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:

Genome-wide characterization of mammalian promoters with distal enhancer functions

Abstract

Gene expression in mammals is precisely regulated by the combination of promoters and gene-distal regulatory regions, known as enhancers. Several studies have suggested that some promoters might have enhancer functions. However, the extent of this type of promoters and whether they actually function to regulate the expression of distal genes have remained elusive. Here, by exploiting a high-throughput enhancer reporter assay, we unravel a set of mammalian promoters displaying enhancer activity. These promoters have distinct genomic and epigenomic features and frequently interact with other gene promoters. Extensive CRISPR–Cas9 genomic manipulation demonstrated the involvement of these promoters in the cis regulation of expression of distal genes in their natural loci. Our results have important implications for the understanding of complex gene regulation in normal development and disease.

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

Figure 1: Comparison of proximal and distal DHSs with enhancer activity in two mouse cell lines.
Figure 2: CapStarr-seq with human promoters.
Figure 3: Genomic and epigenomic properties of Epromoters.
Figure 4: Expression of neighboring genes and promoter–promoter interactions.
Figure 5: Epromoters function as bona fide enhancers and regulate distal gene expression.
Figure 6: Epromoters are involved in a long-range response to IFN-α signaling.
Figure 7: eQTL association within Epromoters.

Similar content being viewed by others

Accession codes

Primary accessions

Gene Expression Omnibus

References

  1. Kim, T.K. & Shiekhattar, R. Architectural and Functional Commonalities between Enhancers and Promoters. Cell 162, 948–959 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

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

    Article  PubMed  Google Scholar 

  3. 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).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  4. Pekowska, A. et al. H3K4 tri-methylation provides an epigenetic signature of active enhancers. EMBO J. 30, 4198–4210 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  5. Koch, F. et al. Transcription initiation platforms and GTF recruitment at tissue-specific enhancers and promoters. Nat. Struct. Mol. Biol. 18, 956–963 (2011).

    Article  CAS  PubMed  Google Scholar 

  6. 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 

  7. Leung, D. et al. Integrative analysis of haplotype-resolved epigenomes across human tissues. Nature 518, 350–354 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  8. Schaffner, W. Enhancers, enhancers - from their discovery to today's universe of transcription enhancers. Biol. Chem. 396, 311–327 (2015).

    Article  CAS  PubMed  Google Scholar 

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

    Article  CAS  PubMed  Google Scholar 

  10. 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).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  11. Nguyen, T.A. et al. High-throughput functional comparison of promoter and enhancer activities. Genome Res. 26, 1023–1033 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  12. 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 

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

    Article  CAS  PubMed  Google Scholar 

  14. Schoenfelder, S. et al. The pluripotent regulatory circuitry connecting promoters to their long-range interacting elements. Genome Res. 25, 582–597 (2015).

    Article  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).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  16. Kowalczyk, M.S. et al. Intragenic enhancers act as alternative promoters. Mol. Cell 45, 447–458 (2012).

    Article  CAS  PubMed  Google Scholar 

  17. Dailey, L. High throughput technologies for the functional discovery of mammalian enhancers: new approaches for understanding transcriptional regulatory network dynamics. Genomics 106, 151–158 (2015).

    Article  CAS  PubMed  Google Scholar 

  18. Vanhille, L. et al. High-throughput and quantitative assessment of enhancer activity in mammals by CapStarr-seq. Nat. Commun. 6, 6905 (2015).

    Article  CAS  PubMed  Google Scholar 

  19. Nejepinska, J., Malik, R., Moravec, M. & Svoboda, P. Deep sequencing reveals complex spurious transcription from transiently transfected plasmids. PLoS One 7, e43283 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  20. Duttke, S.H. et al. Human promoters are intrinsically directional. Mol. Cell 57, 674–684 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  21. Roy, A.L. & Singer, D.S. Core promoters in transcription: old problem, new insights. Trends Biochem. Sci. 40, 165–171 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  22. Forrest, A.R. et al. A promoter-level mammalian expression atlas. Nature 507, 462–470 (2014).

    Article  CAS  PubMed  Google Scholar 

  23. Mahat, D.B., Salamanca, H.H., Duarte, F.M., Danko, C.G. & Lis, J.T. Mammalian Heat Shock Response and Mechanisms Underlying Its Genome-wide Transcriptional Regulation. Mol. Cell 62, 63–78 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  24. Hardison, R.C. & Taylor, J. Genomic approaches towards finding cis-regulatory modules in animals. Nat. Rev. Genet. 13, 469–483 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  25. Michaud, J. et al. HCFC1 is a common component of active human CpG-island promoters and coincides with ZNF143, THAP11, YY1, and GABP transcription factor occupancy. Genome Res. 23, 907–916 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

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

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  27. 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 

  28. 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  Google Scholar 

  29. Feuerborn, A. & Cook, P.R. Why the activity of a gene depends on its neighbors. TIG 31, 483–490 (2015).

    Article  CAS  PubMed  Google Scholar 

  30. Kagey, M.H. et al. Mediator and cohesin connect gene expression and chromatin architecture. Nature 467, 430–435 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  31. Mombaerts, P., Terhorst, C., Jacks, T., Tonegawa, S. & Sancho, J. Characterization of immature thymocyte lines derived from T-cell receptor or recombination activating gene 1 and p53 double mutant mice. Proc. Natl. Acad. Sci. USA 92, 7420–7424 (1995).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  32. Dao, L.T.M., Vanhille, L., Griffon, A., Fernandez, N. & Spicuglia, S. CapStarr-seq protocol. Protocol Exchange http://dx.doi.org/10.1038/protex.2015.096 (2015).

  33. Glusman, G., Caballero, J., Robinson, M., Kutlu, B. & Hood, L. Optimal scaling of digital transcriptomes. PLoS One 8, e77885 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  34. Reimand, J. et al. g:Profiler-a web server for functional interpretation of gene lists (2016 update). Nucleic Acids Res. 44, W1 W83-9 (2016).

    Article  Google Scholar 

  35. McLean, C.Y. et al. GREAT improves functional interpretation of cis-regulatory regions. Nat. Biotechnol. 28, 495–501 (2010).

    CAS  PubMed  PubMed Central  Google Scholar 

  36. Saeed, A.I. et al. TM4: a free, open-source system for microarray data management and analysis. Biotechniques 34, 374–378 (2003).

    Article  CAS  PubMed  Google Scholar 

  37. Kuznetsova, T. et al. Glucocorticoid receptor and nuclear factor kappa-b affect three-dimensional chromatin organization. Genome Biol. 16, 264 (2015).

    Article  PubMed  PubMed Central  Google Scholar 

  38. Tang, Z. et al. CTCF-Mediated Human 3D Genome Architecture Reveals Chromatin Topology for Transcription. Cell 163, 1611–1627 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  39. Naito, Y., Hino, K., Bono, H. & Ui-Tei, K. CRISPRdirect: software for designing CRISPR/Cas guide RNA with reduced off-target sites. Bioinformatics 31, 1120–1123 (2015).

    Article  CAS  PubMed  Google Scholar 

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

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  41. Thorvaldsdóttir, H., Robinson, J.T. & Mesirov, J.P. Integrative Genomics Viewer (IGV): high-performance genomics data visualization and exploration. Brief. Bioinform. 14, 178–192 (2013).

    Article  PubMed  Google Scholar 

  42. Stadhouders, R. et al. Multiplexed chromosome conformation capture sequencing for rapid genome-scale high-resolution detection of long-range chromatin interactions. Nat. Protoc. 8, 509–524 (2013).

    Article  CAS  PubMed  Google Scholar 

  43. Stadhouders, R. et al. HBS1L-MYB intergenic variants modulate fetal hemoglobin via long-range MYB enhancers. J. Clin. Invest. 124, 1699–1710 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  44. Vieux-Rochas, M., Fabre, P.J., Leleu, M., Duboule, D. & Noordermeer, D. Clustering of mammalian Hox genes with other H3K27me3 targets within an active nuclear domain. Proc. Natl. Acad. Sci. USA 112, 4672–4677 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  45. Stadhouders, R. et al. Control of developmentally primed erythroid genes by combinatorial co-repressor actions. Nat. Commun. 6, 8893 (2015).

    Article  CAS  PubMed  Google Scholar 

  46. Thongjuea, S., Stadhouders, R., Grosveld, F.G., Soler, E. & Lenhard, B. r3Cseq: an R/Bioconductor package for the discovery of long-range genomic interactions from chromosome conformation capture and next-generation sequencing data. Nucleic Acids Res. 41, e132 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  47. Klein, F.A. et al. FourCSeq: analysis of 4C sequencing data. Bioinformatics 31, 3085–3091 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  48. Walter, C., Schuetzmann, D., Rosenbauer, F. & Dugas, M. Basic4Cseq: an R/Bioconductor package for analyzing 4C-seq data. Bioinformatics 30, 3268–3269 (2014).

    Article  CAS  PubMed  Google Scholar 

  49. Rusinova, I. et al. Interferome v2.0: an updated database of annotated interferon-regulated genes. Nucleic Acids Res. 41, D1040–D1046 (2013).

    Article  CAS  PubMed  Google Scholar 

  50. Lawrence, M. et al. Software for computing and annotating genomic ranges. PLOS Comput. Biol. 9, e1003118 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  51. Benaglia, T., Chauveau, D., Hunter, D.R. & Young, D.S. mixtools: An R Package for Analyzing Finite Mixture Models. J. Stat. Softw. 32, 1–29 (2009).

    Article  Google Scholar 

Download references

Acknowledgements

We thank J.-C. Andrau and J. Imbert for critical reading of the manuscript. We thank the IBiSA 'Transcriptomics and Genomics Marseille-Luminy' (TGML) platform for sequencing of CapStarr-seq samples and the cell biology platform for management of cell culture. Work in the laboratory of S.S. was supported by recurrent funding from INSERM and Aix-Marseille University and by specific grants from the European Union's FP7 Programme (282510-BLUEPRINT), ARC (PJA 20151203149) and A*MIDEX (ANR-11-IDEX-0001-02). L.T.M.D., A.G. and G.C. were supported, respectively, by Vietnam International Education Development (911), CONACYT and FRM.

Author information

Authors and Affiliations

Authors

Contributions

L.T.M.D. and S.S. conceptualized and designed the experiments. L.T.M.D. performed most experimental work. A.O.G.A. performed most bioinformatics analyses. J.A.C.-M. and J.v.H. performed motif analysis. C.A.-S., T.S., D.M. and E.S. performed 4C–seq experiments and analyses. C.S., A.G. and L.V. performed and analyzed data from mouse CapStarr-seq. J.A., M.T. and N.F. contributed to CRISPR screening and analyses of allelic expression. G.C. and D.P. performed ChIA-PET analyses. A.M.R. performed eQTL analysis. All authors contributed to reading, discussion and commenting on the manuscript. L.T.M.D. and S.S. wrote the manuscript.

Corresponding author

Correspondence to Salvatore Spicuglia.

Ethics declarations

Competing interests

The authors declare no competing financial interests.

Integrated supplementary information

Supplementary Figure 1 Analysis of DHS STARR-seq in the P5424 cell line.

(a) Luciferase enhancer assays of proximal DHSs defined as active or inactive enhancers by STARR-seq in P5424 cells. For each candidate, both orientations were tested. Data represent the normalized fold change over the vector control. Error bars show s.d. from three independent transfections (***P < 0.001, **P < 0.01, *P < 0.1; two-sided Student’s t test). (b) Enrichment score of lymphoid transcription factors at distal and proximal DHSs based on ChIP–seq data from developing thymocytes. The enrichment score was calculated as the −log10 (P value) obtained with a hypergeometric test (depletion is represented by negative values).

Supplementary Figure 2 CapStarr-seq experimental control and epigenomic profiles of Epromoters in K562 and HeLa cells.

(a) IGV screenshots of STARR-seq signals for four STARR-seq-positive controls in HeLa cells. (b) FACS analysis of GFP expression in K562 (left) and HeLa (right) cells transfected with a human promoter library or empty vector. Controls were untransfected cells. The increase in GFP expression in transfected cells with the promoter library indicates potential enhancer activity in the pooled library. (c) Overlap with CpG islands (50%) and regions conserved in placental mammals (10%) using the EpiExplorer tool. The control is non-Epromoters with equal levels of gene expression as Epromoters in the same cell type. (d) Average profiles of epigenomic features for Epromoters and control promoters with the same expression pattern of associated genes. Statistical significance was calculated in a region centered on the TSS (±1 kb) using two-sided Mann–Whitney U tests. Only significant differences (P < 0.001) are shown. (e) Percentage of chromatin state 4 (strong enhancers) found in K562 Epromoters (left) and HeLa Epromoters (right) across ENCODE cell lines using the EpiExplorer tool. (f,g) Density plots of TSS positions corresponding to the selected promoter regions using CAGE peaks from ENCODE data in HeLa (f, top) and K562 (f, bottom) cells and data from FANTOM3 (g, top) and FANTOM5 (g, bottom) (Kolmogorov test). (h) Percentage of TSSs assigned to RefSeq-defined TSSs using different CAGE databases (from data in Supplementary Table 2b). (i) Comparison of the number of different RefSeq-defined TSSs per coding gene (one-sided Mann–Whitney U test).

Supplementary Figure 3 Assessment of the IFN-αβ signaling pathway.

(a) Cumulative plot of normalized RNA levels (FPKM) for genes from the IFN-αβ signaling pathway (Reactome), based on RNA–seq data from 23 cell lines. The HeLa and K562 cell lines are highlighted (Kolmogorov test). (b) Heat map showing RNA–seq relative expression (FPKM) for genes from the IFN-αβ signaling pathway (Reactome) expressed at significantly higher levels in HeLa cells as compared to the 22 remaining cell lines (SAM analysis; α = 0.5). (c) Transcription signatures related to stress/interferon response significantly enriched in the set of Epromoter-associated genes in HeLa cells (GREAT tool).

Supplementary Figure 4 Enrichment of transcription factors at Epromoters.

(a,b) Average profiles of ChIP–seq signals for ENCODE transcription factors enriched at Epromoters in K562 (a) and HeLa (b) cells. Statistical significances were calculated in a region centered on the TSS (±250 bp) using two-sided Mann–Whitney U tests.

Supplementary Figure 5 Motif enrichment at Epromoters.

(a,b) Heat maps showing the enrichment distribution (log2 (observed/expected)) of the non-redundant collection of motifs obtained by combining transcription factor binding motif (TFBM) databases (Jaspar vertebrates and Hocomoco Human). TFBMs were used to scan the extended Epromoter-associated TSS from −1 kb to +1 kb and clustering was performed based on the binding profiles in K562 (a) and HeLa (b) cells. Motifs enriched around the TSS (black line) were selected. (c,d) Significantly enriched motifs in K562 (c) and HeLa (d) cells were identified by comparing the binding enrichment within the promoter region (−200 bp to +50 bp with respect to the TSS; highlighted as orange boxes) between Epromoters and the non-Epromoters. Binding site distribution (left), motif logos (middle) and E values (right) are shown only for significantly enriched motifs (E < 0.001; χ2 test). (e) Enrichment of Epromoters and non-Epromoters as a function of the number of different TFBMs found. The enrichment score was calculated as the −log10 (P value) obtained by hypergeometric test.

Supplementary Figure 6 Proximal and distal correlations of Epromoters with gene expression.

(a) Scatterplots showing the Pearson correlation between the STARR-seq signal of Epromoters and the expression of associated genes. (b) Examples of consistent promoter–promoter interactions observed with different ChIA-PET data sets in K562 cells.

Supplementary Figure 7 Generation of knockout and knock-in cell clones via CRISPR–Cas9.

(a) General strategy for the generation of (E)promoter knockouts. Two gRNAs, G1 and G2, were designed flanking the genomic target to delete the intervening DNA segment. The CRISPR–Cas9 system creates two double-strand breaks (DSBs) at 3 - 4 nt upstream of the PAM sequences (red) and releases the excised DNA (purple). The resulting DSB is repaired by the NHEJ pathway. The genomic deletion is detected by PCR using primers P1 and P2. (bh) Assessment of (E)promoter knockout. Left, IGV screenshots showing the DNase-seq (ENCODE) and RefSeq tracks for targeted regions. The locations of gRNAs (red boxes) and the expected sizes of deleted regions are indicated. Right, PCR validation of biallelic deletion in corresponding cell clones. Details on the gRNA sequences, PCR primers and expected PCR fragments are provided in Supplementary Table 9. (i) Strategy for the generation of the inverted FAF2 Epromoter knock-in. The two gRNAs, G1 and G2, used to generate DSBs are as in the knockout experiment. The repair template contains upstream and downstream homologous arms (HAs) flanking the inverted FAF2 Epromoter. The HDR-mediated repair pathway generates the inverted FAF2 Epromoter knock-in, which is detected by PCR with the combination of two primer pairs (1 + 2) and (2 + 3). (j) PCR validation of a successful inverted FAF2 Epromoter knock-in cell clone using the combination of primers shown in i. (k) RT–PCR detection of antisense (AS) transcription in an Inv.Ep.FAF2 clone. GAPDH was used as a cDNA loading control.

Supplementary Figure 8 Interaction and epigenetic co-regulation of FAF2 and RNF44.

(a) Genomic tracks showing the 4C–seq analysis of interactions between the FAF2 (a) and RNF44 (b) promoters in WT and knockout K562 clones. The viewpoint from the FAF2 Epromoter is indicated by an arrowhead. The specific interaction between the FAF2 and RNF44 promoters is highlighted by the orange box. (b) UCSC Genome Browser tracks for H3K4me3, H3K27ac and H3K36me3 at the FAF2 locus and nearby regions across the HMEC, K562, HeLa and DND41 cell lines.

Supplementary Figure 9 Additional validations of distal gene regulation by Epromoters.

(a,b) IGV tracks for RNA–seq, ChIP–seq and ChIA-PET Pol II data in K562 and HeLa cells at the CSDE1 (a) and TAGLN2 (b) loci and nearby regions. The promoter–promoter interactions for Epromoters are highlighted in red. (c,d) qPCR analysis of gene expression in WT, ΔEp.CSDE1 (c) and ΔEp.TAGLN2 (d) clones. The number following the gene name is the number of independent cell clones. (e) Allelic frequency of the A versus T variant (chr1:160000435) in PIGM transcripts in WT, ΔEp.TAGLN2 homozygous and ΔEp.TAGLN2 heterozygous K562 clones. The total number of reads is indicated for each sample. The significant deviation of allelic frequency in heterozygous clones with respect to homozygous samples was calculated by performing a one-sided Student’s t test. (f) IGV screenshot showing tracks for RNA–seq, ChIP–seq and ChIA-PET Pol II data in K562 cells at the BAZ2B locus and the nearby region. (g) qPCR analysis of gene expression in WT and ΔEp.BAZ2B clones. Knockout of the BAZ2B Epromoter resulted in significant reduction of MARCH7 expression but had no effect on the nearby gene WDSUB1 or BAZ2B (using primers 1 and 2 shown in h). (h) UCSC Genome Browser tracks showing the different BAZ2B transcripts and primers used in g and i. (i) Alternative promoter usage for the BAZ2B gene was assessed by RT–PCR in K562 cells. The smaller fragment size observed in ΔEp.BAZ2B clones corresponds to the deletion of exon 1 (asterisk in h). (j) IGV tracks for FANTOM3 and ENCODE CAGE data, CapStarr-seq regions and RNA junctions around the TSS of the indicated gene. The red color in CapStarr-seq tracks represents active Epromoters. For c, d and g, error bars show s.d. (n = 3 independent RNA/cDNA preparation; ***P < 0.001, **P < 0.01, *P < 0.1, two-sided Student’s t test).

Supplementary Figure 10 Epromoters involved in IFN-α signaling in K562 cells.

(a) Distribution of expression correlation for ChIA-PET interacting gene pairs including at least one Epromoter (red) or excluding Epromoters (gray) and randomly rewired gene pairs (green) using RNA–seq data from ENCODE. Statistical significance was assessed by Kolmogorov test. (b) Examples of clusters of interferon response genes (green labels) associated with Epromoters (red arrows) in HeLa cells. (ce) qPCR analysis of gene expression in WT, ΔEp.YPEL4 (c,d) and ΔEp.METTL21A (e) cell clones. Error bars show s.d (n = 3 independent RNA/cDNA preparations). (f) Allelic frequency of the T versus C variant (chr11:57319339) in UBE2L6 transcripts in WT, ΔEp.YPEL4 homozygous and ΔEp.YPEL4 heterozygous K562 clones. The total number of reads is indicated for each sample. The significant deviation of allelic frequency in heterozygous clones with respect to homozygous samples was calculated by performing a one-sided Student’s t test.

Supplementary information

Supplementary Text and Figures

Supplementary Figures 1–10 and Supplementary Note. (PDF 2563 kb)

Supplementary Table 1

Mouse DHS CapStarr-seq data. (XLSX 418 kb)

Supplementary Table 2

Human promoter CapStarr-seq data. (XLSX 7980 kb)

Supplementary Table 3

Functional enrichment of Epromoter-associated genes. (XLSX 18 kb)

Supplementary Table 4

Transcription factor enrichment. (XLSX 40 kb)

Supplementary Table 5

Summary of gene pairs with promoter–promoter interactions detected by ChIA-PET. (XLSX 361 kb)

Supplementary Table 6

List of gene-pairs with ChIA-PET interaction and involved in the interferon response. (XLSX 10 kb)

Supplementary Table 7

ChIA-PET gene pairs that harbor an eQTL within the ±1.5-kb expanded promoter region and that affect the expression of the distal gene. (XLSX 5009 kb)

Supplementary Table 8

Summary of CapStarr-seq sequencing data. (XLSX 8 kb)

Supplementary Table 9

Summary of ENCODE data used in this study. (XLSX 57 kb)

Supplementary Table 10

Oligonucleotide sequences used in this study. (XLSX 23 kb)

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Dao, L., Galindo-Albarrán, A., Castro-Mondragon, J. et al. Genome-wide characterization of mammalian promoters with distal enhancer functions. Nat Genet 49, 1073–1081 (2017). https://doi.org/10.1038/ng.3884

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/ng.3884

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