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:

Comparative cofactor screens show the influence of transactivation domains and core promoters on the mechanisms of transcription

Abstract

Eukaryotic transcription factors (TFs) activate gene expression by recruiting cofactors to promoters. However, the relationships between TFs, promoters and their associated cofactors remain poorly understood. Here we combine GAL4-transactivation assays with comparative CRISPR–Cas9 screens to identify the cofactors used by nine different TFs and core promoters in human cells. Using this dataset, we associate TFs with cofactors, classify cofactors as ubiquitous or specific and discover transcriptional co-dependencies. Through a reductionistic, comparative approach, we demonstrate that TFs do not display discrete mechanisms of activation. Instead, each TF depends on a unique combination of cofactors, which influences distinct steps in transcription. By contrast, the influence of core promoters appears relatively discrete. Different promoter classes are constrained by either initiation or pause-release, which influences their dynamic range and compatibility with cofactors. Overall, our comparative cofactor screens characterize the interplay between TFs, cofactors and core promoters, identifying general principles by which they influence transcription.

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: Comparative CRISPR screens identify the cofactors needed by nine different ADs.
Fig. 2: TFs display a diverse range of activation mechanisms.
Fig. 3: A direct co-dependent relationship between the tail 2 and kinase modules of the Mediator complex.
Fig. 4: TFs use different cofactors to regulate different steps in transcription.
Fig. 5: Comparative screens show discrete cofactor preferences dictated by core promoter elements.
Fig. 6: Cofactor–promoter compatibility is influenced by the rate-limiting step in transcription.

Similar content being viewed by others

Data availability

All high-throughput sequencing data relevant to this study have been deposited in the NCBI Gene Expression Omnibus under primary accession code GSE198944. All of the relevant source data has been provided with the paper. Source data are provided with this paper.

Code availability

The paper does not include any custom code beyond the implementation of pre-existing publicly available software packages. All computational analysis can be reproduced from the descriptions provided in Methods using the listed publicly available software.

References

  1. Lambert, S. A. et al. The human transcription factors. Cell 172, 650–665 (2018).

    Article  CAS  PubMed  Google Scholar 

  2. Ptashne, M. & Gann, A. Transcriptional activation by recruitment. Nature 386, 569–577 (1997).

    Article  CAS  PubMed  Google Scholar 

  3. Roeder, R. G. Transcriptional regulation and the role of diverse coactivators in animal cells. FEBS Lett. 579, 909–915 (2005).

    Article  CAS  PubMed  Google Scholar 

  4. Ferrie, J. J., Karr, J. P., Tjian, R. & Darzacq, X. “Structure”–function relationships in eukaryotic transcription factors: the role of intrinsically disordered regions in gene regulation. Mol. Cell 82, 3970–3984 (2022).

    Article  CAS  PubMed  Google Scholar 

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

    Article  CAS  PubMed  Google Scholar 

  6. Levine, M., Cattoglio, C. & Tjian, R. Looping back to leap forward: transcription enters a new era. Cell 157, 13–25 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

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

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  8. Neumayr, C. et al. Differential cofactor dependencies define distinct types of human enhancers. Nature 606, 406–413 (2022).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  9. Stampfel, G. et al. Transcriptional regulators form diverse groups with context-dependent regulatory functions. Nature 528, 147–151 (2015).

    Article  CAS  PubMed  Google Scholar 

  10. 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 

  11. Alerasool, N., Leng, H., Lin, Z.-Y., Gingras, A.-C. & Taipale, M. Identification and functional characterization of transcriptional activators in human cells. Mol. Cell 82, 677–695.e7 (2022).

    Article  CAS  PubMed  Google Scholar 

  12. Nemčko, F. & Stark, A. Proteome-scale identification of transcriptional activators in human cells. Mol. Cell 82, 497–499 (2022).

    Article  PubMed  Google Scholar 

  13. Donczew, R., Warfield, L., Pacheco, D., Erijman, A. & Hahn, S. Two roles for the yeast transcription coactivator SAGA and a set of genes redundantly regulated by TFIID and SAGA. eLife 9, e50109 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  14. Bergman, D. T. et al. Compatibility rules of human enhancer and promoter sequences. Nature 106, 176–184 (2022).

    Article  Google Scholar 

  15. Martinez-Ara, M., Comoglio, F., van Arensbergen, J. & van Steensel, B. Systematic analysis of intrinsic enhancer–promoter compatibility in the mouse genome. Mol. Cell 82, 2519–2531.e6 (2022).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  16. van Arensbergen, J., van Steensel, B. & Bussemaker, H. J. In search of the determinants of enhancer–promoter interaction specificity. Trends Cell Biol. 24, 695–702 (2014).

    Article  PubMed  PubMed Central  Google Scholar 

  17. Galouzis, C. C. & Furlong, E. E. M. Regulating specificity in enhancer–promoter communication. Curr. Opin. Cell Biol. 75, 102065 (2022).

    Article  CAS  PubMed  Google Scholar 

  18. Kim, S. & Wysocka, J. Deciphering the multi-scale, quantitative cis-regulatory code. Mol. Cell 83, 373–392 (2023).

    Article  CAS  PubMed  Google Scholar 

  19. Schmid-Burgk, J. L., Höning, K., Ebert, T. S. & Hornung, V. CRISPaint allows modular base-specific gene tagging using a ligase-4-dependent mechanism. Nat. Commun. 7, 12338 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  20. Pattabiraman, D. R. et al. Interaction of c-Myb with p300 is required for the induction of acute myeloid leukemia (AML) by human AML oncogenes. Blood 123, 2682–2690 (2014).

    Article  CAS  PubMed  Google Scholar 

  21. Vojnic, E. et al. Structure and VP16 binding of the mediator Med25 activator interaction domain. Nat. Struct. Mol. Biol. 18, 404–409 (2011).

    Article  CAS  PubMed  Google Scholar 

  22. Mittler, G. et al. A novel docking site on mediator is critical for activation by VP16 in mammalian cells. EMBO J. 22, 6494–6504 (2003).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  23. Yang, F., DeBeaumont, R., Zhou, S. & Näär, A. M. The activator-recruited cofactor/mediator coactivation subunit ARC92 is a functionally important target of the VP16 transcriptional activator. Proc. Natl Acad. Sci. USA 101, 2339–2344 (2004).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  24. Sandberg, M. L. et al. c-Myb and p300 regulate hematopoietic stem cell proliferation and differentiation. Dev. Cell 8, 153–166 (2005).

    Article  CAS  PubMed  Google Scholar 

  25. Xu, Y. et al. A TFIID-SAGA perturbation that targets MYB and suppresses acute myeloid leukemia. Cancer Cell 33, 13–28.e8 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  26. Donner, A. J., Szostek, S., Hoover, J. M. & Espinosa, J. M. CDK8 is a stimulus-specific positive coregulator of p53 target genes. Mol. Cell 27, 121–133 (2007).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  27. Chung, C. Y. et al. Cbx8 acts non-canonically with Wdr5 to promote mammary tumorigenesis. Cell Rep. 16, 472–486 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  28. Zheng, H. et al. Identification of integrator-PP2A complex (INTAC), an RNA polymerase II phosphatase. Science 370, eabb5872 (2020).

    Article  CAS  PubMed  Google Scholar 

  29. El Khattabi, L. et al. A pliable mediator acts as a functional rather than an architectural bridge between promoters and enhancers. Cell 178, 1145–1158 (2019).

    Article  CAS  PubMed  Google Scholar 

  30. Abdella, R. et al. Structure of the human mediator-bound transcription preinitiation complex. Science 372, 52–56 (2021).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  31. Jaeger, M. G. et al. Selective mediator dependence of cell-type-specifying transcription. Nat. Genet. 52, 719–727 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  32. Warfield, L., Donczew, R., Mahendrawada, L. & Hahn, S. Yeast mediator facilitates transcription initiation at most promoters via a tail-independent mechanism. Mol. Cell 82, 4033–4048.e7 (2022).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  33. Nabet, B. et al. The dTAG system for immediate and target-specific protein degradation. Nat. Chem. Biol. 14, 431–441 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

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

    Article  CAS  PubMed  Google Scholar 

  35. Aoi, Y. et al. NELF regulates a promoter-proximal step distinct from RNA Pol II pause-release. Mol. Cell 78, 261–274.e5 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  36. Steinparzer, I. et al. Transcriptional responses to IFN-γ require mediator kinase-dependent pause release and mechanistically distinct CDK8 and CDK19 functions. Mol. Cell 76, 485–499.e8 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  37. Leach, K. M. et al. Characterization of the human β-globin downstream promoter region. Nucleic Acids Res. 31, 1292–1301 (2003).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  38. Stewart, J. J., Fischbeck, J. A., Chen, X. & Stargell, L. A. Non-optimal TATA elements exhibit diverse mechanistic consequences. J. Biol. Chem. 281, 22665–22673 (2006).

    Article  CAS  PubMed  Google Scholar 

  39. Stewart, J. J. & Stargell, L. A. The stability of the TFIIA–TBP–DNA complex is dependent on the sequence of the TATAAA element. J. Biol. Chem. 276, 30078–30084 (2001).

    Article  CAS  PubMed  Google Scholar 

  40. Smale, S. T. & Kadonaga, J. T. The RNA polymerase II core promoter. Annu. Rev. Biochem. 72, 449–479 (2003).

    Article  CAS  PubMed  Google Scholar 

  41. Kadonaga, J. T. Perspectives on the RNA polymerase II core promoter. Wiley Interdiscip. Rev. Dev. Biol. 1, 40–51 (2012).

    Article  CAS  PubMed  Google Scholar 

  42. Morachis, J. M., Murawsky, C. M. & Emerson, B. M. Regulation of the p53 transcriptional response by structurally diverse core promoters. Genes Dev. 24, 135–147 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  43. 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).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  44. Gilchrist, D. A. et al. Pausing of RNA polymerase II disrupts DNA-specified nucleosome organization to enable precise gene regulation. Cell 143, 540–551 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  45. Core, L. & Adelman, K. Promoter-proximal pausing of RNA polymerase II: a nexus of gene regulation. Genes Dev. 33, 960–982 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  46. Parry, T. J. et al. The TCT motif, a key component of an RNA polymerase II transcription system for the translational machinery. Genes Dev. 24, 2013–2018 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  47. Patel, A. B. et al. Structure of human TFIID and mechanism of TBP loading onto promoter DNA. Science 362, eaau8872 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  48. Petrenko, N., Jin, Y., Dong, L., Wong, K. H. & Struhl, K. Requirements for RNA polymerase II preinitiation complex formation in vivo. eLife 8, e43654 (2019).

    Article  PubMed  PubMed Central  Google Scholar 

  49. Louder, R. K. et al. Structure of promoter-bound TFIID and model of human pre-initiation complex assembly. Nature 531, 604–609 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  50. Petrenko, N. & Struhl, K. Comparison of transcriptional initiation by RNA polymerase II across eukaryotic species. eLife 10, e67964 (2021).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  51. Hoopes, B. C., LeBlanc, J. F. & Hawley, D. K. Contributions of the TATA box sequence to rate-limiting steps in transcription initiation by RNA polymerase II. J. Mol. Biol. 277, 1015–1031 (1998).

    Article  CAS  PubMed  Google Scholar 

  52. Yean, D. & Gralla, J. Transcription reinitiation rate: a special role for the TATA box. Mol. Cell. Biol. 17, 3809–3816 (1997).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  53. Serebreni, L. et al. Functionally distinct promoter classes initiate transcription via different mechanisms reflected in focused versus dispersed initiation patterns. EMBO J. 42, e113519 (2023).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  54. Wang, Y. L. et al. TRF2, but not TBP, mediates the transcription of ribosomal protein genes. Genes Dev. 28, 1550–1555 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  55. Klumpe, H. E. et al. The context-dependent, combinatorial logic of BMP signaling. Cell Syst. 13, 388–407.e10 (2022).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  56. Su, C. J. et al. Ligand–receptor promiscuity enables cellular addressing. Cell Syst. 13, 408–425.e12 (2022).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  57. Klumpe, J. E., Garcia-Ojalvo, J., Elowitz, M. B. & Antebi, Y. E. The computational capabilities of many-to-many protein interaction networks. Cell Syst. 14, 430–446 (2023).

    Article  CAS  PubMed  Google Scholar 

  58. Danko, C. G. et al. Signaling pathways differentially affect RNA polymerase II initiation, pausing, and elongation rate in cells. Mol. Cell 50, 212–222 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  59. Blau, J. et al. Three functional classes of transcriptional activation domain. Mol. Cell. Biol. 16, 2044–2055 (1996).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  60. Harden, T. T., Vincent, B. J. & DePace, A. H. Transcriptional activators in the early Drosophila embryo perform different kinetic roles. Cell Syst. 14, 258–272 (2023).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  61. Herschlag, D. & Johnson, F. B. Synergism in transcriptional activation: a kinetic view. Genes Dev. 7, 173–179 (1993).

    Article  CAS  PubMed  Google Scholar 

  62. Martinez-Corral, R. et al. Transcriptional kinetic synergy: a complex landscape revealed by integrating modeling and synthetic biology. Cell Syst. 14, 324–339 (2023).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  63. Scholes, C., DePace, A. H. & Sánchez, Á. Combinatorial gene regulation through kinetic control of the transcription cycle. Cell Syst. 4, 97–108.e9 (2017).

    Article  CAS  PubMed  Google Scholar 

  64. Dudnyk, K., Cai, D., Shi, C., Xu, J. & Zhou, J. Sequence basis of transcription initiation in human genome. Science 384, 6694 (2024).

    Article  Google Scholar 

  65. Li, X. C., Fuqua, T., van Breugel, M. E. & Crocker, J. Mutational scans reveal differential evolvability of Drosophila promoters and enhancers. Philos. Trans. R. Soc. Lond. B 378, 20220054 (2023).

    Article  CAS  Google Scholar 

  66. Sahu, B. et al. Sequence determinants of human gene regulatory elements. Nat. Genet. 54, 283–294 (2022).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  67. Yang, C., Bolotin, E., Jiang, T., Sladek, F. M. & Martinez, E. Prevalence of the initiator over the TATA box in human and yeast genes and identification of DNA motifs enriched in human TATA-less core promoters. Gene 389, 52–65 (2007).

    Article  CAS  PubMed  Google Scholar 

  68. Burr, M. L. et al. An evolutionarily conserved function of polycomb silences the MHC class I antigen presentation pathway and enables immune evasion in cancer. Cancer Cell 36, 385–401.e8 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  69. Joung, J. et al. Genome-scale CRISPR–Cas9 knockout and transcriptional activation screening. Nat. Protoc. 12, 828–863 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

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

    Article  Google Scholar 

  71. Langmead, B. & Salzberg, S. L. Fast gapped-read alignment with Bowtie 2. Nat. Methods 9, 357–359 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  72. Michlits, G. et al. CRISPR–UMI: single-cell lineage tracing of pooled CRISPR–Cas9 screens. Nat. Methods 14, 1191–1197 (2017).

    Article  CAS  PubMed  Google Scholar 

  73. Parnas, O. et al. A genome-wide CRISPR screen in primary immune cells to dissect regulatory networks. Cell 162, 675–686 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  74. Replogle, J. M. et al. Mapping information-rich genotype–phenotype landscapes with genome-scale perturb-seq. Cell 185, 2559–2575.e28 (2022).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  75. Li, W. et al. MAGeCK enables robust identification of essential genes from genome-scale CRISPR–Cas9 knockout screens. Genome Biol. 15, 554 (2014).

    Article  PubMed  PubMed Central  Google Scholar 

  76. Zhu, R., del Rio-Salgado, J. M., Garcia-Ojalvo, J. & Elowitz, M. B. Synthetic multistability in mammalian cells. Science 375, eabg9765 (2022).

    Article  CAS  PubMed  Google Scholar 

  77. Liu, M. et al. Genomic discovery of potent chromatin insulators for human gene therapy. Nat. Biotechnol. 33, 198–203 (2015).

    Article  PubMed  Google Scholar 

  78. Yusa, K., Zhou, L., Li, M. A., Bradley, A. & Craig, N. L. A hyperactive piggyBac transposase for mammalian applications. Proc. Natl Acad. Sci. USA 108, 1531–1536 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  79. Sakuma, T., Nakade, S., Sakane, Y., Suzuki, K. I. T. & Yamamoto, T. MMEJ-assisted gene knock-in using TALENs and CRISPR–Cas9 with the PITCh systems. Nat. Protoc. 11, 118–133 (2016).

    Article  CAS  PubMed  Google Scholar 

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

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  81. He, Q., Johnston, J. & Zeitlinger, J. ChIP-nexus enables improved detection of in vivo transcription factor binding footprints. Nat. Biotechnol. 33, 395–401 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

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

    Article  PubMed  PubMed Central  Google Scholar 

  83. Ye, C. et al. DRUG-seq for miniaturized high-throughput transcriptome profiling in drug discovery. Nat. Commun. 9, 4307 (2018).

    Article  PubMed  PubMed Central  Google Scholar 

  84. Kaminow, B., Yunusov, D., Dobin, A. & Spring, C. STARsolo: accurate, fast and versatile mapping/quantification of single-cell and single-nucleus RNA-seq data. Preprint at bioRxiv https://doi.org/10.1101/2021.05.05.442755 (2021).

  85. Hao, Y. et al. Integrated analysis of multimodal single-cell data. Cell 184, 3573–3587.e29 (2021).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  86. Love, M. I., Huber, W. & Anders, S. Moderated estimation of fold change and dispersion for RNA-seq data with DESeq2. Genome Biol. 15, 550 (2014).

    Article  PubMed  PubMed Central  Google Scholar 

  87. Herzog, V. A. et al. Thiol-linked alkylation of RNA to assess expression dynamics. Nat. Methods 14, 1198–1204 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  88. Neumann, T. et al. Quantification of experimentally induced nucleotide conversions in high-throughput sequencing datasets. BMC Bioinf. 20, 258 (2019).

    Article  Google Scholar 

  89. Robinson, M. D., McCarthy, D. J. & Smyth, G. K. edgeR: a Bioconductor package for differential expression analysis of digital gene expression data. Bioinformatics 26, 139–140 (2009).

    Article  PubMed  PubMed Central  Google Scholar 

  90. Law, C. W., Chen, Y., Shi, W. & Smyth, G. K. voom: precision weights unlock linear model analysis tools for RNA-seq read counts. Genome Biol. 15, R29 (2014).

    Article  PubMed  PubMed Central  Google Scholar 

Download references

Acknowledgements

We thank all members of the Dawson Lab for their support and intellectual input throughout the project. We would also like to acknowledge the Peter MacCallum Cancer Centre Flow Cytometry and Genomics core facilities for their assistance with the research. This research was supported by Cancer Council Victoria Postdoctoral Fellowship (C.C.B.), NHMRC Investigator Grant (1196749) (M.A.D.), Cancer Council Victoria Dunlop Fellowship (M.A.D.), Howard Hughes Medical Institute international research scholarship (55008729) (M.A.D.) and ARC project grant (DP220103927) (M.A.D.).

Author information

Authors and Affiliations

Authors

Contributions

C.C.B., O.G. and M.A.D. designed the research and interpreted data. M.A.D. supervised the research, with assistance from C.C.B. C.C.B. and M.A.D. wrote the paper with helpful input from all the authors. C.C.B. performed the experiments with assistance from J.J.B., L.S., C.-S.A. and O.G. G.J.F. provided critical research support and input. L.T. performed the bioinformatic analysis with assistance from A.G. and E.Y.N.L. and input from C.C.B. and M.A.D.

Corresponding authors

Correspondence to Charles C. Bell or Mark A. Dawson.

Ethics declarations

Competing interests

M.A.D. has been a member of advisory boards for GSK, CTX CRC, Storm Therapeutics, Celgene and Cambridge Epigenetix and receives research funding from Pfizer. The other authors declare no competing interests.

Peer review

Peer review information

Nature Genetics thanks the anonymous reviewer(s) for their contribution to the peer review of this work. Peer reviewer reports are available.

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 Establishing the screening system.

(a) Schematic of the GAL4-based transactivation screening platform. ADs of interest are cloned downstream of the GAL4 DNA binding domain and recruit their respective cofactors to activate a reporter containing a fast-degrading GFP. Both constructs are lentivirally integrated. (b) Alpha-fold predictions of the structure of the 9 transcription factors used in the CRISPR screens. The specific AD region used is highlighted in green. (c) Flow cytometry analysis of demonstrating loss of GFP prior to cell death based on FSC-A and SSC-A and GFP signal at different timepoints after addition of triptolide (10uM). (d) Quantification of the reduction in GFP signal (average fluorescent signal, M.F.I) at different timepoints after the addition of triptolide (10uM). (e) Flow cytometry analysis of GFP signal at day 5 after infection with indicated sgRNAs in GAL4-VP64 and GAL4-MYB AD lines. FSC-A and SSC-A plots shown below to demonstrate predominantly live cells at the timepoint analysed. All flow analysis is performed on at least 10000 cells.

Source data

Extended Data Fig. 2 Reproducibility and validity of AD screening system.

(a) Quantification of change in GFP at day 5 after infection with sgRNAs targeting indicated cofactors in two independent AAVS reporter knock-in clones activated by different ADs. CDK8i was applied at a dose of 10uM for 24 hrs. Change in GFP in polyclonal reporter lines provided for direct comparison. Analysis of polyclonal lines was also performed 5 days after infection with sgRNAs targeting indicated cofactors. Error bars represent S.E.M of two independent sgRNAs or drug treatments. (b) Correlation plot of the fold enrichment between two independent biological replicates of the GAL4-NF-kB mCMV screens. Each replicate was performed using an independently derived reporter line likely to contain different lentiviral integration sites. r = Pearson correlation. Error bands reflect 95% confidence interval of Pearson correlation. (c) Correlation plot of the average fold enrichment scores between two independent biological replicates of VP64-AD screens. One replicate was performed using the GAL4-VP64 and a lentiviral reporter. The second replicate was performed using ZFP-VP64 and an insulated piggyBac reporter. r = Pearson correlation. Error bands reflect 95% confidence interval of Pearson correlation. (d) Flow cytometry analysis of GFP signal in each of the GAL4-AD cell lines at day 5 after infection with control sgRNAs and two independent sgRNAs targeting GAL4.

Source data

Extended Data Fig. 3 Comparative CRISPR screens across 9 ADs.

(a) Schematic of the screening strategy used to identify the cofactors required for each AD. Dropout analysis assesses the guides depleted over time in the screen to identify cofactors required for cell viability. (b) STRING analysis on the highest stringency setting using all the genes classified as a hit for at least one of the ADs. (c) Heatmap of the fold enrichment for all of the genes classified as a hit across any of the 9 AD screens. n = 239 genes. (d) Heatmap of fold enrichment for the genes defined as essential for cell viability based on dropout analysis of day 14 samples. n = 230 genes. (e) Correlation between the average fold enrichment and average effect on cell growth (across all cell lines in the DEPMAP database). Components of the RNA polymerase 2 complex are labelled. Genes in red box with blue dot are required for cell viability, but not enriched in any AD screen. n = 239 genes which are hits. (f) Quantification of the change in representation (guide counts) for essential genes and non-essential genes at each of the screen timepoints. Essential genes defined using day 14 samples, same list as (d). n = 230 essential genes. Line extends from minima to maxima, with box representing 25th percentile, 50th percentile, 75th percentile. (g) Quantification of the change in representation (guide counts) for essential genes and non-essential genes at each of the screen timepoints. Essential genes defined using DEPMAP database. Any gene with an average reduction of −0.5 across all cell lines in DEPMAP was defined as essential. n= ~2000 genes. Line extends from minima to maxima, with box representing 25th percentile, 50th percentile, 75th percentile. (h) Coefficient of variation of essential and non-essential genes across the 9 AD screens. Genes defined as essential by either day 14 dropout (left) (n = 230 genes) or DEPMAP (right) (n= ~2000 genes). > indicates that non-essential genes are higher than essential genes. n.s. = not significant, *=p-value < 0.05, paired two-sided t-test. Line extends from minima to maxima, with box representing 25th percentile, 50th percentile, 75th percentile.

Extended Data Fig. 4 INTS5 KO prevents NF-κB activity.

(a) Spoke and wheel plot of Integrator subunits across each of the ADs, demonstrating a disproportionate requirement for the NF-kB-AD. (b) Flow cytometry analysis of two TNF target genes, ICAM1 and CD69 in K562 cells with and without TNF treatment at day 4 after infection with control or INTS5 sgRNAs. (c) RNA-seq heatmap (left) and NF-kB ChIP-seq (right) in K562 cells treated for 6 hrs with TNF. RNA-seq heatmap displays genes upregulated by at least 1.5-fold, with an associated p65 ChIP-seq peak that increases by at least 2-fold. ChIP-seq heatmap displays all increased p65 peaks within 10 kb of a TNF target gene (defined above). (d) Quantification of change in RNA polymerase 2 ChIP-seq signal across TNF target genes (defined as above) and other genes, with and without TNF treatment (6 hrs) at day 4 after infection with SAFE or INTS5 sgRNAs. P-value calculated using a two-sided paired t-test between Safe + TNF and INTS5 KO + TNF. n = 1 sgRNA per gene. (e) Waterfall plot of change RNA polymerase 2 levels in INTS5 KO K562 cells treated with TNF, relative to SAFE guide control K562 cells treated with TNF. TNF target genes are highlighted in red. (f) IGV snapshot of NF-kB and RNA polymerase 2 occupancy at a target gene (ICAM1) in control and INTS5 KO cells, with and without TNF. (g) IGV snapshot of NF-kB or RNA polymerase 2 occupancy at a non-target gene (MYC) in control and INTS5 KO cells, with and without TNF.

Extended Data Fig. 5 Flow cytometry validation of heterogeneous cofactors.

Flow cytometry plots of GFP signal in each of the GAL4-AD lines at day 5 after infection with safe guide control or guides targeting a heterogeneously required cofactor. All plots are presented on the same scale and were analysed concurrently. These raw data were used to calculate the fold change in the validation heatmap in Fig. 2d. All analyses were performed on at least 10000 cells. n = two sgRNAs per gene.

Extended Data Fig. 6 Validation of cofactor heterogeneity at early timepoint.

(a) Quantification of change in GFP signal in each of the GAL4-AD lines after CDK8i treatment (24 hrs) or infection with guides targeting CDK8 and CCNC (day 5 after sgRNA infection). Fold change was calculated by comparison to SAFE guide control or DMSO treatment. N = 3 guides or independent treatments. Error bars = S.E.M. Representative FACS plots that were used to calculate the change in GFP signal are shown. (b) Quantification demonstrating concordant change in GFP signal in each of the GAL4-AD lines at 24 hrs after treatment with dTAGV-1 to degrade MED25, and infection with guides targeting MED25 at day 5 after sgRNA infection. Fold change was calculated by comparison to SAFE guide control or DMSO treatment. N = 3 guides or independent treatments. Error bars = S.E.M. Representative FACS plots that were used to calculate the change in GFP signal are shown.

Source data

Extended Data Fig. 7 Cofactor heterogeneity unlikely to be explained by regulation of GAL4 expression, differences in protein expression, stability or chromatin occupancy.

(a) Schematic of design and rationale behind EF1a screen. (b) Heatmap displaying all genes that were identified as significantly enriched in the EF1a screen and the corresponding enrichment across each of the AD screens. Box indicates the many cofactors required for EF1a expression that were not required for all ADs and did not display simple patterns that would be indicative of indirect effects. (c) Heatmap displaying the 50 most heterogeneously required cofactors across the 9 AD screens ranked by their degree of enrichment in the EF1a screen. Box highlights the large proportion of the heterogeneously required hits that are not significantly enriched in the EF1a screen, suggesting that heterogeneity is not the result of indirect effects on regulation of the GAL4-ADs. Validated hits are highlighted in bold. (d) Western blot to assess expression of the GAL4-ADs. FLAG-GAL4-AD lines were treated for 24 hrs with DMSO or cycloheximide (100 mg/ml) to assess GAL4-AD protein expression and stability. Lamin-B used as loading control. Representative blot of two biological replicates. (e) ChIP-qPCR for GAL4-AD occupancy at the UAS binding site of the reporter construct, and at a negative control region in each of the GAL4-AD cell lines. Cells where GAL4 had been removed by KO were used as a negative control. Error bars represent S.E.M of 3 technical replicates.

Source data

Extended Data Fig. 8 Co-dependencies are largely restricted to large coactivator complexes.

(a) Pearson correlation matrix of the 100 cofactors with the highest coefficient of variation across the 9 AD screens. Correlation score based on Pearson correlation distance. Clear co-dependent clusters are highlighted. Other independently validated, heterogeneous hits that do not display correlated patterns of requirement are also highlighted. (b) Fold enrichment for individual cofactors contained in co-dependent gene clusters. r = average Pearson correlation between subunits. (c) Western blot showing levels of MED12, MED14, MED25, MYC and aTubulin (loading control) after 4 hrs of addition of dTAGV-1 (500 nM) or CDK8i (10uM) in the respective dTAG knock-in cell lines. Representative blot of two biological replicates. (d) Quantification of the number of metabolically labelled SLAM-seq reads in an unlabelled control and MED14 dTAG cell line treated with DMSO or dTAGV-1 (500 nM) for 4 hrs. * = p < 0.05, two-sided paired t-test. Error bars = S.E.M of 3 biological replicates. (e) Correlation plot between fold change in RNA polymerase 2 ChIP-seq signal upon MED25 or MED14 degradation with red dots indicating which genes defined as significantly downregulated at the nascent transcriptional level upon MED14 degradation (SLAM-seq). (f) Quantification of the change in RNA polymerase 2 ChIP-seq signal upon MED12, MED14 or MED25 degradation at genes downregulated at the nascent transcriptional level upon MED14 degradation (SLAM-seq).

Source data

Extended Data Fig. 9 Elongation ratio correlates with requirement for CDK8 after accounting for the p53-AD.

(a) Correlation between the elongation index (calculated by ChIP-nexus) and the reduction in GFP signal upon CDK8i treatment (10uM), or infection with sgRNAs targeting CDK8 or CCNC (day 5 after sgRNA infection). p53 is indicated as a clear outlier. Fold reduction calculated by comparing GFP signal to matched SAFE guide or DMSO control. N = 3 guides or independent treatments. Error bars = S.E.M. (b) Same as (a) without P53 on the plot, which in a dramatically improves the correlation. N = 3 guides or independent treatments. Error bars = S.E.M. n.s = not significant, *=p < 0.05, **=p < 0.01, ***=p < 0.001.

Extended Data Fig. 10 Assessing the impact of altering the core promoter on cofactor use.

(a) Schematic of the GAL4-based promoter screening platform. Core promoters of interest are cloned downstream of the UAS and are activated by NF-kB and its associated cofactors. Both constructs are lentivirally integrated. (b) Sequences of the different core promoters that were screened. The TSS is indicated with an arrow. TATA box sequences are shown in bold and the Initiator sequence is underlined. (c) Luciferase assays performed with different promoter constructs activated by GAL4-NF-kB. Error bars represent the S.E.M of 4 replicates.

Source data

Supplementary information

Supplementary Information

Supplementary Note 1.

Reporting Summary

Peer Review File

Supplementary Table 1

Table 1. AD amino acid composition. Table 2. AD and promoter sequences. Table 3. Genes in sgRNA library. Table 4. AD screen data. Table 5. Promoter screen data. Table 6. Oligonucleotides and sgRNAs. Table 7. MOI of screens

Source data

Source Data Fig. 2

Statistical source data.

Source Data Fig. 3

Unprocessed western blot.

Source Data Fig. 4

Statistical source data.

Source Data Fig. 5

Statistical source data.

Source Data Fig. 6

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. 6

Statistical source data.

Source Data Extended Data Fig. 7

Statistical source data.

Source Data Extended Data Fig. 7

Unprocessed western blot.

Source Data Extended Data Fig. 8

Unprocessed western blot.

Source Data Extended Data Fig. 10

Statistical source data.

Rights and permissions

Springer Nature or its licensor (e.g. a society or other partner) holds exclusive rights to this article under a publishing agreement with the author(s) or other rightsholder(s); author self-archiving of the accepted manuscript version of this article is solely governed by the terms of such publishing agreement and applicable law.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Bell, C.C., Balic, J.J., Talarmain, L. et al. Comparative cofactor screens show the influence of transactivation domains and core promoters on the mechanisms of transcription. Nat Genet (2024). https://doi.org/10.1038/s41588-024-01749-z

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1038/s41588-024-01749-z

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