoter. Colors indicate enhancers shared by both promoters (blue), active with only one promoter (red), or inactive with both promoters (gray). g, Percent of elements from each ChromHMM class with significant enhancer activity for both MYC promoter and SCP1. Error bars indicate standard error calculated for a sample of binary trials, centered on the observed probability. h, Venn diagram showing overlap of the MYC promoter and SCP1 active enhancer sets.

Extended Data Fig. 3 Validation of strand bias and TSS function from HiDRA.

a, Pie chart indicating the fraction of HiDRA fragments tested in one (gray) or both (gold) orientations. Some fragments have pairings with more than one fragment in the opposing orientation, providing 763,000 distinct pairs. b, Comparison of HiDRA enhancer activities from opposing orientations of fragment pairs. Color indicates the number of pairs. Gray lines denote approximate statistical cut-off for active enhancers. Quadrants II and III denote orientation-dependent ‘enhancer’ fragment pairs; quadrant IV fragments are active in both orientations. c, Pie chart indicating the percent of HiDRA fragment pairs classified as inactive, orientation-dependent, and orientation-independent. de, Bar charts indicating the percentage of orientation-independent enhancer calls from HiDRA fragments sample from DHSs within the indicated ChromHMM classes. d, fragments are further classified as untranscribed or transcribed (contains divergent GRO-cap TSSs). P-values are from two-sided Fisher’s exact test between indicated ratio and total enhancer ratio (140/4,367). e, fragments are sampled from different areas around unpaired GRO-cap TSSs (see cartoon and Methods). Raw fragment counts are shown above each bar. Gray line marks the average percent activity of all fragments. P-values are from two-sided Fisher’s exact test between indicated ratio and total enhancer ratio (402/11,579). All error bars indicate standard error calculated for a sample of binary trials, centered on the observed probability.

Extended Data Fig. 4 Orientation dependence in the HiDRA dataset.

a, Comparison of forward vs reverse cloning orientation for HiDRA fragments overlapping GM12878 DHS peaks. Data points are shown as log2 fold-change of RNA vs DNA read counts. Elements with significantly elevated activity in both orientations are called orientation-independent enhancers (green). Elements with significantly elevated activity in one orientation are called orientation-dependent (black). Remaining fragments are called inactive (gray). b-c, Percent of orientation-dependent (b) or -independent (c) fragments within each GRO-cap and ChromHMM class. Raw fragment counts are shown above each bar. Gray line marks the percent activity of all fragments judged by the same criteria. P-values are from two-sided Fisher’s exact test between indicated ratio and total enhancer ratio (372/4,367 for b, 41/767 for c). Error bars indicate standard error calculated for a sample of binary trials, centered on the observed probability.

Extended Data Fig. 5 Features of eSTARR-seq enhancers.

a, Scatterplot of activity vs GRO-cap reads from eSTARR enhancers in K562 cells. b, Metaplots of average H3K27ac, H3K4me3, and H3K4me1 ChIP-seq signal from different element classes defined in K562 cells. Promoters are defined as GRO-cap divergent TSSs within 500 bp of GENCODE gene start, whereas enhancers are defined as GRO-cap divergent TSSs with significant eSTARR activity. Below, ChIP-seq to GRO-cap signal ratio is shown within the window. c, Metaplots of average H3K27ac, H3K4me3, and H3K4me1 ChIP-seq signal from different element classes defined in GM12878 cells. Promoters are defined as GRO-cap divergent TSSs within 500 bp of GENCODE gene start, whereas enhancers are defined as GRO-cap divergent TSSs with significant HiDRA activity. Below, ChIP-seq to GRO-cap signal ratio is shown within the window. n = 860 promoter DHS, 119 transcribed enhancer DHS, 1,100 untranscribed DHS.

Extended Data Fig. 6 Functional dissection of genomic TSS clusters.

a, Comparison of forward vs reverse cloning orientation for all tested TSS clusters. Data points are shown as log2 fold-change vs negative controls (magenta), averaged from three replicates. Positive controls (black) are known MYC or viral enhancers. Clusters with significantly elevated activity in both orientations are called enhancers (green). All other clusters are called inactive (gray). b, Comparison of sub-element activities within active enhancer clusters. The stronger sub-element is always chosen to be e1, and the weaker sub-element is e2. Gray lines indicate approximate significance cut-offs.

Extended Data Fig. 7 Design and evaluation of synthetic unit pairs.

a, Comparison of sub-element activities within synthetic enhancer clusters. The stronger sub-element is always chosen to be e1, and the weaker sub-element is e2. Gray lines indicate approximate significance cut-offs. b, Correlation between individual eSTARR-seq activities tested previously and re-tested as controls in the synthetic fusion screen (n = 48 elements). c, Agreement between predicted and observed cluster activities (‘C’) for enhancer-containing synthetic pairs. d, Agreement between predicted and observed cluster activities (‘C’) for enhancer-less synthetic pairs.

Extended Data Fig. 8 Genotyping of Cas9 deletion clones.

a, Illustration of genotyping PCR amplicon design and size relative to elements targeted for deletion. b, Table listing expected amplicon sizes from various genotypes. ‘-‘ indicates that no amplification is expected. c, Gel images from K562 clonal lines used for qRT-PCR experiments in Fig. 6. (eNMU clones were generated, genotyped and generously provided by the Shendure lab.) Genotyping PCRs were performed only once, but biological replication was achieved through independent clones.

Source data

Supplementary information

Reporting Summary

Supplementary Table 1

Primer sequences used in this study.

Source data

Source Data Fig. 1

Statistical source data.

Source Data Fig. 6

Statistical source data.

Source Data Extended Data Fig. 8

Unprocessed stained DNA gels.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Tippens, N.D., Liang, J., Leung, A.KY. et al. Transcription imparts architecture, function and logic to enhancer units. Nat Genet 52, 1067–1075 (2020). https://doi.org/10.1038/s41588-020-0686-2

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41588-020-0686-2

This article is cited by

Search

Quick links

Nature Briefing: Translational Research

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

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