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:

CRISPR–ChIP reveals selective regulation of H3K79me2 by Menin in MLL leukemia

Abstract

Chromatin regulation involves the selective recruitment of chromatin factors to facilitate DNA repair, replication and transcription. Here we demonstrate the utility of coupling unbiased functional genomics with chromatin immunoprecipitation (CRISPR–ChIP) to identify the factors associated with active chromatin modifications in mammalian cells. Specifically, an integrated reporter containing a cis-regulatory element of interest and a single guide RNA provide a chromatinized template for a direct readout for regulators of histone modifications associated with actively transcribed genes such as H3K4me3 and H3K79me2. With CRISPR–ChIP, we identify all the nonredundant COMPASS complex members required for H3K4me3 and demonstrate that RNA polymerase II is dispensable for the maintenance of H3K4me3. As H3K79me2 has a putative oncogenic function in leukemia cells driven by MLL translocations, using CRISPR–ChIP we reveal a functional partitioning of H3K79 methylation into two distinct regulatory units: an oncogenic DOT1L complex directed by the MLL fusion protein in a Menin-dependent manner and a separate endogenous DOT1L complex, where catalytic activity is directed by MLLT10. Overall, CRISPR–ChIP provides a powerful tool for the unbiased interrogation of the mechanisms underpinning chromatin regulation.

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: Establishing the CRISPR–ChIP method.
Fig. 2: Characterizing the regulation of H3K4me3 using CRISPR–ChIP.
Fig. 3: MLLT10 and global H3K79me2 are dispensable for the survival of MLL-fusion leukemia cells.
Fig. 4: Selective regulation of transcription by MLL–AF9.
Fig. 5: Selective regulation of H3K79me2 at target genes by MLL–AF9.
Fig. 6: DOT1L functions in distinct native and oncogenic complexes in MLL-FP leukemia.
Fig. 7: Disruption of the native DOT1L complex sensitizes MLL–AF9 leukemia cells to Menin inhibition.
Fig. 8: Schematic model of the native and oncogenic DOT1L complexes.

Similar content being viewed by others

Data availability

Sequencing data that support the findings of this study have been deposited into the sequence read archive Gene Expression Omnibus, hosted by the National Center for Biotechnology Information. The accession number for the sequencing data reported in this paper is National Center for Biotechnology Information sequence read archive GSE192562. Source data are provided with this paper. All other data supporting the findings of this study are available from the corresponding authors on reasonable request.

References

  1. Przybyla, L. & Gilbert, L. A. A new era in functional genomics screens. Nat. Rev. Genet. 23, 89–103 (2021).

  2. Shilatifard, A. The COMPASS family of histone H3K4 methylases: mechanisms of regulation in development and disease pathogenesis. Annu. Rev. Biochem. 81, 65–95 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

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

    Article  CAS  PubMed  Google Scholar 

  4. Hu, D. et al. The Mll2 branch of the COMPASS family regulates bivalent promoters in mouse embryonic stem cells. Nat. Struct. Mol. Biol. 20, 1093–1097 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  5. Denissov, S. et al. Mll2 is required for H3K4 trimethylation on bivalent promoters in embryonic stem cells, whereas Mll1 is redundant. Development 141, 526–537 (2014).

    Article  CAS  PubMed  Google Scholar 

  6. Nguyen, A. T. & Zhang, Y. The diverse functions of Dot1 and H3K79 methylation. Genes Dev. 25, 1345–1358 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  7. Bernt, K. M. et al. MLL-rearranged leukemia is dependent on aberrant H3K79 methylation by DOT1L. Cancer Cell 20, 66–78 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  8. Krivtsov, A. V. et al. A Menin-MLL inhibitor induces specific chromatin changes and eradicates disease in models of MLL-rearranged leukemia. Cancer Cell 36, 660–673.e11 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  9. Daigle, S. R. et al. Selective killing of mixed lineage leukemia cells by a potent small-molecule DOT1L inhibitor. Cancer Cell 20, 53–65 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  10. Issa, G. C. et al. The menin inhibitor revumenib in KMT2A-rearranged or NPM1-mutant leukaemia. Nature 615, 920–924 (2023).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  11. Mohan, M. et al. Linking H3K79 trimethylation to Wnt signaling through a novel Dot1-containing complex (DotCom). Genes Dev. 24, 574–589 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  12. Gilan, O. et al. Functional interdependence of BRD4 and DOT1L in MLL leukemia. Nat. Struct. Mol. Biol. 23, 673–681 (2016).

    Article  CAS  PubMed  Google Scholar 

  13. Nelson, J. D., Denisenko, O. & Bomsztyk, K. Protocol for the fast chromatin immunoprecipitation (ChIP) method. Nat. Protoc. 1, 179–185 (2006).

    Article  CAS  PubMed  Google Scholar 

  14. Qin, J. Y. et al. Systematic comparison of constitutive promoters and the doxycycline-inducible promoter. PLoS ONE 5, e10611 (2010).

    Article  PubMed  PubMed Central  Google Scholar 

  15. Lasko, L. M. et al. Discovery of a selective catalytic p300/CBP inhibitor that targets lineage-specific tumours. Nature 550, 128–132 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  16. Santos-Rosa, H. et al. Active genes are tri-methylated at K4 of histone H3. Nature 419, 407–411 (2002).

    Article  CAS  PubMed  Google Scholar 

  17. Ernst, P. & Vakoc, C. R. WRAD: enabler of the SET1-family of H3K4 methyltransferases. Brief. Funct. Genomics 11, 217–226 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  18. Piunti, A. & Shilatifard, A. Epigenetic balance of gene expression by Polycomb and COMPASS families. Science 352, aad9780 (2016).

    Article  PubMed  Google Scholar 

  19. Krogan, N. J. et al. The Paf1 complex is required for histone H3 methylation by COMPASS and Dot1p: linking transcriptional elongation to histone methylation. Mol. Cell 11, 721–729 (2003).

    Article  CAS  PubMed  Google Scholar 

  20. Wu, M. et al. Molecular regulation of H3K4 trimethylation by Wdr82, a component of human Set1/COMPASS. Mol. Cell. Biol. 28, 7337–7344 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  21. Lee, J. H. & Skalnik, D. G. Wdr82 is a C-terminal domain-binding protein that recruits the Setd1A Histone H3-Lys4 methyltransferase complex to transcription start sites of transcribed human genes. Mol. Cell. Biol. 28, 609–618 (2008).

    Article  CAS  PubMed  Google Scholar 

  22. Schneider, J. et al. Molecular regulation of histone H3 trimethylation by COMPASS and the regulation of gene expression. Mol. Cell 19, 849–856 (2005).

    Article  CAS  PubMed  Google Scholar 

  23. Howe, F. S., Fischl, H., Murray, S. C. & Mellor, J. Is H3K4me3 instructive for transcription activation? Bioessays 39, 1–12 (2017).

    Article  CAS  PubMed  Google Scholar 

  24. Pérez-Lluch, S. et al. Absence of canonical marks of active chromatin in developmentally regulated genes. Nat. Genet. 47, 1158–1167 (2015).

    Article  PubMed  PubMed Central  Google Scholar 

  25. Henikoff, S. & Shilatifard, A. Histone modification: cause or cog? Trends Genet. 27, 389–396 (2011).

    Article  CAS  PubMed  Google Scholar 

  26. Nguyen, A. T., Taranova, O., He, J. & Zhang, Y. DOT1L, the H3K79 methyltransferase, is required for MLL–AF9-mediated leukemogenesis. Blood 117, 6912–6922 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  27. Okada, Y. et al. hDOT1L links histone methylation to leukemogenesis. Cell 121, 167–178 (2005).

    Article  CAS  PubMed  Google Scholar 

  28. Zeisig, D. T. et al. The eleven-nineteen-leukemia protein ENL connects nuclear MLL fusion partners with chromatin. Oncogene 24, 5525–5532 (2005).

    Article  CAS  PubMed  Google Scholar 

  29. Mueller, D. et al. A role for the MLL fusion partner ENL in transcriptional elongation and chromatin modification. Blood 110, 4445–4454 (2007).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  30. Wang, X., Chen, C. W. & Armstrong, S. A. The role of DOT1L in the maintenance of leukemia gene expression. Curr. Opin. Genet. Dev. 36, 68–72 (2016).

    Article  PubMed  Google Scholar 

  31. Wang, E. et al. Histone H2B ubiquitin ligase RNF20 is required for MLL-rearranged leukemia. Proc. Natl Acad. Sci. USA 110, 3901–3906 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  32. Briggs, S. D. et al. Gene silencing: trans-histone regulatory pathway in chromatin. Nature 418, 498 (2002).

    Article  CAS  PubMed  Google Scholar 

  33. Ng, H. H., Xu, R. M., Zhang, Y. & Struhl, K. Ubiquitination of histone H2B by Rad6 is required for efficient Dot1-mediated methylation of histone H3 lysine 79. J. Biol. Chem. 277, 34655–34657 (2002).

    Article  CAS  PubMed  Google Scholar 

  34. Wood, A. et al. Bre1, an E3 ubiquitin ligase required for recruitment and substrate selection of Rad6 at a promoter. Mol. Cell 11, 267–274 (2003).

    Article  CAS  PubMed  Google Scholar 

  35. Zhu, B. et al. The human PAF complex coordinates transcription with events downstream of RNA synthesis. Genes Dev. 19, 1668–1673 (2005).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  36. Vos, S. M., Farnung, L., Urlaub, H. & Cramer, P. Structure of paused transcription complex Pol II–DSIF–NELF. Nature 560, 601–606 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  37. Kamura, T. et al. The Elongin BC complex interacts with the conserved SOCS-box motif present in members of the SOCS, ras, WD-40 repeat, and ankyrin repeat families. Genes Dev. 12, 3872–3881 (1998).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  38. Pfleiderer, M. M. & Galej, W. P. Structure of the catalytic core of the Integrator complex. Mol. Cell 81, 1246–1259.e8 (2021).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  39. Elrod, N. D. et al. The integrator complex attenuates promoter-proximal transcription at protein-coding genes. Mol. Cell 76, 738–752.e7 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  40. Chen, F. X. et al. PAF1 regulation of promoter-proximal pause release via enhancer activation. Science 357, 1294–1298 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  41. Chen, S. et al. The PZP domain of AF10 senses unmodified H3K27 to regulate DOT1L-mediated methylation of H3K79. Mol. Cell 60, 319–327 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  42. Wan, L. et al. ENL links histone acetylation to oncogenic gene expression in acute myeloid leukaemia. Nature 543, 265–269 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  43. Erb, M. A. et al. Transcription control by the ENL YEATS domain in acute leukaemia. Nature 543, 270–274 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  44. 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 

  45. Natsume, T., Kiyomitsu, T., Saga, Y. & Kanemaki, M. T. Rapid protein depletion in human cells by auxin-inducible degron tagging with short homology donors. Cell Rep. 15, 210–218 (2016).

    Article  CAS  PubMed  Google Scholar 

  46. Olsen, S. N. et al. MLL::AF9 degradation induces rapid changes in transcriptional elongation and subsequent loss of an active chromatin landscape. Mol. Cell 82, 1140–1155.e11 (2022).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  47. Dawson, M. A. et al. Inhibition of BET recruitment to chromatin as an effective treatment for MLL-fusion leukaemia. Nature 478, 529–533 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  48. Zuber, J. et al. RNAi screen identifies Brd4 as a therapeutic target in acute myeloid leukaemia. Nature 478, 524–528 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  49. Winter, G. E. et al. BET bromodomain proteins function as master transcription elongation factors independent of CDK9 recruitment. Mol. Cell 67, 5–18.e19 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  50. Gilan, O. et al. Selective targeting of BD1 and BD2 of the BET proteins in cancer and immunoinflammation. Science 368, 387–394 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  51. Kuntimaddi, A. et al. Degree of recruitment of DOT1L to MLL–AF9 defines level of H3K79 di- and tri-methylation on target genes and transformation potential. Cell Rep. 11, 808–820 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  52. Deshpande, A. J. et al. AF10 regulates progressive H3K79 methylation and HOX gene expression in diverse AML subtypes. Cancer Cell 26, 896–908 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  53. Furey, T. S. ChIP–seq and beyond: new and improved methodologies to detect and characterize protein–DNA interactions. Nat. Rev. Genet. 13, 840–852 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  54. Sparbier, C. E. et al. Targeting Menin disrupts the KMT2A/B and polycomb balance to paradoxically activate bivalent genes. Nat. Cell Biol. 25, 258–272 (2023).

    CAS  PubMed  PubMed Central  Google Scholar 

  55. Dafflon, C. et al. Complementary activities of DOT1L and Menin inhibitors in MLL-rearranged leukemia. Leukemia 31, 1269–1277 (2017).

    Article  CAS  PubMed  Google Scholar 

  56. 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 

  57. Koike-Yusa, H., Li, Y., Tan, E. P., Velasco-Herrera, M. D. C. & Yusa, K. Genome-wide recessive genetic screening in mammalian cells with a lentiviral CRISPR-guide RNA library. Nat. Biotechnol. 32, 267–273 (2014).

    Article  CAS  PubMed  Google Scholar 

  58. FASTX-Toolkit (Cold Spring Harbor Laboratory, 2010); http://hannonlab.cshl.edu/fastx_toolkit/

  59. Martin, M. Cutadapt removes adapter sequences from high-throughput sequencing reads. EMBnet J. 17, 10–12 (2011).

    Article  Google Scholar 

  60. 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 

  61. Morgens, D. W. et al. Genome-scale measurement of off-target activity using Cas9 toxicity in high-throughput screens. Nat. Commun. 8, 15178 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  62. Wang, T., Wei, J. J., Sabatini, D. M. & Lander, E. S. Genetic screens in human cells using the CRISPR–Cas9 system. Science 343, 80–84 (2014).

    Article  CAS  PubMed  Google Scholar 

  63. Burr, M. L. et al. CMTM6 maintains the expression of PD-L1 and regulates anti-tumour immunity. Nature 549, 101–105 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  64. Aubrey, B. J. et al. An inducible lentiviral guide RNA platform enables the identification of tumor-essential genes and tumor-promoting mutations in vivo. Cell Rep. 10, 1422–1432 (2015).

    Article  CAS  PubMed  Google Scholar 

  65. Bell, C. C. et al. Targeting enhancer switching overcomes non-genetic drug resistance in acute myeloid leukaemia. Nat. Commun. 10, 2723 (2019).

  66. Shen, L., Shao, N., Liu, X. & Nestler, E. ngs.plot: quick mining and visualization of next-generation sequencing data by integrating genomic databases. BMC Genomics 15, 284 (2014).

    Article  PubMed  PubMed Central  Google Scholar 

  67. Garibaldi, A., Carranza, F. & Hertel, K. J. Isolation of newly transcribed RNA using the metabolic label 4-thiouridine. Methods Mol. Biol. 1648, 169–176 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  68. Kim, D., Paggi, J. M., Park, C., Bennett, C. & Salzberg, S. L. Graph-based genome alignment and genotyping with HISAT2 and HISAT-genotype. Nat. Biotechnol. 37, 907–915 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  69. Anders, S., Pyl, P. T. & Huber, W. HTSeq—a Python framework to work with high-throughput sequencing data. Bioinformatics 31, 166–169 (2014).

    Article  PubMed  PubMed Central  Google Scholar 

  70. 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 

Download references

Acknowledgements

We thank the Flow Cytometry facility and Molecular Genomics Core at the Peter MacCallum Cancer Centre and the ARAFlowcore Flow facility at the Australian Centre for Blood Diseases, Monash University. We thank the following funders for fellowship, scholarship and grant support: VCA Mid-Career Research Fellowship (O.G.); Cancer Council Victoria Sir Edward Dunlop Research Fellowship, NHMRC Investigator Grant #1196749 and Howard Hughes Medical Institute International Research Scholarship #55008729 (M.A.D.); and NHMRC Project Grants #1146192 (O.G.), #1085015 / #1106444 (M.A.D.) and #1128984 (M.A.D.). The funders had no role in study design, data collection and analysis, decision to publish or preparation of the manuscript.

Author information

Authors and Affiliations

Authors

Contributions

O.G. and M.A.D. conceived, designed and supervised the research and wrote the manuscript. O.G., C.C.B., D.N., D.T.F., K.K., M.B. and C.D. conducted experiments and/or analyzed data. O.G., C.C.B. and K.K. performed the CRISPR–ChIP screens. E.Y.N.L. and L.T. led the analysis of the genomic data and CRISPR–ChIP screens, with contribution from Y.-C.C.

Corresponding authors

Correspondence to Omer Gilan 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. The Dawson Laboratory is a recipient of grant funding through the emerging science fund administered through Pfizer. The remaining 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: Dimitris Typas, 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 Chromatin regulation of integrated regulatory elements.

a) ChIP-qPCR analysis of H3K27ac ChIP performed in parental K562 cells or K562 cells infected with the CRISPR-ChIP plasmid containing an EF1a promoter. Primers for a negative control region (Neg Ctrl), endogenous EF1a or regions spanning the gRNA/EF1a (Primer 1), and EF1a/Puro (Primer 2) within the lentiviral CRISPR-ChIP vector were used, data represents mean +/− SD from three independent biological replicates. b) ChIP-qPCR analysis of H3K27ac, H3K4me3, H3K9ac and RNA-Pol II ChIP at the endogenous PGK promoter or integrated PGK promoter-Puro reporter, an intergenic region not enriched in active histone marks was used as a negative control (Neg control). Data represents mean +/− SD from three independent biological replicates. c) ChIP-qPCR analysis of H3K27ac, H3K4me3, H3K79me2 and RNA-Pol II ChIP at the endogenous MEIS1 promoter or integrated MEIS1 promoter (~1.2Kb upstream of TSS)-Puro reporter, an intergenic region not enriched in active histone marks was used as a negative control (Neg control). Data represents the mean from two independent biological replicates. d) ChIP-qPCR analysis of H3K27ac, H3K4me3, and H3K9ac ChIP at the endogenous MYC enhancer or integrated MYC enhancer-mCMV-Puro sequence, an intergenic region not enriched in active histone marks was used as a negative control (Neg control). Data represents the mean from two independent biological replicates. e) ChIP-qPCR analysis of H3K79me2 in K562 cells infected with the CRISPR-ChIP plasmid and treated with DOT1L inhibitor (SGC0946) for 9 days. Primers for Negative control region (Neg Ctrl), Meis1, EF1a, and a region spanning the gRNA/EF1a were used, data represents mean +/− SD from three independent biological replicates. f) qPCR analysis of H3K27ac ChIP in K562 cells infected with the CRISPR-ChIP vector and treated with either DMSO or CBP/P300 inhibitor (A-485) for 24hrs. Primers for Negative control region (Neg Ctrl), Meis1, EF1a, and a region spanning the gRNA/EF1a were used, data represents the mean from two independent biological replicates. g) NGS representation of control sgRNAs at day 2, day 10, ChIP input day 10, and H3K79me2 ChIP (Left panel). NGS representation of DOT1L sgRNAs at day 2, day 10, ChIP input day 10, and H3K79me2 ChIP (Right panel).

Source data

Extended Data Fig. 2 Validation of the CRISPR-ChIP method.

a) The CRISPR-ChIP chromatin library was transduced into Cas9 negative K562 cells. H3K27ac ChIP was performed, and guide representation was assessed by NGS in two sampled replicates of library control and two independent ChIP samples. Data shown are correlation plots of library control Rep1 vs library control Rep2 (R=0.998), H3K27ac (50x10^6) Rep1 vs H3K27ac (50x10^6) Rep2 (R=0.993), and Library control Rep1 vs H3K27ac Rep1 (50x10^6) (R=0.985). b) Correlation plot of library guide counts between H3K27ac ChIP and a library control sample taken from Cas9 negative K562 cells. H3K27ac ChIP was performed from different starting cell number, 5 million, 10 million, 25 million, 50 million and 100 million cells. c) Bubble plot of H3K27ac ChIP (50M cells) from Cas9 negative K562 cells, P values calculated using the MAGECK algorithm and adjusted for multiple testing. d) Correlation plot of RBBP5 KO H3K4me3 LFC vs ASH2L KO H3K4me3 LFC. LFC=Log2FC.

Extended Data Fig. 3 Rationale for CRISPR-ChIP after knockout of essential genes.

a) sgRNA negative selection competition assay in K562 Cas9 cells transduced with control sgRNA or two independent sgRNAs against RNA Pol II (Pol II). Percentage of sgRNA positive cells remaining over time. Data represents the mean from n=2 experiments. b) FACS histogram of BFP expression in K562 Cas9 cells transduced with control sgRNA or two independent sgRNAs against RNA Pol II. Analysis performed at day 5 post infection. c) FACS scatter plot of forward scatter (FSC) vs side scatter (SSC) in K562 Cas9 cells transduced with control sgRNA or two independent sgRNAs against RNA Pol II. Analysis performed at day 5 post infection. d) Schematic of vector systems used for sgRNA expression and destabilised GFP (GFP-PEST) expression. e) FACS histogram of GFP expression in K562 Cas9 EF1a-GFP-PEST cells transduced with control sgRNA or two independent sgRNAs against RNA Pol II. Analysis performed at day 5 post infection. f) FACS scatter plot of GFP vs BFP expression in K562 Cas9 EF1a-GFP-PEST cells transduced with control sgRNA or two independent sgRNAs against the RNA Pol II. Analysis performed at day 5 post infection. g) Correlation plot of normalised Pol II levels (rpm) vs normalised H3K4me3 levels (rpm).

Source data

Extended Data Fig. 4 Transcriptional consequences of depletion of the DOT1L complex.

a) Schematic of DOTCOM complex (DOT1L, ENL/MLLT1, MLLT6/AF17, MLLT10). b) Guide counts over time from CRISPR dropout screen in MLL-AF9 cells for 2 MLLT10 and 2 DOT1L sgRNAs. c) PCA analysis of RNA-seq from MLLAF9 Cas9 cells transduced with control (non-targeting), DOT1L sg1, two independent ENL guides, two independent MLLT10 sgRNAs and two independent MLLT6 sgRNAs. d) Heatmap of RNA-seq data described above, showing downregulated genes in the various knockouts as indicated. e) Correlation plot of RNA-seq LFC between two independent MLLT6 sgRNAs and DOT1L KO for the top 100 downregulated genes in the DOT1L KO cells. LFC=Log2FC.

Extended Data Fig. 5 Establishment of a model for the rapid degradation of MLLAF9.

a) Kaplan-Meier curve of mice transplanted with MLLAF9FLAG-mAID cells or Nras + MLLAF9FLAG-mAID. b) Genome browser snapshot of MLLAF9 ChIP-seq replicates at the MEIS1 locus. FLAG ChIP from non-FLAG tagged MLLAF9 cells used as a negative control and input shown. c) Genome browser snapshot of MLLAF9 ChIP-seq replicates at the HOXA cluster. FLAG ChIP from non-FLAG tagged MLLAF9 cells used as a negative control and input shown. d) Proliferation assay of MLLAF9FLAG-mAID Tir1 cells treated with DMSO or IAA (500uM) over 7 days. Eror bars represent mean +/− SD from three independent biological replicates. e) FACS analysis of cell surface Gr1 levels in unstained, stained, DMSO or IAA treated MLLAF9FLAG-mAID Tir1 cells for 4 days. f) Profile plot of MLLAF9 FLAG ChIP-seq in MLLAF9FLAG-mAID cells treated with DMSO, IAA (2hrs and 4hrs). g) Heatmap of 4su-seq (nascent RNA-seq) in MLLAF9FLAG-mAID Tir1 cells treated with DMSO or IAA for 2hrs, showing 260 downregulated genes. h) PCA plot of RNA-seq from MLLAF9FLAG-mAID Tir1 cells treated with DMSO or IAA (2hrs). i) Heatmap of RNA Pol II ser2 ChIP-seq at all genes in MLLAF9FLAG-mAID Tir1 cells treated with DMSO or IAA (2hrs). j) Genome browser snapshots of MLLAF9 targets, EPHA7 and BAZ2B showing RNA-Pol IIS2ph and MLLAF9 ChIP-seq tracks after treatment with either DMSO or IAA (2hrs).

Source data

Extended Data Fig. 6 Menin and DOT1L inhibition evicts MLLAF9 and Menin from chromatin.

a) Heatmap of Menin ChIP-seq in MLLAF9 cells treated with DMSO, SGC0946 (72hrs) or VTP50469 (48hrs) at all genes. b) MLLAF9 FLAG ChIP-seq in MLLAF9 cells treated with DMSO, IBET151 or IBET-VHL for 8hrs at all genes. c) PCA analysis of RNA-seq data in MLLAF9 cells treated with DMSO, SGC0946 (72hrs), IAA (24hrs), VTP50469 (24hrs). d) Heatmap of H3K79me2 ChIP-seq analysis in MLLAF9-F3-mAID cells treated with either DMSO, SGC0946 (6hrs), SGC0946 (48hrs), VTP50469 (6hrs), and VTP50469 (48hrs). e) Genome browser snapshot of the MEIS1 locus from the H3K79me2 ChIP-seq in (D).

Extended Data Fig. 7 H3K79me2 is selectively lost after prolonged Menin inhibition.

a) Profile plot of MLLAF9 ChIP-seq in MLLAF9 cells treated with either DMSO, VTP50469 (6hrs), SGC0946 (48hrs) or VTP50469 (48hrs) b) Heatmap of MLLAF9 and H3K79me2 ChIP-seq data in MLLAF9 cells after treatment with VTIP50469 for 48hrs at genes that show decreased H3K79me2 after VTP50469 treatment. c) Genome browser snapshot of ChIP-seq tracks at canonical MLLAF9 target genes, JMJD1C and RUNX2, showing MLL-AF9 and H3K79me2 after treatment with VTP50469 or SGC0946 for 48hrs.

Extended Data Fig. 8 H3K79me2 is retained at MLLAF9 bound regions in MLLT10 KO cells.

a) Average profile plot of H3K79me2 and input at all genes in MLL-AF9 cells infected with non-silencing guides (Control), 2 independent DOT1L sgRNAs and 2 independent MLLT10 sgRNAs. b) Heatmap of H3K79me2 at direct MLL-AF9 target genes (from nascentRNA-seq after IAA treatment) for the indicated samples. Genes are ranked from highest to lowest coverage of H3K79me2 in the control cells. c) Genome browser snapshot of a canonical MLLAF9 target, HOXA9, from H3K79me2 ChIP-seq in MLLAF9 cells transduced with control sgRNA or two independent targeting MLLT10 or DOT1L. d) Correlation plot of LFC H3K79me2 in MLLT10 depleted vs control cells for sgRNA 1 vs sgRNA 2 (left panel) and sgRNA1 vs sgRNA 3 (right panel). e) Genome browser snapshots of exemplar direct and strongly bound MLL-AF9 target genes that show increased H3K79me2 following MLLT10 KO in MLL-AF9 cells, Jmjd1c and Epha7. f) Genome browser snapshot of exemplar direct MLLAF9 target genes (Meis1, Tsc22d2) in control or Cre deleted MLLT10 in murine MLLAF9 cells from Deshpande et al14. g) Immunoblot analysis of gel filtration fractions (superose 6 column) from nuclear extracts of MV4;11 cells (human MLL-AF4) using antibodies against DOT1L, MLLT10, MLL1, and ENL. Fractions from smallest to largest. Representative image from two independent experiments. h) Immunoblot analysis of gel filtration fractions (superose 6 column) from nuclear extracts of MOLM13 cells (human MLL-AF9) using antibodies against DOT1L, MLLT10, MLL1 and ENL. Fractions from smallest to largest. Representative image from two independent experiments.

Source data

Extended Data Fig. 9 MLLT10 depletion sensitises cells to Menin inhibition.

a) Heatmap of RNA-seq data from MLLAF9 Cas9 cells infected with control or MLLT10 sgRNAs and treated with either DMSO or VTP50469 with 20 or 40nM. b) Proliferation assays using MLL-AF9 Cas9 cells transduced with either a control non-silencing sgRNA or two independent sgRNAs targeting MLLT10 (sg1 and sg2). Cells were treated with either DMSO, VTP50469 25nM or 50nM. Plots represent n=3 biological replicates. Error bars represent mean +/− SD c) Proliferation assay in control and MLLT10 KO MLLAF9FLAG-mAID Cas9 cells treated with either DMSO, VTP50469 (10nM) or SGC0946 (1uM) + VTP50469 (10nM). Cell counts shown at day 3 and day 7. Data represents mean +/− SD from n=3 biological replicates. d) Profile plot of H3K79me2 ChIP-seq in MLLAF9 Cas9 cells infected with control or MLLT10 guides and treated with DMSO or VTP50469 (20nM) for 48hrs. Analysis performed across MLLAF9 target genes. e) Proliferation assay in MOLM13 cells treated with either DMSO, SGC0946 (3uM), VTP50469 (100nM) or combination. Mean +/− SD from n=3 biological replicates. f) Proliferation assay in MLLAF9FLAG-mAID Cas9 cells treated with either DMSO, SGC0946 (3uM), VTP50469 (50nM) or combination. Mean +/− SD from n=3 biological replicates.

Source data

Supplementary information

Reporting Summary

Peer Review File

Supplementary Tables

Supplementary Table 1: Human and mouse chromatin sgRNA library. Supplementary Table 2: CRISPR–ChIP H3K4me3 raw counts and MAGECK analysis. Supplementary Table 3: CRISPR–ChIP H3K79me2 raw counts and MAGECK analysis. Supplementary Table 4: Day 14–day 2 CRISPR screen dropout MAGECK analysis. Supplementary Table 5: ChIP–qPCR and primary transcript qPCR primers.

Source data

Source Data Fig. 1

Statistical source data.

Source Data Fig. 1

Unprocessed western blots.

Source Data Fig. 2

Unprocessed western blots.

Source Data Fig. 3

Statistical source data.

Source Data Fig. 3

Unprocessed western blots.

Source Data Fig. 4

Unprocessed western blots.

Source Data Fig. 5

Statistical source data.

Source Data Fig. 7

Statistical source data.

Source Data Extended Data Fig. 1

Statistical source data.

Source Data Extended Data Fig. 3

Statistical source data.

Source Data Extended Data Fig. 5

Statistical source data.

Source Data Extended Data Fig. 8

Unprocessed western blots.

Source Data Extended Data Fig. 9

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

Gilan, O., Talarmain, L., Bell, C.C. et al. CRISPR–ChIP reveals selective regulation of H3K79me2 by Menin in MLL leukemia. Nat Struct Mol Biol 30, 1592–1606 (2023). https://doi.org/10.1038/s41594-023-01087-4

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41594-023-01087-4

This article is cited by

Search

Quick links

Nature Briefing: Cancer

Sign up for the Nature Briefing: Cancer newsletter — what matters in cancer research, free to your inbox weekly.

Get what matters in cancer research, free to your inbox weekly. Sign up for Nature Briefing: Cancer