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:

Cooperation of chromatin remodeling SWI/SNF complex and pioneer factor AP-1 shapes 3D enhancer landscapes

Abstract

The mechanism controlling the dynamic targeting of SWI/SNF has long been postulated to be coordinated by transcription factors (TFs), yet demonstrating a specific TF influence has proven difficult. Here we take a multi-omics approach to interrogate transient SWI/SNF interactors, chromatin targeting and the resulting three-dimensional epigenetic landscape. We utilize the labeling technique TurboID to map the SWI/SNF interactome and identify the activator protein-1 (AP-1) family members as critical interacting partners for SWI/SNF complexes. CUT&RUN profiling demonstrates SWI/SNF targeting enrichment at AP-1 bound loci, as well as SWI/SNF–AP-1 cooperation in chromatin targeting. HiChIP reveals AP-1–SWI/SNF-dependent restructuring of the three-dimensional promoter–enhancer architecture and generation of enhancer hubs. Through interrogation of the SWI/SNF–AP-1 interaction, we demonstrate an SWI/SNF dependency on AP-1-mediated chromatin localization. We propose that pioneer factors, such as AP-1, bind and target SWI/SNF to inactive chromatin, where it restructures the genomic landscape into an active state through epigenetic rewiring spanning multiple dimensions.

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: SWI/SNF interactome via proximity-based labeling.
Fig. 2: Dynamics of SWI/SNF subcomplex targeting.
Fig. 3: SWI/SNF complex interacts with the AP-1 complex and colocalizes on the genome.
Fig. 4: Cooperative targeting between the SWI/SNF complex and AP-1 complex.
Fig. 5: In vivo remodeling activity of the SWI/SNF complex measured by EChO.
Fig. 6: SWI/SNF-complex-mediated 3D enhancer landscapes.

Similar content being viewed by others

Data availability

All next-generation sequencing data are available from the Gene Expression Omnibus (GEO) under accession nos. GSE175534 and GSE196960. Source data are provided with this paper.

References

  1. Wang, W. et al. Purification and biochemical heterogeneity of the mammalian SWI-SNF complex. EMBO J. 15, 5370–5382 (1996).

    Article  CAS  Google Scholar 

  2. Olave, I. A., Reck-Peterson, S. L. & Crabtree, G. R. Nuclear actin and actin-related proteins in chromatin remodeling. Annu. Rev. Biochem. 71, 755–781 (2002).

    Article  CAS  Google Scholar 

  3. Wang, X., Haswell, J. R. & Roberts, C. W. M. Molecular pathways: SWI/SNF (BAF) complexes are frequently mutated in cancer—mechanisms and potential therapeutic insights. Clin. Cancer Res. 20, 21–27 (2014).

    Article  Google Scholar 

  4. Mittal, P. & Roberts, C. W. M. The SWI/SNF complex in cancer-biology, biomarkers and therapy. Nat. Rev. Clin. Oncol. 17, 435–448 (2020).

    Article  CAS  Google Scholar 

  5. Mashtalir, N. et al. Modular organization and assembly of SWI/SNF family chromatin remodeling complexes. Cell 175, 1272–1288 (2018).

    Article  CAS  Google Scholar 

  6. Alpsoy, A. & Dykhuizen, E. C. Glioma tumor suppressor candidate region gene 1 (GLTSCR1) and its paralog GLTSCR1-like form SWI/SNF chromatin remodeling subcomplexes. J. Biol. Chem. 293, 3892–3903 (2018).

    Article  CAS  Google Scholar 

  7. Michel, B. C. et al. A non-canonical SWI/SNF complex is a synthetic lethal target in cancers driven by BAF complex perturbation. Nat. Cell Biol. 20, 1410–1420 (2018).

    Article  CAS  Google Scholar 

  8. Gatchalian, J. et al. A non-canonical BRD9-containing BAF chromatin remodeling complex regulates naive pluripotency in mouse embryonic stem cells. Nat. Commun. 9, 5139 (2018).

    Article  Google Scholar 

  9. Wang, X. et al. BRD9 defines a SWI/SNF sub-complex and constitutes a specific vulnerability in malignant rhabdoid tumors. Nat. Commun. 10, 1881 (2019).

    Article  Google Scholar 

  10. Pan, J. et al. The ATPase module of mammalian SWI/SNF family complexes mediates subcomplex identity and catalytic activity-independent genomic targeting. Nat. Genet. 51, 618–626 (2019).

    Article  CAS  Google Scholar 

  11. Hodges, H. C. et al. Dominant-negative SMARCA4 mutants alter the accessibility landscape of tissue-unrestricted enhancers. Nat. Struct. Mol. Biol. 25, 61–72 (2018).

    Article  CAS  Google Scholar 

  12. Schick, S. et al. Acute BAF perturbation causes immediate changes in chromatin accessibility. Nat. Genet. 53, 269–278 (2021).

    Article  CAS  Google Scholar 

  13. Iurlaro, M. et al. Mammalian SWI/SNF continuously restores local accessibility to chromatin. Nat. Genet. 53, 279–287 (2021).

    Article  CAS  Google Scholar 

  14. Kelso, T. Chromatin accessibility underlies synthetic lethality of SWI/SNF subunits in ARID1A-mutant cancers. eLife 6, e30506 (2017).

    Article  Google Scholar 

  15. Yu, Y. et al. Olig2 targets chromatin remodelers to enhancers to initiate oligodendrocyte differentiation. Cell 152, 248–261 (2013).

    Article  CAS  Google Scholar 

  16. Ho, L. et al. An embryonic stem cell chromatin remodeling complex, esBAF, is an essential component of the core pluripotency transcriptional network. Proc. Natl Acad. Sci. USA 106, 5187–5191 (2009).

    Article  CAS  Google Scholar 

  17. Wang, X. et al. SMARCB1-mediated SWI/SNF complex function is essential for enhancer regulation. Nat. Genet. 49, 289–295 (2017).

    Article  CAS  Google Scholar 

  18. Alver, B. H. et al. The SWI/SNF chromatin remodelling complex is required for maintenance of lineage specific enhancers. Nat. Commun. 8, 14648 (2017).

    Article  Google Scholar 

  19. Nakayama, R. T. et al. SMARCB1 is required for widespread BAF complex-mediated activation of enhancers and bivalent promoters. Nat. Genet. 49, 1613–1623 (2017).

    Article  CAS  Google Scholar 

  20. Weintraub, A. S. et al. YY1 is a structural regulator of enhancer-promoter loops. Cell 171, 1573–1588 (2017).

    Article  CAS  Google Scholar 

  21. Di Giammartino, D. C. et al. KLF4 is involved in the organization and regulation of pluripotency-associated three-dimensional enhancer networks. Nat. Cell Biol. 21, 1179–1190 (2019).

    Article  Google Scholar 

  22. Dekker, J. & Misteli, T. Long-range chromatin interactions. Cold Spring Harb. Perspect. Biol. 7, a019356 (2015).

    Article  Google Scholar 

  23. Tang, Z. et al. CTCF-mediated human 3D genome architecture reveals chromatin topology for transcription. Cell 163, 1611–1627 (2015).

    Article  CAS  Google Scholar 

  24. Vierbuchen, T. et al. AP-1 transcription factors and the BAF complex mediate signal-dependent enhancer selection. Mol. Cell 68, 1067–1082 (2017).

    Article  CAS  Google Scholar 

  25. Branon, T. C. et al. Efficient proximity labeling in living cells and organisms with TurboID. Nat. Biotechnol. 36, 880–898 (2018).

    Article  CAS  Google Scholar 

  26. Soutourina, J. Transcription regulation by the Mediator complex. Nat. Rev. Mol. Cell Biol. 19, 262–274 (2018).

    Article  CAS  Google Scholar 

  27. Cenik, B. K. & Shilatifard, A. COMPASS and SWI/SNF complexes in development and disease. Nat. Rev. Genet. 22, 38–58 (2021).

    Article  CAS  Google Scholar 

  28. Bracken, A. P., Brien, G. L. & Verrijzer, C. P. Dangerous liaisons: interplay between SWI/SNF, NURD, and polycomb in chromatin regulation and cancer. Genes Dev. 33, 936–959 (2019).

    Article  CAS  Google Scholar 

  29. Park, J.-H. et al. Mammalian SWI/SNF complexes facilitate DNA double-strand break repair by promoting γ-H2AX induction. EMBO J. 25, 3986–3997 (2006).

    Article  CAS  Google Scholar 

  30. Brookes, E. & Shi, Y. Diverse epigenetic mechanisms of human disease. Annu. Rev. Genet. 48, 237–268 (2014).

    Article  CAS  Google Scholar 

  31. Skene, P. J. & Henikoff, S. An efficient targeted nuclease strategy for high-resolution mapping of DNA binding sites. eLife 6, e21856 (2017).

    Article  Google Scholar 

  32. Yu, G., Wang, L. G. & He, Q. Y. ChIP seeker: an R/Bioconductor package for ChIP peak annotation, comparison and visualization. Bioinformatics 31, 2382–2383 (2015).

    Article  CAS  Google Scholar 

  33. Kaczynski, J., Cook, T. & Urrutia, R. Sp1- and Krüppel-like transcription factors. Genome Biol. 4, 206 (2003).

    Article  Google Scholar 

  34. Hollenhorst, P. C., Shah, A. A., Hopkins, C. & Graves, B. J. Genome-wide analyses reveal properties of redundant and specific promoter occupancy within the ETS gene family. Genes Dev. 21, 1882–1894 (2007).

    Article  CAS  Google Scholar 

  35. Gao, F. et al. Heterozygous mutations in SMARCA2 reprogram the enhancer landscape by global retargeting of SMARCA4. Mol. Cell 75, 891–904 (2019).

    Article  CAS  Google Scholar 

  36. Herdegen, T. et al. Lasting N-terminal phosphorylation of c-Jun and activation of c-Jun N-terminal kinases after neuronal injury. J. Neurosci. 18, 5124–5135 (1998).

    Article  CAS  Google Scholar 

  37. Shaulian, E. & Karin, M. AP-1 as a regulator of cell life and death. Nat. Cell Biol. 4, E131–E136 (2002).

    Article  CAS  Google Scholar 

  38. Orlando, K. A. et al. Re-expression of SMARCA4/BRG1 in small cell carcinoma of ovary, hypercalcemic type (SCCOHT) promotes an epithelial-like gene signature through an AP-1-dependent mechanism. eLife 9, e59073 (2020).

    Article  CAS  Google Scholar 

  39. Meers, M. P., Janssens, D. H. & Henikoff, S. Pioneer factor-nucleosome binding events during differentiation are motif encoded. Mol. Cell 75, 562–575 (2019).

    Article  CAS  Google Scholar 

  40. Mumbach, M. R. et al. HiChIP: efficient and sensitive analysis of protein-directed genome architecture. Nat. Methods 13, 919–922 (2016).

    Article  CAS  Google Scholar 

  41. Mumbach, M. R. et al. Enhancer connectome in primary human cells identifies target genes of disease-associated DNA elements. Nat. Genet. 49, 1602–1612 (2017).

    Article  CAS  Google Scholar 

  42. Fang, R. et al. Mapping of long-range chromatin interactions by proximity ligation-assisted ChIP-seq. Cell Res. 26, 1345–1348 (2016).

    Article  CAS  Google Scholar 

  43. Servant, N. et al. HiC-Pro: an optimized and flexible pipeline for Hi-C data processing. Genome Biol. 16, 259 (2015).

    Article  Google Scholar 

  44. Lareau, C. A. & Aryee, M. J. hichipper: a preprocessing pipeline for calling DNA loops from HiChIP data. Nat. Methods 15, 155–156 (2018).

    Article  CAS  Google Scholar 

  45. Cho, K. F. et al. Proximity labeling in mammalian cells with TurboID and split-TurboID. Nat. Protoc. 15, 3971–3999 (2020).

    Article  CAS  Google Scholar 

  46. Kettenbach, A. N. et al. Global assessment of its network dynamics reveals that the kinase Plk1 inhibits the phosphatase PP6 to promote Aurora A activity. Sci. Signal. 11, eaaq1441 (2018).

    Article  Google Scholar 

  47. Eng, J. K., Jahan, T. A. & Hoopmann, M. R. Comet: an open-source MS/MS sequence database search tool. Proteomics 13, 22–24 (2013).

    Article  CAS  Google Scholar 

  48. Elias, J. E. & Gygi, S. P. Target-decoy search strategy for mass spectrometry-based proteomics. Methods Mol. Biol. 604, 55–71 (2010).

    Article  CAS  Google Scholar 

  49. Paek, J. et al. Multidimensional tracking of GPCR signaling via peroxidase-catalyzed proximity labeling. Cell 169, 338–349 (2017).

    Article  CAS  Google Scholar 

  50. Skene, P. J., Henikoff, J. G. & Henikoff, S. Targeted in situ genome-wide profiling with high efficiency for low cell numbers. Nat. Protoc. 13, 1006–1019 (2018).

    Article  CAS  Google Scholar 

  51. Meers, M. P., Bryson, T. D., Henikoff, J. G. & Henikoff, S. Improved CUT&RUN chromatin profiling tools. eLife 8, e46314 (2019).

    Article  Google Scholar 

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

    Article  CAS  Google Scholar 

  53. Li, H. et al. The Sequence alignment/map format and SAMtools. Bioinformatics 25, 2078–2079 (2009).

    Article  Google Scholar 

  54. Amemiya, H. M., Kundaje, A. & Boyle, A. P. The ENCODE blacklist: identification of problematic regions of the genome. Sci. Rep. 9, 9354 (2019).

    Article  Google Scholar 

  55. Zhang, Y. et al. Model-based analysis of ChIP-Seq (MACS). Genome Biol. 9, R137 (2008).

    Article  Google Scholar 

  56. Stovner, E. B. & Sætrom, P. Epic2 efficiently finds diffuse domains in ChIP-seq data. Bioinformatics 35, 4392–4393 (2019).

    Article  CAS  Google Scholar 

  57. Zang, C. et al. A clustering approach for identification of enriched domains from histone modification ChIP-Seq data. Bioinformatics 25, 1952–1958 (2009).

    Article  CAS  Google Scholar 

  58. Lareau, C. A. & Aryee, M. J. diffloop: a computational framework for identifying and analyzing differential DNA loops from sequencing data. Bioinformatics 34, 672–674 (2018).

    Article  CAS  Google Scholar 

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

    Article  CAS  Google Scholar 

  60. Ramírez, F., Dündar, F., Diehl, S., Grüning, B. A. & Manke, T. deepTools: a flexible platform for exploring deep-sequencing data. Nucleic Acids Res. 42, W187–W191 (2014).

    Article  Google Scholar 

  61. 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  Google Scholar 

  62. Kent, W. J. et al. The human genome browser at UCSC. Genome Res. 12, 996–1006 (2002).

    Article  CAS  Google Scholar 

Download references

Acknowledgements

We thank all members of the Wang laboratory for insightful and helpful discussions. We thank S. Henikoff (Fred Hutchinson Cancer Research Center) for providing the pA/G-MNase fusion protein, M. Meers (Fred Hutchinson Cancer Research Center) for helping with EChO analysis, and X. Zhang (Huntsman Cancer Institute) for helping with HiChIP analysis. This work was supported by the US National Institutes of Health (grants nos. R00CA197640 (X.W.), R01CA259850 (X.W.) and R01GM122846 (S.A.G.)), the Cancer Prevention Research Institute of Texas (CPRIT, RR180061 to C.C.) and a Dartmouth College Norris Cotton Cancer Center Support Grant (P30CA023108).The Andrew McDonough B+ (Be Positive) Foundation provided additional support for X.W. C.C. is a CPRIT Scholar in Cancer Research.

Author information

Authors and Affiliations

Authors

Contributions

B.K.W. and X.W. conceived the experiments and study design. A.M. and W.H.H. designed and performed the TurboID experiments. B.W.K., A.M. and X.W. performed the CUT&RUN, ATAC-seq, RNA-seq and HiChIP experiments. L.T.D., I.S.L. and S.A.G. helped with the TurboID experiments. B.K.W. and Y.Z. performed computational analyses. N.W.S. analyzed the RNA-seq data. C.C. supervised the computational analysis. All authors contributed to the interpretation of experiments. B.K.W., A.M. and X.W. wrote the manuscript, with input from all co-authors.

Corresponding authors

Correspondence to Chao Cheng or Xiaofeng Wang.

Ethics declarations

Competing interests

B.K.W. consults for SQZ Biotechnologies. The other authors declare no competing interests.

Peer review

Peer review information

Nature Structural & Molecular Biology thanks the anonymous reviewers for their contribution to the peer review of this work. Primary Handling editor: Carolina Perdigoto, in collaboration with the Nature Structural & Molecular Biology team. 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 Co-immunoprecipitation of SMARCC1 with indicated TurboID fusion subunits.

Co-immunoprecipitation of SMARCC1 with indicated TurboID fusion subunits and endogenous subunits showing how fusion of TurboID could impair subunit interaction with core subunit SMARCC1, indicating impaired ability to form intact complex.

Source data

Extended Data Fig. 2 SWI/SNF interactome via proximity-based labeling.

a. Glycerol sedimentation (10–35%) assay of major SWI/SNF core subunits (SMARCC1, SMARCC2, SMARCA4, SMARCD1, ACTL6A), subcomplex specific subunits (BAF: ARID1A, ARID1B, DPF2; PBAF: ARID2, PBRM1, PHF10; GBAF/ncBAF: BRD9, GLTSCR1) in G401 cells without SMARCB1 (Day 0) and SMARCB1 re-expression at Day 1, 3, and 7, showing rapid assembly of BAF and PBAF complexes and shift of subunits from GBAF to BAF and PBAF. b. Glycerol sedimentation assay of G401 SMARCD1-miniTurbo cells with and without SMARCB1, demonstrating SMARCD1-miniTurboID migrates similarly to endogenous SMARCD1. c. Western blot of Streptavidin IP of SMARCD1-TurboID and SMARCD1-miniTurboID nuclear lysate after different timepoints of exogenous biotin treatment. d. Volcano plot of TMT values for peptides labeled by SMARCD1-TurboID in G401 cells (Day 3 of SMARCB1 addback) compared to control NLS-TurboID G401 cells. e. Volcano plot of TMT values for peptides labeled by SMARCD1-TurboID in G401 cells (Day 3 of SMARCB1 addback) compared to SMARCD1-miniTurboID. T test function was used with each group’s protein expression data as input, and p values were generated by two-sided tests, no adjustment p values were used.

Source data

Extended Data Fig. 3 Dynamics of SWI/SNF subcomplexes targeting.

a. CUT&RUN peak numbers of SWI/SNF subunits: SMARCC1, ARID1A, PBRM1, and BRD9 in G401 cells without SMARCB1 (Day 0) and with SMARCB1 addback time points (Day 1, 3, and 7). b. Venn diagram comparison of CUT&RUN peak numbers of SWI/SNF subunits: SMARCC1, ARID1A, PBRM1, and BRD9 in G401 cells without SMARCB1 (Day 0) and with SMARCB1 addback time points (Day 1, 3, and 7). c. Genomic features of ARID1A regions overlapped with PBRM1, BRD9, and SMARCC1 (Top left), SMARCC1 regions overlapped with PBR1, BRD9, and ARID1A (Top right), BRD9 regions overlapped with PBRM1, ARID1A, and SMARCC1 (Bottom left), PBRM1 regions overlapped with ARID1A, BRD9, and SMARCC1 (Bottom right). d. Heatmap of HOMER identified motifs enriched in ARID1A, BRD9, PBRM1, and SMARCC1 CUT&RUN peaks over a time-course of SMARCB1 addback. T test function was used and p values were generated by two-sided tests, no adjustment p values were used. e. Heatmap of the percentage of peaks containing the motifs identified in part D for each peak set.

Source data

Extended Data Fig. 4 Dynamics of SWI/SNF subcomplexes targeting.

a. Heatmaps of GBAF specific subunit BRD9 with histone marks aligned to all BRD9 peaks in G401 cells. TSS-proximal is defined ± 2 Kb of TSS, TSS-distal is all regions outside 2 Kb of TSS. b. Heatmaps of PBAF specific subunit PBRM1 with histone marks aligned to all PBRM1 peaks in G401 cells. TSS-proximal is defined ±2 Kb of TSS, TSS-distal is all regions outside 2 Kb of TSS. c. Heatmaps of profiled SWI/SNF subunits and histone marks aligned to all SMARCC1 peaks. TSS-proximal is defined ±2 Kb of TSS, TSS-distal is all regions outside 2 Kb of TSS. d. Number of differential K27ac / K4me1 peaks over our time course (by DiffBind) at ARID1A, PBRM1, and BRD9 bound regions. The respective histone’s time point was compared against Day 0, specifically at peak regions contained by the respective SWI/SNF subunit. Differentially gained regions are defined as regions with an FDR < = 0.05, and log2FC > 0. e. Additional examples of SWI/SNF subunits targeting and histone marks using WashU Epigenome Browser, showing gained ARID1A, and/or PBRM1, and/or BRD9 binding upon SMARCB1 addback co-occurring with gained enhancer mark K27ac and K4me1.

Extended Data Fig. 5 SWI/SNF complex interacts with AP-1 complex and colocalizes on genome.

a. Western blot of G401 cells at Day 0 and 3 of SMARCB1 addback Immuno-Precipitated with AP-1 family proteins. b. Venn diagram comparison of CUT&RUN peak numbers of AP-1 subunits pJUN, JUND, and JUNB with ARID1A, PBRM1, and BRD9 without SMARCB1 (Day 0) and SMARCB1 re-expression (Day 1 and Day 3). c. Heat maps of SWI/SNF subunits ARID1A, PBRM1, BRD9, SMARCC1, and AP-1 subunits pJUN and JUNB aligned to all JUND peaks in G401 cells. TSS-proximal is defined ±2 Kb of TSS, TSS-distal is all regions outside 2 Kb of TSS. d. Heat maps of SWI/SNF subunits ARID1A, PBRM1, BRD9, SMARCC1, and AP-1 subunits pJUN and JUNB aligned to all JUNB peaks in G401 cells. TSS-proximal is defined ±2 Kb of TSS, TSS-distal is all regions outside 2 Kb of TSS. e. Average profile of SMARCC1 and ARID1A CUT&RUN signal before and after SMARCB1 addback following AP-1 depletion compared to a non-targeting control.

Extended Data Fig. 6 Cooperative targeting between SWI/SNF complex and AP-1 complex.

a. Proportion of fragments >120 bp in ARID1A, BRD9, PBRM1, and SMARCC1 CUT&RUN samples at indicated time points demonstrating the remodeling capacity of ARID1A-BAF complex (related to Fig. 5c). b. Proportion of fragments >120 bp in pJUN, JUND, and JUNB CUT&RUN samples at indicated time points (related to Fig. 5d). c. ChIPseeker genomic feature analysis of ATAC-seq peaks identified in G401 cells at each timepoint of SMARCB1 addback. d. HOMER Motif enrichment for ATAC-seq peaks identified in G401 cells at each timepoint of SMARCB1 addback. e. Specific examples using the WashU Epigenome browser showing regions that gain SWI/SNF and AP-1 binding after SMARCB1 addback, resulting in increased chromatin accessibility.

Source data

Extended Data Fig. 7 SWI/SNF complex mediated 3-D enhancer landscapes.

a. Summary of loops identified in K27ac and ARID1A HiChIP experiments. b. Additional example of SWI/SNF subunits SMARCC1, ARID1A, PBRM1, and BRD9 binding positions (CUT&RUN), and K27ac and ARID1A HiChIP using WashU Epigenome Browser, highlighting colocalization of SWI/SNF complex enrichment at enhancer anchors as well as overlapping loops between K27ac and ARID1A HiChIP (Left); Shown in the volcano plots highlighting genes in this region (CEBPB, KCNB1, PREX1, NCOA3) that are upregulated (Right). P values were generated by two-sided tests. c-j. Density heat-map showing correlation of altered K27ac loops with various factors’ CUT&RUN signal change at loop anchors: correlation of K27ac loop log fold change compared to BRD9 CUT&RUN signal log2 fold change (Day3/Day0) at corresponding anchor positions (C); correlation of K27ac loop log fold change compared to BRD9 CUT&RUN signal log2 fold change (Day3/Day0) at corresponding anchor positions that had static K27ac CUT&RUN signal across time (−1 < log2FC < 1) (D); correlation of K27ac loop log fold change compared to PBRM1 CUT&RUN signal log2 fold change (Day3/Day0) at corresponding anchor positions (E); correlation of K27ac loop log fold change compared to PBRM1 CUT&RUN signal log2 fold change (Day3/Day0) at corresponding anchor positions that had static K27ac CUT&RUN signal across time (−1 < log2FC < 1) (F); correlation of K27ac loop log fold change compared to JUNB CUT&RUN signal log2 fold change (Day3/Day0) at corresponding anchor positions (G); correlation of K27ac loop log fold change compared to JUNB CUT&RUN signal log2 fold change (Day3/Day0) at corresponding anchor positions that had static K27ac CUT&RUN signal across time (−1 < log2FC < 1) (H); correlation of K27ac loop log fold change compared to JUND CUT&RUN signal log2 fold change (Day3/Day0) at corresponding anchor positions (I); correlation of K27ac loop log fold change compared to JUND CUT&RUN signal log2 fold change (Day3/Day0) at corresponding anchor positions that had static K27ac CUT&RUN signal across time (−1 < log2FC < 1) (J). k. Differential K27ac HiChIP Enhancer-Promoter, Enhancer-Enhancer, and Promoter-Promoter loops upon AP-1 depletion.

Source data

Supplementary information

Reporting Summary

Peer Review File

Supplementary Table 1

Supplementary table for raw mass spectrometry data.

Source data

Source Data Fig. 1

Unprocessed western blots for Fig. 1.

Source Data Fig. 2

Statistical source data for Fig. 2a.

Source Data Fig. 3

Unprocessed western blots for Fig. 3.

Source Data Fig. 5

Statistical source data for Fig. 5b-d.

Source Data Fig. 6

Statistical source data for Fig. 6c,k.

Source Data Extended Data Fig. 1

Unprocessed western blots for Extended Data Fig. 1.

Source Data Extended Data Fig. 2

Unprocessed western blots for Extended Data Fig. 2.

Source Data Extended Data Fig. 3

Statistical source data for Extended Data Fig. 3c.

Source Data Extended Data Fig. 6

Statistical source data for Extended Data Fig. 6a-c.

Source Data Extended Data Fig. 7

Statistical source data for Extended Data Fig. 7a,k.

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

Wolf, B.K., Zhao, Y., McCray, A. et al. Cooperation of chromatin remodeling SWI/SNF complex and pioneer factor AP-1 shapes 3D enhancer landscapes. Nat Struct Mol Biol 30, 10–21 (2023). https://doi.org/10.1038/s41594-022-00880-x

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41594-022-00880-x

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