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:

Spt5-mediated enhancer transcription directly couples enhancer activation with physical promoter interaction

Abstract

Active enhancers are frequently transcribed, yet the regulatory role of enhancer transcription remains debated. Here, we depleted the RNA polymerase II pausing and elongation factor Spt5 in activated mouse B cells and found that approximately 50% of enhancer–gene pairs showed co-regulated transcription, consistent with a potential functional requirement for enhancer transcription. In particular, Spt5 depletion led to loss of super-enhancer–promoter physical interaction and gene expression at the immunoglobulin heavy-chain locus (Igh), abrogating antibody class switch recombination. This defect correlated strictly with loss of enhancer transcription but did not affect acetylation of histone H3 at lysine 27, chromatin accessibility and occupancy of Mediator and cohesin at the enhancer. Strikingly, CRISPRa-mediated rescue of enhancer transcription in Spt5-depleted cells restored Igh gene expression. Our work suggests that Spt5-mediated enhancer transcription underlies the physical and functional interaction between a subset of active enhancers and their target promoters.

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: Spt5 depletion leads to loss of transcription at the Igh locus.
Fig. 2: Igh 3′RR super-enhancer function is dependent on Spt5.
Fig. 3: The Igh 3′RR super-enhancer chromatin is in an accessible and acetylated state in Spt5dep cells.
Fig. 4: CRISPRa-mediated transcriptional activation of the 3′RR partially restores Ighg1 transcription in Spt5dep cells.
Fig. 5: The maintenance of pre-established 3′RR interactions is independent of transcription, Pol II pausing and Spt5.
Fig. 6: Model for the role of Spt5-dependent enhancer transcription in Igh 3′RR–promoter interactions during antibody maturation.

Similar content being viewed by others

Data availability

All raw next-generation sequencing data (GRO-Seq, ChIP-Seq, ATAC-Seq, PRO-cap and mRNA-Seq) have been deposited in the Gene Expression Omnibus under accession number GSE132029. Source data for Figs. 1 and 2 are presented with the paper.

References

  1. Reiter, F., Wienerroither, S. & Stark, A. Combinatorial function of transcription factors and cofactors. Curr. Opin. Genet. Dev. 43, 73–81 (2017).

    CAS  PubMed  Google Scholar 

  2. Shlyueva, D., Stampfel, G. & Stark, A. Transcriptional enhancers: from properties to genome-wide predictions. Nat. Rev. Genet. 15, 272–286 (2014).

    CAS  PubMed  Google Scholar 

  3. Core, L. J. et al. Defining the status of RNA polymerase at promoters. Cell Rep. 2, 1025–1035 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  4. 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 

  5. 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 

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

    CAS  PubMed  PubMed Central  Google Scholar 

  7. Lai, F. et al. Activating RNAs associate with mediator to enhance chromatin architecture and transcription. Nature 494, 497–501 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  8. Schaukowitch, K. et al. Enhancer RNA facilitates NELF release from immediate early genes. Mol. Cell 56, 29–42 (2014).

    CAS  PubMed  PubMed Central  Google Scholar 

  9. Bose, D. A. et al. RNA binding to CBP stimulates histone acetylation and transcription. Cell 168, 135–149.e22 (2017).

    CAS  PubMed  PubMed Central  Google Scholar 

  10. Li, W. et al. Functional roles of enhancer RNAs for oestrogen-dependent transcriptional activation. Nature 498, 516–520 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  11. 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 

  12. Young, R. S., Kumar, Y., Bickmore, W. A. & Taylor, M. S. Bidirectional transcription initiation marks accessible chromatin and is not specific to enhancers. Genome Biol. 18, 242 (2017).

    PubMed  PubMed Central  Google Scholar 

  13. Paralkar, V. R. et al. Unlinking an lncRNA from its associated cis element. Mol. Cell 62, 104–110 (2016).

    CAS  PubMed  PubMed Central  Google Scholar 

  14. 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 

  15. Lloret-Llinares, M. et al. The RNA exosome contributes to gene expression regulation during stem cell differentiation. Nucleic Acids Res. 46, 11502–11513 (2018).

    CAS  PubMed  PubMed Central  Google Scholar 

  16. 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 

  17. Chen, H. et al. Dynamic interplay between enhancer–promoter topology and gene activity. Nat. Genet. 50, 1296–1303 (2018).

    CAS  PubMed  PubMed Central  Google Scholar 

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

    PubMed  PubMed Central  Google Scholar 

  19. Benabdallah, N. S. et al. Decreased enhancer–promoter proximity accompanying enhancer activation. Mol. Cell 76, 473–484.e7 (2019).

    CAS  PubMed  PubMed Central  Google Scholar 

  20. Yamaguchi, Y. et al. Structure and function of the human transcription elongation factor DSIF. J. Biol. Chem. 274, 8085–8092 (1999).

    CAS  PubMed  Google Scholar 

  21. Ivanov, D., Kwak, Y. T., Guo, J. & Gaynor, R. B. Domains in the SPT5 protein that modulate its transcriptional regulatory properties. Mol. Cell. Biol. 20, 2970–2983 (2000).

    CAS  PubMed  PubMed Central  Google Scholar 

  22. Core, L. J., Waterfall, J. J. & Lis, J. T. Nascent RNA sequencing reveals widespread pausing and divergent initiation at human promoters. Science 322, 1845–1848 (2008).

    CAS  PubMed  PubMed Central  Google Scholar 

  23. Fitz, J., Neumann, T. & Pavri, R. Regulation of RNA polymerase II processivity by Spt5 is restricted to a narrow window during elongation. EMBO J. 37, e97965 (2018).

    PubMed  PubMed Central  Google Scholar 

  24. Kieffer-Kwon, K. R. et al. Interactome maps of mouse gene regulatory domains reveal basic principles of transcriptional regulation. Cell 155, 1507–1520 (2013).

    CAS  PubMed  Google Scholar 

  25. Gressel, S. et al. CDK9-dependent RNA polymerase II pausing controls transcription initiation. eLife 6, e29736 (2017).

    PubMed  PubMed Central  Google Scholar 

  26. Shao, W. & Zeitlinger, J. Paused RNA polymerase II inhibits new transcriptional initiation. Nat. Genet. 49, 1045–1051 (2017).

    CAS  PubMed  Google Scholar 

  27. Pinaud, E. et al. The IgH locus 3′ regulatory region: pulling the strings from behind. Adv. Immunol. 110, 27–70 (2011).

    CAS  PubMed  Google Scholar 

  28. Vincent-Fabert, C. et al. Genomic deletion of the whole IgH 3′ regulatory region (hs3a, hs1,2, hs3b, and hs4) dramatically affects class switch recombination and Ig secretion to all isotypes. Blood 116, 1895–1898 (2010).

    CAS  PubMed  Google Scholar 

  29. Muramatsu, M. et al. Class switch recombination and hypermutation require activation-induced cytidine deaminase (AID), a potential RNA editing enzyme. Cell 102, 553–563 (2000).

    CAS  PubMed  Google Scholar 

  30. Chaudhuri, J. & Alt, F. W. Class-switch recombination: interplay of transcription, DNA deamination and DNA repair. Nat. Rev. Immunol. 4, 541–552 (2004).

    CAS  PubMed  Google Scholar 

  31. Stavnezer, J. & Schrader, C. E. IgH chain class switch recombination: mechanism and regulation. J. Immunol. 193, 5370–5378 (2014).

    CAS  PubMed  Google Scholar 

  32. Wuerffel, R. et al. S–S synapsis during class switch recombination is promoted by distantly located transcriptional elements and activation-induced deaminase. Immunity 27, 711–722 (2007).

    CAS  PubMed  PubMed Central  Google Scholar 

  33. Manis, J. P. et al. Class switching in B cells lacking 3′ immunoglobulin heavy chain enhancers. J. Exp. Med. 188, 1421–1431 (1998).

    CAS  PubMed  PubMed Central  Google Scholar 

  34. Manis, J. P., Michaelson, J. S., Birshtein, B. K. & Alt, F. W. Elucidation of a downstream boundary of the 3′ IgH regulatory region. Mol. Immunol. 39, 753–760 (2003).

    CAS  PubMed  Google Scholar 

  35. Bebin, A. G. et al. In vivo redundant function of the 3′ IgH regulatory element HS3b in the mouse. J. Immunol. 184, 3710–3717 (2010).

    CAS  PubMed  Google Scholar 

  36. Pinaud, E. et al. Localization of the 3′ IgH locus elements that effect long-distance regulation of class switch recombination. Immunity 15, 187–199 (2001).

    CAS  PubMed  Google Scholar 

  37. Cogne, M. et al. A class switch control region at the 3′ end of the immunoglobulin heavy chain locus. Cell 77, 737–747 (1994).

    PubMed  Google Scholar 

  38. Saintamand, A. et al. Elucidation of IgH 3′ region regulatory role during class switch recombination via germline deletion. Nat. Commun. 6, 7084 (2015).

    CAS  PubMed  Google Scholar 

  39. Pavri, R. et al. Activation-induced cytidine deaminase targets DNA at sites of RNA polymerase II stalling by interaction with Spt5. Cell 143, 122–133 (2010).

    CAS  PubMed  PubMed Central  Google Scholar 

  40. Thomas-Claudepierre, A. S. et al. Mediator facilitates transcriptional activation and dynamic long-range contacts at the IgH locus during class switch recombination. J. Exp. Med. 213, 303–312 (2016).

    CAS  PubMed  PubMed Central  Google Scholar 

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

    CAS  PubMed  PubMed Central  Google Scholar 

  42. Thomas-Claudepierre, A. S. et al. The cohesin complex regulates immunoglobulin class switch recombination. J. Exp. Med. 210, 2495–2502 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  43. Gunal-Sadik, G. et al. Stage-specific binding profiles of cohesin in resting and activated B lymphocytes suggest a role for cohesin in immunoglobulin class switching and maturation. PLoS ONE 9, e111748 (2014).

    PubMed  PubMed Central  Google Scholar 

  44. Seitan, V. C. et al. A role for cohesin in T-cell-receptor rearrangement and thymocyte differentiation. Nature 476, 467–471 (2011).

    CAS  PubMed  PubMed Central  Google Scholar 

  45. Merkenschlager, M. & Odom, D. T. CTCF and cohesin: linking gene regulatory elements with their targets. Cell 152, 1285–1297 (2013).

    CAS  PubMed  Google Scholar 

  46. Trimarchi, T. et al. Genome-wide mapping and characterization of Notch-regulated long noncoding RNAs in acute leukemia. Cell 158, 593–606 (2014).

    CAS  PubMed  PubMed Central  Google Scholar 

  47. Chavez, A. et al. Highly efficient Cas9-mediated transcriptional programming. Nat. Methods 12, 326–328 (2015).

    CAS  PubMed  PubMed Central  Google Scholar 

  48. He, M., Cortizas, E. M., Verdun, R. E. & Severinson, E. Cyclin-dependent kinases regulate Ig class switching by controlling access of AID to the switch region. J. Immunol. 194, 4231–4239 (2015).

    CAS  PubMed  PubMed Central  Google Scholar 

  49. Nagano, T. et al. Cell-cycle dynamics of chromosomal organization at single-cell resolution. Nature 547, 61–67 (2017).

    CAS  PubMed  PubMed Central  Google Scholar 

  50. Naumova, N. et al. Organization of the mitotic chromosome. Science 342, 948–953 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  51. Titov, D. V. et al. XPB, a subunit of TFIIH, is a target of the natural product triptolide. Nat. Chem. Biol. 7, 182–188 (2011).

    CAS  PubMed  PubMed Central  Google Scholar 

  52. Chao, S. H. & Price, D. H. Flavopiridol inactivates P-TEFb and blocks most RNA polymerase II transcription in vivo. J. Biol. Chem. 276, 31793–31799 (2001).

    CAS  PubMed  Google Scholar 

  53. Jonkers, I., Kwak, H. & Lis, J. T. Genome-wide dynamics of Pol II elongation and its interplay with promoter proximal pausing, chromatin, and exons. eLife 3, e02407 (2014).

    PubMed  PubMed Central  Google Scholar 

  54. Wang, Y., Lu, J. J., He, L. & Yu, Q. Triptolide (TPL) inhibits global transcription by inducing proteasome-dependent degradation of RNA polymerase II (Pol II). PLoS ONE 6, e23993 (2011).

    CAS  PubMed  PubMed Central  Google Scholar 

  55. Steurer, B. et al. Live-cell analysis of endogenous GFP-RPB1 uncovers rapid turnover of initiating and promoter-paused RNA Polymerase II. Proc. Natl Acad. Sci. USA 115, E4368–E4376 (2018).

    CAS  PubMed  PubMed Central  Google Scholar 

  56. Wada, T. et al. DSIF, a novel transcription elongation factor that regulates RNA polymerase II processivity, is composed of human Spt4 and Spt5 homologs. Genes Dev. 12, 343–356 (1998).

    CAS  PubMed  PubMed Central  Google Scholar 

  57. Nair, S. J. et al. Phase separation of ligand-activated enhancers licenses cooperative chromosomal enhancer assembly. Nat. Struct. Mol. Biol. 26, 193–203 (2019).

    CAS  PubMed  PubMed Central  Google Scholar 

  58. Melo, C. A. et al. eRNAs are required for p53-dependent enhancer activity and gene transcription. Mol. Cell 49, 524–535 (2013).

    CAS  PubMed  Google Scholar 

  59. Mousavi, K. et al. eRNAs promote transcription by establishing chromatin accessibility at defined genomic loci. Mol. Cell 51, 606–617 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  60. Kaikkonen, M. U. et al. Remodeling of the enhancer landscape during macrophage activation is coupled to enhancer transcription. Mol. Cell 51, 310–325 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  61. Wang, D. et al. Reprogramming transcription by distinct classes of enhancers functionally defined by eRNA. Nature 474, 390–394 (2011).

    CAS  PubMed  PubMed Central  Google Scholar 

  62. Rennie, S. et al. Transcription start site analysis reveals widespread divergent transcription in D. melanogaster and core promoter-encoded enhancer activities. Nucleic Acids Res. 46, 5455–5469 (2018).

    CAS  PubMed  PubMed Central  Google Scholar 

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

    CAS  PubMed  PubMed Central  Google Scholar 

  64. Rao, S. S. et al. A 3D map of the human genome at kilobase resolution reveals principles of chromatin looping. Cell 159, 1665–1680 (2014).

    CAS  PubMed  PubMed Central  Google Scholar 

  65. Braikia, F. Z. et al. Inducible CTCF insulator delays the IgH 3′ regulatory region-mediated activation of germline promoters and alters class switching. Proc. Natl Acad. Sci. USA 114, 6092–6097 (2017).

    CAS  PubMed  PubMed Central  Google Scholar 

  66. Marina-Zarate, E., Perez-Garcia, A. & Ramiro, A. R. CCCTC-binding factor locks premature IgH germline transcription and restrains class switch recombination. Front. Immunol. 8, 1076 (2017).

    PubMed  PubMed Central  Google Scholar 

  67. Pefanis, E. et al. RNA exosome-regulated long non-coding RNA transcription controls super-enhancer activity. Cell 161, 774–789 (2015).

    CAS  PubMed  PubMed Central  Google Scholar 

  68. Lim, J. et al. Nuclear proximity of Mtr4 to RNA exosome restricts DNA mutational asymmetry. Cell 169, 523–537.e15 (2017).

    CAS  PubMed  PubMed Central  Google Scholar 

  69. Germier, T. et al. Real-time imaging of a single gene reveals transcription-initiated local confinement. Biophys. J. 113, 1383–1394 (2017).

    CAS  PubMed  PubMed Central  Google Scholar 

  70. McBride, K. M. et al. Regulation of hypermutation by activation-induced cytidine deaminase phosphorylation. Proc. Natl Acad. Sci. USA 103, 8798–8803 (2006).

    CAS  PubMed  PubMed Central  Google Scholar 

  71. Core, L. J. & Lis, J. T. Transcription regulation through promoter-proximal pausing of RNA polymerase II. Science 319, 1791–1792 (2008).

    CAS  PubMed  PubMed Central  Google Scholar 

  72. Schoeberl, U. E., Kurth, H. M., Noto, T. & Mochizuki, K. Biased transcription and selective degradation of small RNAs shape the pattern of DNA elimination in Tetrahymena. Genes Dev. 26, 1729–1742 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  73. Mahat, D. B. et al. Base-pair-resolution genome-wide mapping of active RNA polymerases using precision nuclear run-on (PRO-Seq). Nat. Protoc. 11, 1455–1476 (2016).

    PubMed  PubMed Central  Google Scholar 

  74. 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 

  75. Buenrostro, J. D., Giresi, P. G., Zaba, L. C., Chang, H. Y. & Greenleaf, W. J. Transposition of native chromatin for fast and sensitive epigenomic profiling of open chromatin, DNA-binding proteins and nucleosome position. Nat. Methods 10, 1213–1218 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  76. 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 

  77. Feldman, S. et al. Constraints contributed by chromatin looping limit recombination targeting during Ig class switch recombination. J. Immunol. 194, 2380–2389 (2015).

    CAS  PubMed  Google Scholar 

  78. Baek, S., Sung, M. H. & Hager, G. L. Quantitative analysis of genome-wide chromatin remodeling. Methods Mol. Biol. 833, 433–441 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

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

    CAS  PubMed  PubMed Central  Google Scholar 

  80. Langmead, B., Trapnell, C., Pop, M. & Salzberg, S. L. Ultrafast and memory-efficient alignment of short DNA sequences to the human genome. Genome Biol. 10, R25 (2009).

    PubMed  PubMed Central  Google Scholar 

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

    PubMed  PubMed Central  Google Scholar 

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

    CAS  PubMed  PubMed Central  Google Scholar 

Download references

Acknowledgements

We are grateful to the Vienna Biocenter Core Facilities for the next-generation sequencing, and to the IMP/IMBA core facilities, especially Comparitive Medicine, BioOptics, Molecular Biology Service and Protein Chemistry. We thank K. Uzunova for the Tn5 enzyme preparation, K. Mochizuki for the GRO-Seq reagents, and C. Umkehrer and A. Obenauf for the dual sgRNA vector. We thank A. Stark, M. Busslinger, C. Bücker and A. Andersen (Life Science Editors) for critical reading of the manuscript. This work was funded by Boehringer Ingelheim and the Austrian Research Promotion Agency (headquarter grant FFG-834223). U.E.S. is supported by the L’Oréal For Women in Science Austria fellowship (2015) and Austrian Science Fund (FWF T 795-B30).

Author information

Authors and Affiliations

Authors

Contributions

J.F. designed and performed the experiments and co-wrote the manuscript. T.N. performed all of the bioinformatic analyses. M.S., E.-M.W., A.C.G., A.A. and U.E.S. designed and performed the experiments. R.P. conceived of the project, designed and performed the experiments and wrote the manuscript.

Corresponding author

Correspondence to Rushad Pavri.

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 Targeting strategy for generating Spt5-depleted (Spt5dep) mice and characterization of Spt5dep primary B cells.

a, Scheme for obtaining Spt5dep B cells (described in detail in Fitz et al. 23). Briefly, the targeting vector (Sanger Consortium) was used to generate knock-in mice, which were bred to Flp-expressing followed by EIIA-Cre-expressing mice to generate Supt5hF/+ and Supt5h-/+ genotypes, respectively. Supt5hF/F mice were crossed with Supt5h-/+Rosa26 Cre-ERT2/Cre-ERT2 mice to obtain the final genotype, Supt5hF/-Rosa26Cre-ERT2/+. b, Generation of WT (Rosa26Cre-ERT2/+) and Spt5dep (Supt5h-/-Rosa26Cre-ERT2/+) B cells. Primary B cells were isolated from spleens and cultured for 60 h with IL4 and LPS. 32 h prior to harvesting, 2 μM 4-hydroxytamoxifen (4-HT) was added to induce deletion of the floxed Supt5h exons. c, Multiplex PCR-based genotyping from genomic DNA from B cells of three Spt5dep mice at indicated times after 4-HT addition. d, Spt5 protein analysis from 10 μg nuclear extracts of WT and Spt5dep primary B cells (two mice each) treated with 4-HT for 0 h, 24 h and 32 h. Relative quantification was performed by normalization to histone H3 by setting the WT condition of each timepoint to 100%. e, Assessment of cell proliferation. Primary WT and Spt5dep B cells were stained with Cell Trace Violet upon isolation and analyzed by flow cytometry 32 h post 4-HT treatment. f, Viability curve using Trypan Blue exclusion for WT and Spt5dep primary B cells treated with 4-HT at the indicated time-points. The data represent the mean ± s.d. from eight replicates per genotype. g, Analysis of Pol II, Spt5 and histone H3 protein levels from nuclear extracts of WT and Spt5dep primary B cells treated with 4-HT for 32 h. Two replicate mice were used per genotype and a two-fold titration (3 μg and 10 μg) of nuclear extract was performed.

Extended Data Fig. 2 Replicate correlation analysis for all next-generation sequencing data used in this study.

a, Replicate correlation plots for total Pol II, Ser5P-Pol II, Ser2P-Pol II and Spt5 ChIP-seq datasets, as well as for GRO-seq, ATAC-seq and PRO-cap datasets from WT and Spt5dep primary B cells. Spearmann’s correlation coefficient (r) is indicated within the plots. b, Replicate correlation plots for mRNA-seq done in triplicate with Spearmann’s correlation coefficient indicated.

Extended Data Fig. 3 Identification of differentially expressed enhancer (DEE) groups.

a, Heatmaps in WT activated B cells centered on extragenic DNase hypersensitive site (DHS) peaks and ordered by decreasing DHS read density. Read densities of the following datasets are shown as reads per million (RPM): DHS-seq, ChIP-seq of total Pol II, serine 5-phosphorylated Pol II (Ser5P-Pol II), serine 2-phosphorylated Pol II (Ser2P-Pol II) and Spt5 and GRO-seq (sense and anti-sense strand reads are separated). DHS-seq data is from Kieffer-Kwon et al., 2013. b, Generating differentially expressed enhancer (DEE) groups. GRO-seq fold-changes (Spt5dep/WT) at extragenic DHSs (considered as putative enhancers) were clustered into three DEE groups: up-regulated, moderately down-regulated and strongly down-regulated (shown in Extended Data Fig. 5a). DEEs were linked to genes using Pol II ChIA-PET (Kieffer-Kwon et al., 2013). DEEs with (linked) or without (unlinked) PET links were divided into H3K27ac-positive (H3K27ac-pos) enhancers, H3K27ac-negative (H3K27ac-neg) enhancers and super-enhancer (SE) classes based on histone H3 lysine 27 acetylation (H3K27ac) (Whyte et al., 79). Finally, these enhancer classes were assigned to differentially expressed genes (DEGs) based on direct PET linkages (as in Extended Data Fig. 5c, d). A detailed explanation of this workflow is provided in the Methods under the Bioinformatics section. c, Upper panel: Diagram showing the calculation of the 5’ pausing index (5’PI). Note that we do not employ the conventional PI where the whole gene body is considered because of the processivity defects within gene bodies in Spt5dep cells seen in Extended Data Fig. 5a and described previously in Spt5dep MEFs (Fitz et al., 23). Lower panel: 5’ PI of DEGs assigned to DEEs. Each bar represents DEGs linked to the indicated DEE group. Each linked DEE group is split into the three enhancer classes as described in b. Statistical significance is calculated using the Student’s t test. Asterisks indicate P < 0.05 and ns = not significant.

Extended Data Fig. 4 Spt5 depletion results in elongation defects at long genes but does not lead to a global down-regulation of mRNA synthesis.

a, GRO-seq Spt5dep /WT ratio heatmap of all expressed minus-strand genes <100 kb in length (n=2943) generated using 1025 static bins and ordered by increasing gene length. Read densities are in reads per million (RPM). The transcription start site (TSS), and transcription termination site (TTS) are indicated. b, UCSC browser snapshots at one short and one long gene in WT and Spt5dep B cells. The tracks are overlaid with blue and red representing the WT and Spt5dep samples, respectively, and the overlapping region in black. The Y axis shows normalized read counts. c, UCSC browser snapshots of a representative up-regulated DEG (top) and down-regulated DEG (bottom). The WT and Spt5dep tracks are overlaid as described in b. d, Volcano plot of mRNA-seq data from WT and Spt5dep primary B cells performed in triplicate and calculated with the DEseq2 software. Genes having an expression fold-change >2 and an adjusted P value of < 0.1 are highlighted in red and considered as up- or down-regulated genes in this study. e, Box plots showing the distribution of mRNA fold-changes (Spt5dep/WT) within the corresponding GRO-seq-based DEG classes shown in Extended Data Fig. 5c. f, RT-qPCR analysis for mRNA abundance in WT and Spt5dep cells of five up-regulated and five down-regulated genes. Drosophila S2 cells were spiked in to normalize WT and Spt5dep qPCR values to the transcripts of the Drosophila-specific housekeeping gene, Act5c. The plotted data represents the mean of eight independent WT and Spt5dep samples ± s.d. Statistical significance in e and f were calculated using the two-tailed Student’s t test. Asterisks indicate P < 0.01.

Extended Data Fig. 5 A subset of enhancer-promoter pairs shows correlated transcriptional changes upon Spt5 depletion.

a, Left panel: WT GRO-seq read densities within the DEE groups (Extended Data Fig. 3b). DEEs linked to gene promoters using Pol II ChIA-PET, as well as DEEs with no PET links (unlinked DEEs), were further divided into three enhancer classes based on H3K27ac signal. DEEs with no PET links were termed unlinked enhancers. The number of DHSs in each row is indicated below the plot. Right panel: Same as before but now showing the fold-change in GRO-seq signal upon Spt5 depletion (Spt5dep/WT). Statistical significance for both plots is calculated using the two-tailed Student’s t test. Asterisks indicate P < 0.05 and ns = not significant. b, Mediator subunit, Med12, and histone H3 lysine 4 monomethylation (H3K4me1) levels within the DEE groups displayed exactly as in a. c, Left panel: Whole gene GRO-seq composite metagene analysis of WT and Spt5dep tracks representing the three differentially expressed gene (DEG) classes (Methods). Right panel: A magnified view of the gene body region (boxed in the left panel plots). d, Each bar represents DEGs linked to the indicated DEE group shown in a. Each linked DEE group is split into the three enhancer classes as described in b. The Y axis is set to 100% and the number of DEGs per bar is indicated above the bar. The numbers in white within the bars indicate the percentage of that DEG class relative to the total DEGs in that bar. Note the relatively equal distribution of DEGs showing up-, down- or intermediate regulation within each DEE group irrespective of linkage to an H3K27ac-pos enhancer, H3K27ac-neg enhancer or SE.

Extended Data Fig. 6 Effect of Spt5 depletion on B cell activation, expression of major B cells transcription factors and IgH GLTs under various stimulation conditions.

a-b, Flow cytometry analysis of a panel of B cell activation markers in WT and Spt5dep primary B cells after 32h of 4-HT treatment. Representative histogram plots for six activation markers is shown in (A) and bar graphs of mean fluorescence intensity for all tested activation markers is shown in b. Data in b represent the mean of eight independent WT and Spt5dep samples ± s.d. No statistically significant differences were detected between WT and Spt5dep activated B cells. c, RT-qPCR analysis for Igh germline transcript abundance in WT and Spt5dep cells activated with different stimulation conditions, as mentioned above each graph. For each condition, the expressed and expected Igh transcripts were analyzed, as indicated. The data represents the mean of three independent WT and Spt5dep samples ± s.d. Statistical significance was calculated using the two tailed Student’s t test (asterisks indicate P < 0.01 and ns = not significant). d, Scatter plot of mRNA-seq data for selected B cell transcription factors and cofactors as well as general transcription factors, Pol II, Mediator and cohesin (see Supplementary Table 1 for a complete list). The data plotted represent the mean reads per kilobase per million mapped reads (RPKM) of three independent WT and Spt5dep primary B cells samples. Genes showing 2-fold significant differential expression are marked in green (up-regulated) and red (down-regulated).

Extended Data Fig. 7 dCas9-VPR cannot stimulate Igh transcription in the absence of transcription.

a, Snapshot of the Igh locus showing the location of the uIghε sgRNA (~3 kb upstream of the Igh-ε promoter) relative to the IghH-γ1 and Igh-ε genes, and the 3’RR. b, Anti-Cas9 ChIP-qPCR at the hs4 and uIghε in cells infected with the indicated dCas9-VPR/sgRNA viruses, as described in Fig. 4a. c, RT-qPCR at hs4 and uIghε from the experiment in (B). For the uIghε, two sets of primers were used probing the region upstream and downstream of the sgRNA binding site. d, RT-qPCR for Igh-γ1 and Igh-μ transcripts in WT and Spt5dep cells infected with dCas9-VPR and uIghε sgRNA, as in Fig. 4a. In all cases, four independent experiments were performed and the two-tailed Student’s t test was used for statistical significance. Asterisks indicate P < 0.05 and ns, not significant.

Supplementary information

Supplementary Information

Supplementary Tables 1 and 2, and Discussion

Reporting Summary

Source data

Source Data Fig. 1

Unprocessed western blots.

Source Data Fig. 2

Gating strategy for flow cytometry.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Fitz, J., Neumann, T., Steininger, M. et al. Spt5-mediated enhancer transcription directly couples enhancer activation with physical promoter interaction. Nat Genet 52, 505–515 (2020). https://doi.org/10.1038/s41588-020-0605-6

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41588-020-0605-6

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