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:

Subtype-specific 3D genome alteration in acute myeloid leukaemia

Abstract

Acute myeloid leukaemia (AML) represents a set of heterogeneous myeloid malignancies, and hallmarks include mutations in epigenetic modifiers, transcription factors and kinases1,2,3,4,5. The extent to which mutations in AML drive alterations in chromatin 3D structure and contribute to myeloid transformation is unclear. Here we use Hi-C and whole-genome sequencing to analyse 25 samples from patients with AML and 7 samples from healthy donors. Recurrent and subtype-specific alterations in A/B compartments, topologically associating domains and chromatin loops were identified. RNA sequencing, ATAC with sequencing and CUT&Tag for CTCF, H3K27ac and H3K27me3 in the same AML samples also revealed extensive and recurrent AML-specific promoter–enhancer and promoter–silencer loops. We validated the role of repressive loops on their target genes by CRISPR deletion and interference. Structural variation-induced enhancer-hijacking and silencer-hijacking events were further identified in AML samples. Hijacked enhancers play a part in AML cell growth, as demonstrated by CRISPR screening, whereas hijacked silencers have a downregulating role, as evidenced by CRISPR-interference-mediated de-repression. Finally, whole-genome bisulfite sequencing of 20 AML and normal samples revealed the delicate relationship between DNA methylation, CTCF binding and 3D genome structure. Treatment of AML cells with a DNA hypomethylating agent and triple knockdown of DNMT1, DNMT3A and DNMT3B enabled the manipulation of DNA methylation to revert 3D genome organization and gene expression. Overall, this study provides a resource for leukaemia studies and highlights the role of repressive loops and hijacked cis elements in human diseases.

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: Genome organization and compartment analysis in primary AML samples.
Fig. 2: AML and subtype-specific chromatin loops.
Fig. 3: Identification and validation of repressive loops.
Fig. 4: Identification, characterization and screening of enhancer hijacking in AML.
Fig. 5: Identification and validation of silencer hijacking in AML.
Fig. 6: Inhibition of DNA methylation restores chromatin structure and gene expression.

Similar content being viewed by others

Data availability

The processed data from CUT&Tag, ATAC-seq, Hi-C and WGBS of primary samples and cell lines and the raw sequencing data for all cell lines have been deposited at Gene Expression Omnibus with the open session GSE152136. All processed data were generated by mapping to human reference genome GRCh38. Source data are provided with this paper.

Code availability

The code for stripe identification is available at GitHub with the following link: https://github.com/XiaoTaoWang/StripeCaller.

References

  1. Dohner, H. et al. Diagnosis and management of AML in adults: 2017 ELN recommendations from an international expert panel. Blood 129, 424–447 (2017).

    Article  PubMed  PubMed Central  Google Scholar 

  2. Dohner, H., Weisdorf, D. J. & Bloomfield, C. D. Acute myeloid leukemia. N. Engl. J. Med. 373, 1136–1152 (2015).

    Article  PubMed  Google Scholar 

  3. Arber, D. A. et al. The 2016 revision to the World Health Organization classification of myeloid neoplasms and acute leukemia. Blood 127, 2391–2405 (2016).

    Article  CAS  PubMed  Google Scholar 

  4. Papaemmanuil, E. et al. Genomic classification and prognosis in acute myeloid leukemia. N. Engl. J. Med. 374, 2209–2221 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  5. Assi, S. A. et al. Subtype-specific regulatory network rewiring in acute myeloid leukemia. Nat. Genet. 51, 151–162 (2019).

    Article  CAS  PubMed  Google Scholar 

  6. McKeown, M. R. et al. Superenhancer analysis defines novel epigenomic subtypes of non-APL AML, including an RARα dependency targetable by SY-1425, a potent and selective RARα agonist. Cancer Discov. 7, 1136–1153 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  7. Harris, W. J. et al. The histone demethylase KDM1A sustains the oncogenic potential of MLL-AF9 leukemia stem cells. Cancer Cell 21, 473–487 (2012).

    Article  CAS  PubMed  Google Scholar 

  8. Luo, H. et al. CTCF boundary remodels chromatin domain and drives aberrant HOX gene transcription in acute myeloid leukemia. Blood 132, 837–848 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  9. Ghasemi, R., Struthers, H., Wilson, E. R. & Spencer, D. H. Contribution of CTCF binding to transcriptional activity at the HOXA locus in NPM1-mutant AML cells. Leukemia 35, 404–416 (2020).

    Article  PubMed  PubMed Central  Google Scholar 

  10. Lieberman-Aiden, E. et al. Comprehensive mapping of long-range interactions reveals folding principles of the human genome. Science 326, 289–293 (2009).

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  11. Dixon, J. R. et al. Topological domains in mammalian genomes identified by analysis of chromatin interactions. Nature 485, 376–380 (2012).

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  12. Nora, E. P. et al. Spatial partitioning of the regulatory landscape of the X-inactivation centre. Nature 485, 381–385 (2012).

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  13. Yan, J. et al. Histone H3 lysine 4 monomethylation modulates long-range chromatin interactions at enhancers. Cell Res. 28, 204–220 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  14. Rao, S. S. P. et al. Cohesin loss eliminates all loop domains. Cell 171, 305–320.e24 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  15. Viny, A. D. et al. Cohesin members Stag1 and Stag2 display distinct roles in chromatin accessibility and topological control of HSC self-Renewal and differentiation. Cell Stem Cell 25, 682–696.e8 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  16. Yang, M. et al. Proteogenomics and Hi-C reveal transcriptional dysregulation in high hyperdiploid childhood acute lymphoblastic leukemia. Nat. Commun. 10, 1519 (2019).

    Article  ADS  PubMed  PubMed Central  Google Scholar 

  17. Diaz, N. et al. Chromatin conformation analysis of primary patient tissue using a low input Hi-C method. Nat. Commun. 9, 4938 (2018).

    Article  ADS  PubMed  PubMed Central  Google Scholar 

  18. Kloetgen, A. et al. Three-dimensional chromatin landscapes in T cell acute lymphoblastic leukemia. Nat. Genet. 52, 388–400 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  19. Yang, H. et al. Noncoding genetic variation in GATA3 increases acute lymphoblastic leukemia risk through local and global changes in chromatin conformation. Nat. Genet. 54, 170–179 (2022).

  20. Dixon, J. R. et al. Integrative detection and analysis of structural variation in cancer genomes. Nat. Genet. 50, 1388–1398 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  21. Groschel, S. et al. A single oncogenic enhancer rearrangement causes concomitant EVI1 and GATA2 deregulation in leukemia. Cell 157, 369–381 (2014).

    Article  CAS  PubMed  Google Scholar 

  22. Northcott, P. A. et al. Enhancer hijacking activates GFI1 family oncogenes in medulloblastoma. Nature 511, 428–434 (2014).

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  23. Spassov, B. V. et al. Wilms’ tumor protein and FLT3-internal tandem duplication expression in patients with de novo acute myeloid leukemia. Hematology 16, 37–42 (2011).

    Article  CAS  PubMed  Google Scholar 

  24. Sondka, Z. et al. The COSMIC Cancer Gene Census: describing genetic dysfunction across all human cancers. Nat. Rev. Cancer 18, 696–705 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  25. Metzeler, K. H. et al. Spectrum and prognostic relevance of driver gene mutations in acute myeloid leukemia. Blood 128, 686–698 (2016).

    Article  CAS  PubMed  Google Scholar 

  26. Akdemir, K. C. et al. Disruption of chromatin folding domains by somatic genomic rearrangements in human cancer. Nat. Genet. 52, 294–305 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  27. Salameh, T. J. et al. A supervised learning framework for chromatin loop detection in genome-wide contact maps. Nat. Commun. 11, 3428 (2020).

  28. Jaatinen, T. et al. Global gene expression profile of human cord blood-derived CD133+ cells. Stem Cells 24, 631–641 (2006).

    Article  CAS  PubMed  Google Scholar 

  29. Diaz-Blanco, E. et al. Molecular signature of CD34+ hematopoietic stem and progenitor cells of patients with CML in chronic phase. Leukemia 21, 494–504 (2007).

    Article  CAS  PubMed  Google Scholar 

  30. Ngan, C. Y. et al. Chromatin interaction analyses elucidate the roles of PRC2-bound silencers in mouse development. Nat. Genet. 52, 264–272 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  31. Cai, Y. et al. H3K27me3-rich genomic regions can function as silencers to repress gene expression via chromatin interactions. Nat. Commun. 12, 719 (2021).

  32. Kataoka, K. et al. Integrated molecular analysis of adult T cell leukemia/lymphoma. Nat. Genet. 47, 1304–1315 (2015).

    Article  CAS  PubMed  Google Scholar 

  33. Goldman, M. J. et al. Visualizing and interpreting cancer genomics data via the Xena platform. Nat. Biotechnol. 38, 675–678 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

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

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  35. Morris, V. A., Cummings, C. L., Korb, B., Boaglio, S. & Oehler, V. G. Deregulated KLF4 expression in myeloid leukemias alters cell proliferation and differentiation through microRNA and gene targets. Mol. Cell. Biol. 36, 559–573 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  36. Wang, X. et al. Genome-wide detection of enhancer-hijacking events from chromatin interaction data in rearranged genomes. Nat. Methods 18, 661–668 (2021).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  37. Phanstiel, D. H. et al. Static and dynamic DNA loops form AP-1-bound activation hubs during macrophage development. Mol. Cell 67, 1037–1048.e6 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  38. 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 

  39. Achinger-Kawecka, J. et al. Epigenetic reprogramming at estrogen-receptor binding sites alters 3D chromatin landscape in endocrine-resistant breast cancer. Nat. Commun. 11, 320 (2020).

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  40. Bell, A. C. & Felsenfeld, G. Methylation of a CTCF-dependent boundary controls imprinted expression of the Igf2 gene. Nature 405, 482–485 (2000).

    Article  ADS  CAS  PubMed  Google Scholar 

  41. Figueroa, M. E. et al. DNA methylation signatures identify biologically distinct subtypes in acute myeloid leukemia. Cancer Cell 17, 13–27 (2010).

    Article  MathSciNet  CAS  PubMed  PubMed Central  Google Scholar 

  42. Flavahan, W. A. et al. Altered chromosomal topology drives oncogenic programs in SDH-deficient GISTs. Nature 575, 229–233 (2019).

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  43. Flavahan, W. A. et al. Insulator dysfunction and oncogene activation in IDH mutant gliomas. Nature 529, 110–114 (2016).

    Article  ADS  CAS  PubMed  Google Scholar 

  44. Tovy, A. et al. Tissue-biased expansion of DNMT3A-mutant clones in a mosaic individual is associated with conserved epigenetic erosion. Cell Stem Cell 27, 326–335.e4 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  45. Zhang, X. et al. Large DNA methylation nadirs anchor chromatin loops maintaining hematopoietic stem cell identity. Mol. Cell 78, 506–521.e6 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  46. Akalin, A. et al. Base-pair resolution DNA methylation sequencing reveals profoundly divergent epigenetic landscapes in acute myeloid leukemia. PLoS Genet. 8, e1002781 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  47. Neri, F. et al. Intragenic DNA methylation prevents spurious transcription initiation. Nature 543, 72–77 (2017).

    Article  ADS  CAS  PubMed  Google Scholar 

  48. Sanborn, A. L. et al. Chromatin extrusion explains key features of loop and domain formation in wild-type and engineered genomes. Proc. Natl Acad. Sci. USA 112, E6456–E6465 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  49. Yang, T. et al. HiCRep: assessing the reproducibility of Hi-C data using a stratum-adjusted correlation coefficient. Genome Res. 27, 1939–1949 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  50. Spracklin, G. et al. Heterochromatin diversity modulates genome compartmentalization and loop extrusion barriers. Preprint at bioRxiv https://doi.org/10.1101/2021.08.05.455340 (2021).

  51. Cesana, D. et al. Retrieval of vector integration sites from cell-free DNA. Nat. Med. 27, 1458–1470 (2021).

    Article  MathSciNet  CAS  PubMed  Google Scholar 

  52. Loke, J. et al. RUNX1-ETO and RUNX1-EVI1 differentially reprogram the chromatin landscape in t(8;21) and t(3;21) AML. Cell Rep. 19, 1654–1668 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  53. Kaya-Okur, H. S. et al. CUT&Tag for efficient epigenomic profiling of small samples and single cells. Nat. Commun. 10, 1930 (2019).

    Article  ADS  PubMed  PubMed Central  Google Scholar 

  54. Buenrostro, J. D., Wu, B., Chang, H. Y. & Greenleaf, W. J. ATAC-seq: a method for assaying chromatin accessibility genome-wide. Curr. Protoc. Mol. Biol. 109, 21.29.1–21.29.9 (2015).

    Article  Google Scholar 

  55. Diao, Y. et al. A tiling-deletion-based genetic screen for cis-regulatory element identification in mammalian cells. Nat. Methods 14, 629–635 (2017).

    Article  CAS  PubMed  PubMed Central  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. Bolland, D. J., King, M. R., Reik, W., Corcoran, A. E. & Krueger, C. Robust 3D DNA FISH using directly labeled probes. J. Vis. Exp. https://doi.org/10.3791/50587 (2013).

  58. Riccardi, C., & Nicoletti, I. Analysis of apoptosis by propidium iodide staining and flow cytometry. Nat. Protoc. 1, 1458–1461 (2006).

    Article  CAS  PubMed  Google Scholar 

  59. Tarasov, A., Vilella, A. J., Cuppen, E., Nijman, I. J. & Prins, P. Sambamba: fast processing of NGS alignment formats. Bioinformatics 31, 2032–2034 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  60. Cingolani, P. et al. A program for annotating and predicting the effects of single nucleotide polymorphisms, SnpEff: SNPs in the genome of Drosophila melanogaster strain w1118; iso-2; iso-3. Fly (Austin) 6, 80–92 (2012).

    Article  CAS  Google Scholar 

  61. Rausch, T. et al. DELLY: structural variant discovery by integrated paired-end and split-read analysis. Bioinformatics 28, i333–i339 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  62. Chiang, C. et al. SpeedSeq: ultra-fast personal genome analysis and interpretation. Nat. Methods 12, 966–968 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  63. Abdennur, N. & Mirny, L. A. Cooler: scalable storage for Hi-C data and other genomically labeled arrays. Bioinformatics 36, 311–316 (2020).

    Article  CAS  PubMed  Google Scholar 

  64. Kerpedjiev, P. et al. HiGlass: web-based visual exploration and analysis of genome interaction maps. Genome Biol. 19, 125 (2018).

    Article  PubMed  PubMed Central  Google Scholar 

  65. Durand, N. C. et al. Juicebox provides a visualization system for Hi-C contact maps with unlimited zoom. Cell Syst. 3, 99–101 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  66. Wang, Y. et al. The 3D Genome Browser: a web-based browser for visualizing 3D genome organization and long-range chromatin interactions. Genome Biol. 19, 151 (2018).

    Article  PubMed  PubMed Central  Google Scholar 

  67. Dixon, J. R. et al. Chromatin architecture reorganization during stem cell differentiation. Nature 518, 331–336 (2015).

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  68. Varoquaux, N., Ay, F., Noble, W. S. & Vert, J. P. A statistical approach for inferring the 3D structure of the genome. Bioinformatics 30, i26–i33 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  69. 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 

  70. Krueger, F. & Andrews, S. R. Bismark: a flexible aligner and methylation caller for Bisulfite-Seq applications. Bioinformatics 27, 1571–1572 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  71. Schultz, M. D. et al. Human body epigenome maps reveal noncanonical DNA methylation variation. Nature 523, 212–216 (2015).

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  72. Corces, M. R. et al. The chromatin accessibility landscape of primary human cancers. Science 362, eaav1898 (2018).

    Article  ADS  PubMed  PubMed Central  Google Scholar 

  73. Dobin, A. et al. STAR: ultrafast universal RNA-seq aligner. Bioinformatics 29, 15–21 (2013).

    Article  CAS  PubMed  Google Scholar 

  74. Li, B. & Dewey, C. N. RSEM: accurate transcript quantification from RNA-seq data with or without a reference genome. BMC Bioinformatics 12, 323 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  75. Juric, I. et al. MAPS: model-based analysis of long-range chromatin interactions from PLAC-seq and HiChIP experiments. PLoS Comput. Biol. 15, e1006982 (2019).

    Article  PubMed  PubMed Central  Google Scholar 

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

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  77. Knight, P. A. & Ruiz, D. A fast algorithm for matrix balancing. IMA J. Numer. Anal. 33, 1029–1047 (2012).

    Article  MathSciNet  MATH  Google Scholar 

Download references

Acknowledgements

F.Y. is supported by NIH grants R35GM124820, 1R01HG009906, R01HG011207, U01CA200060 and R24DK106766 (R.C.H. and F.Y.). We acknowledge the use of the Integrated Genomics Operation Core, funded by the Memorial Sloan Kettering Cancer Center Support Grant NIH P30 CA008748. This work was supported by National Cancer Institute R35 CA197594-01A1 (R.L.L.), National Cancer Institute R01 CA216421 (R.L.L.) and National Cancer Institute PS-OC U54 CA143869-05 (R.L.L). A.D.V. is supported by National Cancer Institute career development grant K08 CA215317, the William Raveis Charitable Fund Fellowship of the Damon Runyon Cancer Research Foundation (DRG 117-15) and an Evans MDS Young Investigator grant from the Edward P. Evans Foundation. T.Y. is supported by U01DA053691. The CUT&Tag reagent pA–Tn5 was provided as a gift from S. Henikoff’s Lab at Fred Hutchinson Cancer Research Center. The dCas9-UTX plasmid was provided as a gift by S. M. Offer’s Lab at Mayo Clinic.

Author information

Authors and Affiliations

Authors

Contributions

F.Y. conceived and supervised the project. J.X. and F.S. led the investigation. J.X., H.L., B.Z. and M.K. performed Hi-C, ATAC-seq, CUT&Tag, ChIP-seq, RNA-seq and 5-AZA-related experiments, and prepared DNA for WGS and WGBS. J.X. and H.L. performed CRISPR screening. J.X. performed 4C. H.L. and J.X. performed CRISPRi. M.K. and L.S. performed CRIPSR deletion. Z.Z. performed DNA FISH. Y.Y. and T.Y. supervised DNA FISH analyses. Qi Jin helped with FACS. J.H.W. performed reporter assays. H.Y. and Qiushi Jin performed DNMT TKD and HiChIP. F.S., J.X., Y.L., Y.F. and Q.W. conducted data analysis. X.W. developed the algorithm for stripe detection. Y.D., S.D., R.C.H. and J.R.B. contributed biological insights. H.Z., B.J., D.C., J.R.B., R.L.L., A.D.V. and L.C.P. provided the samples and clinical insights. J.X., F.S. and F.Y. wrote the manuscript with input from all authors.

Corresponding authors

Correspondence to Hong Zheng or Feng Yue.

Ethics declarations

Competing interests

F.Y. is a co-founder of Sariant Therapeutics, Inc. The other authors declare no competing interests.

Peer review

Peer review information

Nature thanks the anonymous reviewers for their contribution to the peer review of this work.

Additional information

Publisher’s note Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Extended data figures and tables

Extended Data Fig. 1

List of genomic experiments performed in this study.

Extended Data Fig. 2 Examples of genes with A/B compartment switch.

a and c, PC1 values and the expressions for WT1, POU2AF1, FGF13, and BCL11B genes. b, A compartment associated with higher H3K27ac and ATAC-seq signals and lower H3K27me3 intensities. Shown in figure are the ATAC-seq, CUT&Tag for H3K27ac and H3K27me3 signals at WT1 gene promoter, normalized to sequencing depths. Promoter is defined as within +/-1Kb of TSS. P value by two-sided Wilcoxon rank-sum test. PC1: A (n = 15 samples) B (n = 7 samples); H3K27ac: A (n = 15) B (n = 6); H3K27me3: A (n = 15) B (n = 6); ATAC-seq: A (n = 14) B (n = 7). Box plot: middle line denotes the median, top/bottom of boxes denotes first/third quartiles and whiskers extend to 1.5 times the interquartile range.

Source Data

Extended Data Fig. 3 TAD boundary alteration and transcription.

a, Illustration for how alteration of TAD boundary is defined in this study. b, COSMIC cancer-related genes that are located in TADs with recurrent change of boundary. Y axis is the number of incidences across different samples. c, Differential expression analysis of genes located inside recurrently altered TADs. P value by two-sided Student t-test.

Source Data

Extended Data Fig. 4 AML-specific loop analysis.

a, Overlap of AML-specific loops with CNVs. Color scheme: neutral (blue), gain (orange), or loss (green) of copies. b, Genome-wide CNV profiles for all the 10kb bins. c, GSEA analysis for differentially expressed genes associated with AML-specific loops. P value is calculated by permutation test.

Source Data

Extended Data Fig. 5 AML-specific loops associated with sample-specific open chromatins and gene expression.

a, Hi-C maps for the MEIS1 and ERG gene regions. Left lower halves are for AML samples and the right upper halves are for HSPC and PBMC. The loops are absent in all three PBMC samples (only one is plotted) and less frequent in the four HSPC samples (only one is plotted). b-c, RNA-Seq, CUT&Tag for H3K27ac and H3K27me3, and ATAC-Seq data for regions surrounding the three genes. Purple arcs indicate chromatin loops predicted in at least two AML samples. Shown in the figure are two representative AML samples.

Extended Data Fig. 6 Analysis of promoter-silencer (P-S) loops.

a, Percentage of P-S loops among all chromatin loops in all samples. b, Size distribution of P-S loops in all samples. c, Percentage of P-S loop anchors overlapping with CTCF binding peaks. d, Normalized EZH2 CUT&Tag signals at H3K27me3 peaks in the P-S loop anchors in THP-1 cells. e, Expression of genes that are located in the P-E vs. P-S loop anchors. P value was calculated by the two-sided Kruskal-Wallis test. P-E loops (n = 31332), P-S loops (n = 7912). Box plot: middle line denotes the median, top/bottom of boxes denotes first/third quartiles and whiskers extend to 1.5 times the interquartile range. f, Expression of genes in different patient samples with different distance to the nearest non-looping H3K27me3-marked silencer. P value calculated by one-sided Wilcoxon rank-sum test. N = 3809, 4403, 3865, 3217, 2490 genes (from left to right each category). Box plot: middle line denotes the median, top/bottom of boxes denotes first/third quartiles and whiskers extend to 1.5 times the interquartile range. g. Distance distribution between the IKZF2 promoter to control or its looped silencers in Kasumi-1 (n = 112 alleles). P value by two-sided Wilcoxon rank-sum test. Box plot: middle line denotes the median, top/bottom of boxes denotes first/third quartiles and whiskers extend to 1.5 times the interquartile range.

Source Data

Extended Data Fig. 7 Validation of a P-S loop for the RTTN gene.

a, Hi-C and H3K27me3 CUT&Tag data in AML patient sample 496 and Kasumi-1 cells, which have the P-S loop for RTTN. AML sample 018 has the same loop but with less interacting signals and almost no H3K27me3 signals at the loop anchor. The orange tracks are the 4C data for RTTN in Kasumi-1 cells. b, (Right) DNA FISH imaging to measure the distance between RTTN promoters and the looped silencers in THP-1 cells. The control is the region on the other side of the RTTN promoter, with equal linear genomic distance. (Left) distance distribution between promoter and control or its looped silencer from 144 THP-1 cells (P value = 1E-9, two-sided Wilcoxon rank-sum test). Box plot: middle line denotes the median, top/bottom of boxes denotes first/third quartiles and whiskers extend to 1.5 times the interquartile range. c, RNA expression of RTTN across patient samples and normal controls. Normal HSPC&PBMC n = 6 samples, AML w/o P-S loop n = 13, AML with P-S loop n = 10. P value was computed by the two-sided Student’s t-test. Box plot: middle line denotes the median, top/bottom of boxes denotes first/third quartiles and whiskers extend to 1.5 times the interquartile range. d, The design of PCR for detecting CRISPR deletion of the RTTN silencer (vertical bar in the H3K27me3 track). e, Two replicates of PCR results confirming the heterozygous deletion of the RTTN silencer. The red arrow points to the amplification of the CRISPR deletion junction (~700bp), and the blue arrow points to the wildtype band (~300bp). f. Sanger sequencing confirmed the CRISPR deletion of the silencer. g, qPCR results of RTTN expression in the clone with this heterozygous deletion vs. control group. The control group is other clones that underwent the same CRISPR system treatment without deletion (n = 3 technical replicates in 2 biological replicates). P value by two-sided Student’s t-test. Data show mean ± s.e.m. h, CCK-8 assay results for proliferation of Kasumi-1 cells with the RTTN silencer deleted vs. both the wildtype and the CRISPR control cells in panel g (n = 3 biological replicates). i, Stereomicroscope images of CFA cells on day 12 at 7.5X zoom. CFA was performed in n = 3 replicates. j, Size (pixel) distribution of WT colonies (n = 40) and the colonies with RTTN silencer deleted (n = 35), measured by imageJ. P value calculated by two-sided Wilcoxon rank-sum test. Box plot: middle line denotes the median, top/bottom of boxes denotes first/third quartiles and whiskers extend to 1.5 times the interquartile range.

Source Data

Extended Data Fig. 8 Analysis of chromatin stripes.

a, Hi-C maps for the MYC gene regions. Left lower panels are for AML samples and the right upper panels are for HSPC and PBMC. Black arrow marks a stripe. b, RNA-Seq, CUT&Tag for H3K27ac and H3K27me3, and ATAC-Seq data for regions surrounding the MYC genes. c, Classification of stripes based on whether the anchors contain gene promoters and the stripe zones contain enhancers (H3K27ac) or silencer marks (H3K27me3). d, In each patient samples, genes in enhancer stripes have higher expression than genes in silencer stripes. P-value by two-sided Kruskal–Wallis test. P-E stripes n = 4310, P-S stripes n = 2518. Box plot: middle line denotes the median, top/bottom of boxes denotes first/third quartiles and whiskers extend to 1.5 times the interquartile range. e. For each gene involved with a P-E stripe, samples were grouped into two categories: samples with P-E stripe, and samples with neither P-E nor P-S stripe for this gene. So was for P-S stripe analysis. Then the average gene expression (TPM) within each category was calculated. P value by two-sided Kruskal-Wallis test. Left: n = 461 genes; Right: n = 415 genes. Box plot: middle line denotes the median, top/bottom of boxes denotes first/third quartiles and whiskers extend to 1.5 times the interquartile range.

Source Data

Extended Data Fig. 9 Detection and analysis of SV-induced neo-loops.

a, Detection of SVs in AML samples from Hi-C data, marked by the black arrow in Hi-C maps. Left: inv(16). Middle: t(6;9). Right: t(8;13). b, Distribution of genomic distances between translocations/deletions to the nearest cancer-related genes (COSMIC database). Expected value is calculated by random permutation of the SVs in the genome for 1000 times. P value is calculated by two sample Kolmogorov-Smirnov test. TL: translocation. DEL: deletion. c, An example of reconstructed Hi-C maps surrounding the SV breakpoint between chr7 and chr11 in the AML sample 270. We also showed the inter-chromosomal Hi-C map in the HSPC 213, where there is no visible inter-chromosomal interactions. The orientation of the SV is marked by the arrows, which always point from 5’ to 3’. The neo-loop for CDK5 is circled. Above the Hi-C: predicted 3D structure for the region in normal and AML sample visualized by PyMOL. d, Recurrent enhancer hijacking involving the CBL gene and enhancers. e, RNA expression of the CBL gene in all samples.

Source Data

Extended Data Fig. 10 Enhancer hijacking analysis and validation.

a, Motif enrichment analysis for all hijacked enhancers in AML. The number in each box is the P-value (–log10) calculated by binomial test. b, The list of hijacked enhancers being validated, the corresponding sgRNA, and their targeted genes over SV. c, H3K27ac at the sgRNA-targeted loci and the NSMCE3 gene promoter. d, qPCR relative gene expression in THP-1 cells expressing dCas9-KRAB-MeCP2 (n = 3 technical replicates in 2 biological replicates). P value by two-sided Students’ t-test. Data show mean ± s.e.m. e, qPCR results of the targeted gene expression in Kasumi-1 cells with repressing the hijacked enhancers ((n = 3 technical replicates in 2 biological replicates)). P value by two-sided Student’s t test. Data show mean ± s.e.m.

Source Data

Extended Data Fig. 11 Aberrant DNA methylation in AML is associated with changed CTCF binding and chromatin structures.

a, Genome-wide CG methylation level in HSPC, PBMC, and AML samples. Data for HSPC* was downloaded from Tovy et al., Cell Stem Cell, 2020. b, CG methylation levels at TSS regions (upper) and gene bodies (lower) in A and B compartment. (Upper) A (n = 531,112 TSS) B (n = 204,365 TSS), (Lower) A (n = 340,385 genes) B (156,852 genes). Box plot: middle line denotes median, top/bottom of boxes denotes first/third quartiles and whiskers extend to 1.5 times the interquartile range of the first and third quartile. c, (Left) Heatmaps of CTCF CUT&Tag signals in AML samples ranked by normalized peak intensity. (Right) Methylation levels surrounding the CTCF binding sites from the same samples. d, Heatmaps showing dynamic CTCF binding sites and methylation levels in AML 424 vs. HSPC. (Right) Hypermethylated sites in AML 424 vs. HSPC that also overlapped with CTCF binding sites in HSPC. Hypermethylation is defined as CpG methylation levels significantly different between two samples (beta-binomial distribution p-value<0.01) and at least 0.3 higher in AML sample. (Left) CTCF signals at the dynamic methylated regions. e, A model for gain of loops as a result of loop extrusion over loss of CTCF binding. f, For AML 424 and HSPC 213, we separately aggregated their Hi-C plots that are centered at the 142 reduced CTCF binding sites of AML 424 associated with hypermethylation. The two aggregated plots were then normalized by total contacts and distance, and the difference is calculated as log2 of fold change. The diamond-shaped region between the two dashed lines resided the interactions across the lost CTCF sites. g, Differential CTCF binding and chromatin loops. Blue vertical bar marks a differential CTCF binding site that was absent in four AML samples. The CTCF binding motifs were also hypermethylated in the AML samples. We observed additional loops across the lost CTCF binding sites in these AML samples.

Source Data

Extended Data Fig. 12 Effect of 5-AZA treatment on DNA methylation and cell phenotypes.

a, CCK-8 assay started with n = 3 biological replicates of 1,000 cells per well in 96-well plate. The assay completed on day 8 as the DMSO group grow confluent. Data show mean ± s.e.m. MW: molecular weight. b, Western Blot results for Caspase-3. AA2 is Apoptosis Activator 2 (TOCRIS 2098). The experiments were independently repeated twice with similar results. c, Western Blot for yH2AX. UV light condition is detailed in method. The experiments were independently repeated twice with similar results. d, Flow cytometry result for cells treated with either DMSO or 5-AZA 0.5μM. e, Flow cytometry results for cells treated with 1μM, 2μM, 4μM, and 8μM 5-AZA for 2 days.

Source Data

Extended Data Fig. 13 Impact of 5-AZA treatment on chromatin structure.

a, Global CG methylation levels of U937 under DMSO or 5-AZA treatment. b, A snapshot of DNA methylation levels across a genomic region. c-d, DNA methylation levels at CpGi, CpG shores (<2Kb), CpG shelves (2Kb-4Kb), open sea (rest of genome), and CTCF binding sites under 12-day DMSO or 0.5μM 5-AZA treatment. DMSO: n = 2,691,094 CpG types, 5-AZA n = 2,669,959 CpG types. CTCF binding sites were defined from HSPC cells. N = 10,960 CTCF sites. P value by two-sided Wilcoxon rank-sum test. Violin plot: middle line denotes the median, top/bottom of boxes denotes first/third quartiles and boundary extend to 1.5 times the interquartile range. e, Enrichment of CpGi and open sea at each type of switching compartment. P value by two-sided Wilcoxon rank-sum test. From left to right for both CpGi and Open sea: n = 808, n = 372, n = 636, n = 338; Box plot: middle line denotes the median, top/bottom of boxes denotes first/third quartiles and whiskers extend to 1.5 times the interquartile range. f. Methylation levels for each type of switching compartment at CpGi (upper) and open sea (lower). From left to right: CpGi panel: n = 20,566, n = 450, n = 4,651, n = 910; Open sea panel: n = 591, n = 731, n = 4,531, n = 1,398. P value by two-sided Wilcoxon rank-sum test. AB: A in DMSO group and B in 5-AZA group. Violin plot: middle line denotes the median, top/bottom of boxes denotes first/third quartiles and boundary extend to 1.5 times the interquartile range. g, CpGi at one or both anchors of two types of loops under 5-Aza treatment. h, Methylation levels for two types of loops at CpGi and open sea. From left to right: CpGi panel: n = 2,871, n = 596; Open sea panel: n = 117,177, n = 20,499. P value by two-sided Wilcoxon rank-sum test. Violin plot: middle line denotes the median, top/bottom of boxes denotes first/third quartiles and boundary extend to 1.5 times the interquartile range.

Source Data

Supplementary information

Supplementary Figures

This file contains Supplementary Figs. 1 and 2.

Reporting Summary

Supplementary Tables

Supplementary Tables 1–14.

Source data

Rights and permissions

Springer Nature or its licensor 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

Xu, J., Song, F., Lyu, H. et al. Subtype-specific 3D genome alteration in acute myeloid leukaemia. Nature 611, 387–398 (2022). https://doi.org/10.1038/s41586-022-05365-x

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41586-022-05365-x

This article is cited by

Comments

By submitting a comment you agree to abide by our Terms and Community Guidelines. If you find something abusive or that does not comply with our terms or guidelines please flag it as inappropriate.

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