Skip to main content
bioRxiv
  • Home
  • About
  • Submit
  • ALERTS / RSS
Advanced Search
New Results

Mutational synergy coordinately remodels chromatin accessibility, enhancer landscape and 3-Dimensional DNA topology to alter gene expression during leukemia induction

Haiyang Yun, Shabana Vohra, Annalisa Mupo, George Giotopoulos, Daniel Sasca, Sarah J. Horton, Shuchi Agrawal-Singh, Eshwar Meduri, Faisal Basheer, Ludovica Marando, Malgorzata Gozdecka, Oliver M. Dovey, Aracely Castillo-Venzor, Xiaonan Wang, Paolo Gallipoli, Carsten Müller-Tidow, Cameron S. Osborne, George S. Vassiliou, View ORCID ProfileBrian J. P. Huntly
doi: https://doi.org/10.1101/2020.03.20.000059
Haiyang Yun
1Wellcome - MRC Cambridge Stem Cell Institute, Cambridge, United Kingdom
2Department of Haematology, University of Cambridge, Cambridge, United Kingdom
3Department of Medicine V, Hematology, Oncology and Rheumatology, University of Heidelberg, Heidelberg, Germany
  • Find this author on Google Scholar
  • Find this author on PubMed
  • Search for this author on this site
Shabana Vohra
1Wellcome - MRC Cambridge Stem Cell Institute, Cambridge, United Kingdom
2Department of Haematology, University of Cambridge, Cambridge, United Kingdom
  • Find this author on Google Scholar
  • Find this author on PubMed
  • Search for this author on this site
Annalisa Mupo
1Wellcome - MRC Cambridge Stem Cell Institute, Cambridge, United Kingdom
2Department of Haematology, University of Cambridge, Cambridge, United Kingdom
4Haematological Cancer Genetics, Wellcome Sanger Institute, Cambridge, United Kingdom
  • Find this author on Google Scholar
  • Find this author on PubMed
  • Search for this author on this site
George Giotopoulos
1Wellcome - MRC Cambridge Stem Cell Institute, Cambridge, United Kingdom
2Department of Haematology, University of Cambridge, Cambridge, United Kingdom
  • Find this author on Google Scholar
  • Find this author on PubMed
  • Search for this author on this site
Daniel Sasca
1Wellcome - MRC Cambridge Stem Cell Institute, Cambridge, United Kingdom
2Department of Haematology, University of Cambridge, Cambridge, United Kingdom
5Department of Hematology, Oncology and Pneumology, University Medical Center Mainz, Mainz, Germany
  • Find this author on Google Scholar
  • Find this author on PubMed
  • Search for this author on this site
Sarah J. Horton
1Wellcome - MRC Cambridge Stem Cell Institute, Cambridge, United Kingdom
2Department of Haematology, University of Cambridge, Cambridge, United Kingdom
  • Find this author on Google Scholar
  • Find this author on PubMed
  • Search for this author on this site
Shuchi Agrawal-Singh
1Wellcome - MRC Cambridge Stem Cell Institute, Cambridge, United Kingdom
2Department of Haematology, University of Cambridge, Cambridge, United Kingdom
  • Find this author on Google Scholar
  • Find this author on PubMed
  • Search for this author on this site
Eshwar Meduri
1Wellcome - MRC Cambridge Stem Cell Institute, Cambridge, United Kingdom
2Department of Haematology, University of Cambridge, Cambridge, United Kingdom
  • Find this author on Google Scholar
  • Find this author on PubMed
  • Search for this author on this site
Faisal Basheer
1Wellcome - MRC Cambridge Stem Cell Institute, Cambridge, United Kingdom
2Department of Haematology, University of Cambridge, Cambridge, United Kingdom
  • Find this author on Google Scholar
  • Find this author on PubMed
  • Search for this author on this site
Ludovica Marando
1Wellcome - MRC Cambridge Stem Cell Institute, Cambridge, United Kingdom
2Department of Haematology, University of Cambridge, Cambridge, United Kingdom
  • Find this author on Google Scholar
  • Find this author on PubMed
  • Search for this author on this site
Malgorzata Gozdecka
1Wellcome - MRC Cambridge Stem Cell Institute, Cambridge, United Kingdom
2Department of Haematology, University of Cambridge, Cambridge, United Kingdom
4Haematological Cancer Genetics, Wellcome Sanger Institute, Cambridge, United Kingdom
  • Find this author on Google Scholar
  • Find this author on PubMed
  • Search for this author on this site
Oliver M. Dovey
1Wellcome - MRC Cambridge Stem Cell Institute, Cambridge, United Kingdom
2Department of Haematology, University of Cambridge, Cambridge, United Kingdom
4Haematological Cancer Genetics, Wellcome Sanger Institute, Cambridge, United Kingdom
  • Find this author on Google Scholar
  • Find this author on PubMed
  • Search for this author on this site
Aracely Castillo-Venzor
1Wellcome - MRC Cambridge Stem Cell Institute, Cambridge, United Kingdom
  • Find this author on Google Scholar
  • Find this author on PubMed
  • Search for this author on this site
Xiaonan Wang
1Wellcome - MRC Cambridge Stem Cell Institute, Cambridge, United Kingdom
2Department of Haematology, University of Cambridge, Cambridge, United Kingdom
  • Find this author on Google Scholar
  • Find this author on PubMed
  • Search for this author on this site
Paolo Gallipoli
1Wellcome - MRC Cambridge Stem Cell Institute, Cambridge, United Kingdom
2Department of Haematology, University of Cambridge, Cambridge, United Kingdom
6Centre for Haemato-Oncology, Barts Cancer Institute, Queen Mary University of London, London, United Kingdom
  • Find this author on Google Scholar
  • Find this author on PubMed
  • Search for this author on this site
Carsten Müller-Tidow
3Department of Medicine V, Hematology, Oncology and Rheumatology, University of Heidelberg, Heidelberg, Germany
  • Find this author on Google Scholar
  • Find this author on PubMed
  • Search for this author on this site
Cameron S. Osborne
7Department of Medical and Molecular Genetics, King’s College London, London, United Kingdom
  • Find this author on Google Scholar
  • Find this author on PubMed
  • Search for this author on this site
George S. Vassiliou
1Wellcome - MRC Cambridge Stem Cell Institute, Cambridge, United Kingdom
2Department of Haematology, University of Cambridge, Cambridge, United Kingdom
4Haematological Cancer Genetics, Wellcome Sanger Institute, Cambridge, United Kingdom
  • Find this author on Google Scholar
  • Find this author on PubMed
  • Search for this author on this site
Brian J. P. Huntly
1Wellcome - MRC Cambridge Stem Cell Institute, Cambridge, United Kingdom
2Department of Haematology, University of Cambridge, Cambridge, United Kingdom
  • Find this author on Google Scholar
  • Find this author on PubMed
  • Search for this author on this site
  • ORCID record for Brian J. P. Huntly
  • For correspondence: bjph2@cam.ac.uk
  • Abstract
  • Full Text
  • Info/History
  • Metrics
  • Supplementary material
  • Preview PDF
Loading

Abstract

Altered transcription is a cardinal feature of acute myeloid leukemia (AML), however, exactly how mutations synergize to remodel the epigenetic landscape and rewire 3-Dimensional (3-D) DNA topology is unknown. Here we apply an integrated genomic approach to a murine allelic series that models the two most common mutations in AML, Flt3-ITD and Npm1c. We then deconvolute the contribution of each mutation to alterations of the epigenetic landscape and genome organization, and infer how mutations synergize in the induction of AML. These analyses allow the identification of long-range cis-regulatory circuits, including a novel super-enhancer of the Hoxa locus, as well as larger and more detailed gene-regulatory networks, whose importance we demonstrate through perturbation of network members.

Introduction

The functional specification of tissues in metazoans occurs through the generation of cell-specific proteomes, in turn, controlled by diverse transcriptional programs1,2. These transcriptional programs are tightly spatiotemporally regulated through the utilization of tissue-specific cis-regulatory elements such as enhancers, that are licensed and brought into direct communication with their cognate promoters through the functions of transcription factors (TFs), chromatin regulators (CRs) and genome structural proteins3,4. Nowhere is this complexity better demonstrated than in the hematopoietic system, where multiple mature effector cells that fulfil purposes as diverse as oxygen carriage, hemostasis and innate and adaptive immunity, derive from a single cell; the hematopoietic stem cell (HSC)5. Hematopoiesis, the dynamic and reactive process of blood production is exquisitely regulated by a complicated interplay between multiple pioneer, lineage-specific and signal-dependent TFs, CRs and structural genomic proteins6-8. These players are known to remodel the epigenetic landscape and reshape genome topology to allow communication between proximal cis-regulatory promoters and tissue-specific distal enhancer elements.

Further evidence for the importance of TFs, CRs and genomic structural proteins in coordinating normal hematopoiesis also comes from sequencing studies in hematological cancers. In malignancies such as acute myeloid leukemia (AML), recurrent mutations in TFs including CEBPA, GATA2, RUNX1, CRs such as DNMT3A, TET2 and ASXL1 and structural genomic proteins, including components of the Cohesin complex (STAG2, RAD21, SMC1A and SMC3) and CTCF have been demonstrated at high frequencies9-11. AML is the most common acute leukemia in adults, serves as a paradigm for aberrant hematopoietic differentiation and is an aggressive disease with a dismal overall survival rate of < 30%12. However, although aberrant transcription is a cardinal feature of AML13, not all cases carry mutations in the above classes of proteins, suggesting that the effects of other common mutations, such as signaling alterations, indirectly converge on the same epigenetic, transcriptional and genome regulatory machinery, although the mechanistic detail of this remains obscure.

This study addresses these fundamental questions of how AML mutations, even those that do not directly control gene expression, co-opt the transcriptional and epigenetic machinery to alter chromatin states, 3-D DNA topology and communication between enhancers and promoters to generate leukemia-specific transcriptional programs. To do so we have experimentally “deconstructed” AML utilizing an allelic series of mice that model different “transition states” during AML induction; normal, pre-malignant and overt leukemia. In addition, analysis of the single mutant (SM) pre-malignant mice also allows deconvolution of the contribution of individual mutations to altered epigenetic regulation. Using this experimental system we have attempted to deconvolute disease induction and elucidate the interplay between epigenetic and 3D genomic states and the generation of leukemia-specific transcriptional programs.

Results

Prospective modelling of AML induction in mice reveals marked transcriptional synergy between commonly co-occurring mutations

FLT3-ITD and NPM1C mutations are the most common in AML and co-occur in ∼15% of all cases9-11. Mouse models carrying knock-in mutations of Npm1 (conditional Npm1flox-cA/+;Mx1-Cre, hereafter referred to as “Npm1c”) or Flt3 (constitutive Flt3ITD/+, hereafter “Flt3-ITD”) and referred to as single mutant (SM) mice, are associated with individual subtle, non-fatal but obvious pre-malignant phenotypes14,15. However, when combined (Npm1flox-cA/+;Flt3ITD/+;Mx1-Cre, “Npm1c/Flt3-ITD” or double mutant, DM) mice develop an aggressive AML with short latency16. To prospectively assess dynamic remodeling of the cis-regulatory landscape and three-dimensional (3D) genome topology during leukemia development, we utilized an allelic series of wild-type (WT), Npm1c, Flt3-ITD and DM mice to model AML induction (Fig. 1a, upper panel). To this end, we analyzed lineage marker negative (Lin-) hematopoietic stem and progenitor cells (HSPCs) from WT and mutant mice for gene expression (RNA-seq), multiple chromatin activation states by chromatin immunoprecipitation and mass parallel sequencing (ChIP-seq), chromatin accessibility by assay for transposon accessible chromatin (ATAC-seq)17 and promoter-anchored 3-D chromatin interaction by promoter capture HiC (pCHiC)18 prior to integration of these orthogonal datasets to address our research questions (Fig. 1a, lower panel).

Fig. 1:
  • Download figure
  • Open in new tab
Fig. 1: Prospective modelling of AML induction in mice reveals marked transcriptional synergy between Npm1c and Flt3-ITD.

a. Schematic of overall experimental design. b. Principal component analysis (PCA) of mRNA expression data of 16,771 protein-coding (PC) genes. c. Volcano (left) and bar plots (right) showing differential expression between mutant and WT HSPCs using edgeR. adjP, adjusted P values; FC, fold-change. Red and blue, upregulated (adjP < 0.05 and mutants/WT FC >= 1.5) and downregulated genes (adjP < 0.05 and WT/mutants FC >= 1.5) in mutant states vs WT, respectively. d. Overlapping upregulated genes in each mutant. Numbers of genes in each segment and representative genes are indicated. Hypergeometric test p values are shown. e. Heatmap of normalised mRNA expression (z-Score) of representative genes from Fig 1.d and Supplementary Fig. 1b in WT and mutant HSPCs. f. Immune response related Hallmark gene sets from GSEA analysis of differential gene expression in Flt3-ITD or DM HSPC. Selected gene sets were significantly enriched (FDR q-value < 0.05). g. Enrichment plots showing gene set of TNFA signalling via NFKB in Flt3-ITD or DM vs WT by GSEA analysis. NES, normalised enrichment score.

As gene expression patterns determine cellular phenotype, we first analysed differential gene expression between WT, individual pre-malignant SM and leukemic DM stages. At the level of global transcription, the two SM HSPCs were clustered closely to WT, and all three were markedly distinct from transcription in DM leukemic HSPCs (Fig. 1b and Supplementary Fig. 1a). Pairwise comparisons between Npm1c or Flt3-ITD and WT HSPCs further confirmed that each mutation induced only very modest changes of global gene expression in isolation (Fig. 1c). In stark contrast, when combined, Npm1c and Flt3-ITD strongly synergised to induce marked differential gene expression (Fig. 1c). Of note, many of the genes altered in the single mutant HSPCs were also dysregulated in the same direction upon leukemia development (Fig. 1d and Supplementary Fig. 1b). Moreover, they often demonstrated a stepwise change in gene expression related to the mutational synergy (Fig 1e and Supplementary Fig. 1c). Examples of genes upregulated in Npm1c and DM cells included Spib and genes of the Hoxa locus (e.g. Hoxa9, Hoxa10), with genes such as Socs2 and Nov sequentially induced in Flt3-ITD and DM (Fig. 1d,e). A small number of genes were upregulated by both single mutations and included the AP-1 complex members Jun, Fos and Fosb, with their expression increased still further in DM HSPCs. However, some genes such as Spi1, encoding the ETS-family transcription factor Pu.1, Setbp1, and Interferon regulatory factors (e.g. Irf4, Irf8) were upregulated only in DM HSPCs (Fig. 1d,e). Of downregulated genes, tumour suppressor genes Gfi1, Gfi1b, Gata1 and Gata2 expression was sequentially reduced in Npm1c, Flt3-ITD and DM HSPCs (Fig. 1e and Supplementary Fig. 1b). Assessing global signatures by GSEA analysis, Flt3-ITD HSPCs displayed signatures particularly related to immune activation that were shared with and extended in DM Leukemic HSPCs (Fig. 1f,g). Conversely, genes downregulated in either Npm1c or DM HSPCs were significantly enriched for repressive gene signatures associated with recurrent chromosomal abnormalities (e.g. EWSR1-FLI1), transcriptional regulators (e.g. ETO2, E2F3), and leukemia stem cells (LSC) (Supplementary Fig. 1d,e). Taken together, these data suggest that isolated expression of Flt3-ITD or Npm1c during pre-malignancy induces subtle and distinctive alterations of transcription, but that together they display marked mutational synergy, consolidating existing programmes as well as generating entirely novel synergistic changes.

Flt3-ITD but not Npm1c remodels the enhancer landscape during AML evolution

To link gene expression with changes in epigenetic regulation we next assessed dynamic alterations of chromatin state (mono- or tri-methylation of Histone 3 lysine 4, H3K4me1 or H3K4me3 and acetylation of Histone 3 lysine 27, H3K27ac) at cis-regulatory elements in WT and mutant HSPCs using ChIP-seq. Enhancers were considered regions marked by high levels of H3K4me1 and no or low levels of H3K4me3 (Supplementary Fig. 2a) and were designated “primed” when lacking H3K27Ac or “active” if marked with H3K27Ac19. A total of 98,365 unique enhancer peaks were identified across all four HSPC states (Supplementary Fig. 2b). This enhancer repertoire was highly dynamic across the cellular states; 51% (50,318) of all enhancers demonstrated differential H3K4me1 between mutant and WT HSPCs (Supplementary Fig. 2b). Of note, a large number of dynamic alterations of H3K4me1 were observed for both Flt3-ITD and DM HSPCs, with an increased number in the DM leukemic state (Fig. 2a,b) but a marked overlap between the two (Supplementary Fig. 2c). Strikingly, and by contrast, Npm1c HSPCs were virtually indistinguishable from WT in their patterns of H3K4me1 (Fig. 2a, b). Moreover, among all DM leukemia-associated dynamic enhancers, intermediate changes of H3K4me1 were observed in Flt3-ITD but not Npm1c HSPCs (Fig. 2c and Supplementary Fig. 2d). These data demonstrated that Npm1c alone had no effect on enhancer specification, while Flt3-ITD (alone or in combination with Npm1c) greatly altered the enhancer landscape of HSPCs, with the constitutive activation of Flt3 defining an intermediate H3K4me1 landscape.

Fig. 2:
  • Download figure
  • Open in new tab
Fig. 2: Flt3-ITD but not Npm1c remodels the enhancer landscape during AML evolution.

a. Correlation matrix of H3K4me1 reads at each enhancer in mutant vs WT HSPCs. Counts per million (CPM) of H3K4me1 reads were log2 transformed and plotted. Pearson correlation coefficients were computed and are shown. b. MA plots showing enhancers with significantly differential H3K4me1 levels in mutant vs WT HSPCs. Differential analysis was performed using edgeR, significant changes were defined by FDR value < 0.05 and mutants/WT FC >= 1.5 (gain, red) or WT/mutants FC >= 1.5 (loss, blue). c. Profile plots of H3K4me1 enrichment scores at enhancers with gain (upper panel) or loss (lower panel) of H3K4me1 by DM across four cellular states. Plots were generated using computeMatrix and plotProfile of deepTools. d. MA plots showing enhancers with significantly differential H3K27ac levels in mutant vs WT HSPCs. Differential analysis was performed using edgeR, significant changes were defined by FDR value < 0.05 and FC >= 1.5 (gain) or FC <= −1.5 (loss). e. Number of super-enhancers (SE) with increase (UP) or decrease (DOWN) of H3K27ac in mutant vs WT HSPCs. f. Heatmaps of the H3K4me1 and H3K27ac patterns at gained (left) and lost (enhancers) for WT and DM HSPCs and for neutrophils (NE). Patterns 1 and 4 demonstrate similar changes in DM and NE cells while patterns 2 and 3 differ and are “leukemia-specific”.

We next assessed enhancer activation in WT and mutant HSPCs, overlaying the H3K27ac changes on the H3K4me1 landscape. In contrast to H3K4me1, H3K27ac alterations were relatively modest. Only ∼10% (9,456) of the entire enhancer repertoire was active in either WT or mutant HSPCs (Supplementary Fig. 3a), albeit that 59% of these active enhancers (5,575) were dynamic and demonstrated differential H3K27ac signals across the cellular states (Supplementary Fig. 3b). Similarly as for H3K4me1 levels, Flt3-ITD alone and DM demonstrated H3K27ac changes at hundreds of enhancer regions, whereas Npm1c did not significantly alter global enhancer activity (Fig. 2d). Once again, the effects of Flt3-ITD alone were often intermediate in producing differential H3K27ac when compared with DM (Supplementary Fig. 3c). Super-enhancers (SE) are regions enriched for active regulatory elements that regulate oncogenes, tumour suppressors and genes critical for cellular identity20,21. In total, 801 unique SE were identified in WT and mutant HSPCs based on H3K27ac enrichment (Supplementary Fig. 3d). The SE patterns reflected those seen for the global enhancers, with only 3 SE gained and none lost in Npm1c, 69 SE gained and 51 lost in Flt3-ITD and 210 SE gained and 214 lost in DM (Fig. 2e and Supplementary Fig. 3e).

To address whether lost and gained enhancers were truly leukemia-specific or were associated with other hematopoietic states such as normal myeloid differentiation, we replicated our integrated genomic analysis in normal neutrophils and compared enhancer modifications at the same loci. Of note, we did indeed see a degree of overlap with neutrophil enhancer states; for enhancers primed and activated (gain of H3K4me1 and H3K27ac respectively) between WT and DM HSPCs, 24% (5,572/22,801) overlapped with similar H3K4me1 and H3K27ac marks in neutrophils, however the remaining 76% appeared leukemia–specific (Fig. 2f group 2). In contrast, for decommissioned and inactivated enhancers (loss of H3K4me1 and H3K27ac), similar changes were seen in neutrophils for the majority of enhancers (20,949/26,284, 80%), while only ∼20% were leukemia-specific (Fig. 2f group 3).

Both Flt3-ITD and Npm1c significantly alter chromatin accessibility during leukemia induction

To facilitate transcriptional activation, cis-regulatory elements such as enhancers and promoters are depleted for nucleosome occupancy and become accessible to specific transcription factors22. Using ATAC-seq, we assessed the dynamics of chromatin accessibility across our allelic series of WT and mutant mice. As expected, open chromatin regions were located predominantly at enhancers and promoters (Supplementary Fig. 4a). In contrast to its minimal effects on histone modifications, Npm1c significantly affected chromatin accessibility, similar to Flt3-ITD HSPCs, demonstrating thousands of regions of altered accessibility when compared with WT HSPCs (Fig. 3a,b). As with chromatin modification, the combination of mutations in DM HSPCs induced more extensive changes on chromatin accessibility (Fig. 3a,b). In line with our previous observations, the alterations in accessibility were largely preserved between the Flt3-ITD and DM HSPCs (Fig. 3c), with accessibility changes intermediate in the Flt3-ITD SM state (Fig. 3d). In contrast, for Npm1c HSPCs, although there was a reasonable overlap for regions where accessibility was lost between Npm1c and DM HSPCs (493/1130, 43.6%), the overlap between areas that gained accessibility was small (211/2278, 9.3%; Fig. 3c and Supplementary Fig. 4b). However, within this overlap were four regions located in the Hoxa locus (Fig. 3e), where a modest increase in accessibility was associated with an increase in gene expression in both Npm1c and DM HSPCs. Regions that became sequentially less accessible in Npm1c and DM HSPCs included the Gata2 promoter, as well as its distal and proximal enhancers23,24 (Supplementary Fig. 4c), explaining the significantly reduced expression of Gata2 in DM but not Npm1c HSPCs (Fig. 1e). Regions showing dynamic accessibility involved several hundred promoters in SM and a few thousand in DM (Supplementary Fig. 4d). Interestingly, differential accessibility was positively correlated with differential gene expression at the DM leukemia stage but not in pre-malignant SM stages (Supplementary Fig. 4e).

Fig. 3:
  • Download figure
  • Open in new tab
Fig. 3: Both FLT3-ITD and Npm1c significantly alter chromatin accessibility during leukemia induction.

a, PCA of ATAC-seq signals across all four cellular states (WT, Npm1c, Flt3-ITD and DM). Normalised reads as CPM were used for the analysis. b, MA plots showing ATAC-seq peaks with significantly differential accessibility in mutant vs WT HSPCs. Differential analysis was performed using edgeR, significant changes were defined by FDR value < 0.05 and mutants/WT FC >= 2 (gain) or WT/mutants FC >= 2 (loss). c, Venn diagram of overlapping ATAC-seq peaks with accessibility gain or loss induced by each mutant. d, Profile plots of ATAC-seq signals at regions demonstrating gain or loss of accessibility induced by DM across four cellular states. e, Chromatin accessibility at the Hoxa cluster in all four HSPCs and wildtype neutrophils. ATAC-seq tracks were normalised to CPM, scaled to the same level, and visualized on WashU Epigenome Browser. Regions showing differential accessibility by Npmc1 and DM were highlighted.

Transcriptional reprogramming during leukemia evolution utilises existing chromatin interactions and reconfigures the 3-dimensional DNA topology

Current models of transcriptional regulation posit that enhancers and promoters require spatial localisation, via DNA looping, to allow communication of activating signals4. To capture genome-scale alterations in 3-D DNA topology, we employed promoter capture HiC (pCHiC)18 to generate a compendium of promoter-associated DNA interactions and demonstrate how these differed between WT and mutant HSPCs upon leukemia induction. A total of 88,624 high-confidence (HC) interactions were captured across WT, pre-malignant and leukemic HSPCs (Supplementary Fig. 5a). Each promoter had a median of nine interactions, at a median distance of 218 kilobases (Supplementary Fig. 5b,c). Again, both single and double mutant HSPCs displayed interaction profiles distinct from WT by principal component analysis, with single mutant Npm1c and Flt3-ITD more similar to each other than to DM (Fig. 4a).

Fig. 4:
  • Download figure
  • Open in new tab
Fig. 4: Transcriptional reprogramming during leukemia evolution both utilises existing chromatin interactions and reconfigures 3-dimensional DNA topology.

a. PCA of CHiCAGO scores of pCHiC high-confidence (HC) interactions across all four HSPCs. HC interactions were defined as significant interactions (CHiCAGO score > 5) present in both replicates of each state. b. Pan-chromosome correlation of chromatin compartments between mutant and WT HSPCs. c. Number of chromatin compartments showing flipped changes upon mutations. d. Illustration of chromatin compartments A/B levels at a DM-induced “B to A” flipped region (highlighted) containing Setbp1. e. Differential total interaction reads at individual promoter between mutant and WT HSPCs. Red and blue, promoters with increased (adjP < 0.05 and mutants/WT FC > 1) and decreased (adjP < 0.05 and mutants/WT FC < 1) pCHiC reads in mutant states vs WT, respectively. f. Scatter plot showing correlation between promoters with altered interactions induced by either Flt3-ITD (y axis) or DM (x axis). Red or blue, promoters with gained or lost interactions in both cells, respectively. Hypergeometric test p values are shown. g, Proportion of rewired and hard-wired interactions among all HC interactions across four HPSC states. h. Number of gained or lost interactions in the presence of mutations. i, Significant chromatin interactions, represented by arcs, between the Gfi1b promoter and its downstream regulatory region are present in WT, Npm1c and NE, but are lost in Flt3-ITD and DM HSPCs. j, Similarly, at the Irf8 promoter, interactions not present in any other cell type are gained in DM cells.

pCHiC data can further subdivide the genome into A (active) and B (inactive) compartments using principal component analysis25,26. When assessing alterations in compartment structure between WT and mutant HSPCs, no alterations in compartment assignment were demonstrated for either Npm1c or Flt3-ITD HSPCs (Fig. 4b, c). However, in marked contrast, the combination of Npm1c and Flt3-ITD in DM HSPC altered the compartment structure, with several hundred compartments changing their assignment (Fig. 4b, c). A total of 290 regions, across multiple chromosomes, altered their compartment assignment from “inactive” B to “active” A (Figure 4c). These regions contained 475 genes and included known oncogenes such as Setbp1 and Igf1 (Fig. 4d and Supplementary Fig. 5d). Moreover, whole contiguous gene regions, including the interferon-inducible Ifi gene cluster on Chromosome 1 altered their compartment assignment (Supplementary Fig. 5e). Alteration from an active to an inactive compartment, A to B, was also seen, with 274 regions reassigned in this direction (Fig. 4c). These regions contained 511 genes including the tumour suppressor gene Dach1 and the Fancf gene that is silenced by methylation in AML27.

Assessing total interactions at individual promoters, only a few promoters were altered for Npm1c HSPCs in comparison to WT. In contrast, in the presence of the Flt3-ITD, several hundred promoters showed lost or gained interactions, with this number roughly doubling in DM HSPCs (Fig. 4e). Moreover, nearly half of those promoters altered by Flt3-ITD overlapped with the change in DM HSPCs (Fig. 4f). At the level of individual unique interactions, of note, only 11% of all high-confidence interactions were dynamic (gained or lost) by SM or DM, based on stringent alterations of the ranking of CHiCAGO scores28,29 (Fig. 4g), with more interactions lost than newly formed in the mutant states (Fig. 4h). We named these “rewired” interactions, and interactions could be both lost (e.g. at the Gfi1b promoter, Fig. 4i) or gained (e.g. for Irf8 promoter, Fig. 4j). In contrast, the remainder of interactions were already present in WT HSPCs and were termed “hard-wired” DNA contacts (Fig. 4g).

Integration of chromatin alterations, DNA topology and gene expression identifies critical transcriptional network nodes

Gene expression requires both instructive signals from accessible, activated chromatin at proximal and distal cis-regulatory elements, as well as their communication through 3-D interaction4. We therefore integrated our ChIP-Seq and pCHiC analyses using specific DNA interactions, rather than proximity alone, to assign cis-regulatory elements to their specific promoters, thus accurately linking alterations of chromatin modification and accessibility at these elements with changes in gene expression. As expected, alterations of enhancer marks correlated globally with gene expression upon leukemia induction, albeit modestly. For H3K4me1, an increase in gene expression was associated with 18% of primed enhancers and a decrease in gene expression was observed in 15% of decommissioned enhancers with significant H3K4me1 loss (Fig. 5a). Similarly, for H3K27ac, 31% of activated enhancers were associated with an increase in gene expression and 23% of deactivated enhancers with decreased gene expression (Fig. 5a). Gene expression also correlated with chromatin accessibility, but more significantly for increased promoter accessibility in DM HSPC (859/1913, 45% of genes upregulated, Supplementary Fig. 4e). In the “flipped” genome compartments, the gene expression pattern generally correlated with expectations; 166/345 genes (48%) were upregulated in compartments that changed from B to A and 145/453 genes (32%) downregulated in compartments that changed from A to B (Supplementary Fig. 6a).

Fig. 5:
  • Download figure
  • Open in new tab
Fig. 5: Integration of chromatin alterations, DNA topology and gene expression identifies critical transcriptional networks nodes.

a. Percentage of enhancers with H3K4me1/H3K27ac changes showing differential mRNA expression of their physically linked genes in DM vs WT HSPCs. b. De novo motifs significantly enriched at genomic regions gaining accessibility in the presence of Flt3-ITD or DM. c. Comparison of alteration of H3K4me1 between WT and DM states at enhancers that gained (upper graph) or lost (lower graph) interactions. Red and blue, upregulated (adjP < 0.05 and DM/WT FC >= 1.5) and downregulated H3K4me1 readcount (adjP < 0.05 and WT/DM FC >= 1.5) in DM vs WT, respectively. d. Enrichment of H3K4me1 at enhancers involved in flipped chromatin compartments in DM vs WT HSPCs. e. Combined profiles of Spi1 mRNA expression with chromatin states, accessibility at its upstream enhancers (Spi1-URE), as well as 3-D interaction between Spi1 promoter and Spi1-URE in all four HPSCs and WT neutrophils. f. Combined profiles of Hoxa genes expression with chromatin states, accessibility at its long-range upstream super-enhancers (Hoxa-LRSE), as well as 3-D interaction between Hoxa cluster and Hoxa-LSRE in all four HPSCs and WT neutrophils.

To identify TFs within the transcriptional networks that regulated the epigenetic landscape, we performed de novo motif analysis of the differential H3K4me1, H3K27ac and ATAC accessible regions. Similar findings were evident for differential enhancer marks and accessibility: regions that gained modification or accessibility were enriched for PU.1 and RUNX motifs, for motifs of the IRF signal inducible transcription factors and for AP-1 members in both Flt3-ITD and DM HSPCs (Fig. 5b). Of note, expression levels of their cognate TFs Spi1/Pu.1, the AP-1 TFs Jun, Junb, Fos and Fosb and a raft of Irf TFs (Irf1, 2, 4, 5, 6, 7, 8, and 9) also increased in mutant HSPCs (Fig. 1e). For regions that had lost modification or accessibility, enrichment for GATA and CTCF motifs was demonstrated (Supplementary Fig. 6b) and as previously noted DM HSPCs demonstrated decreased expression of both Gata1 and Gata2 (Fig. 1e). In addition, focussing the same analysis on highly specific regions, we could further demonstrate that STAT5 motifs were also enriched in leukemia-specific regions of enhancer gain (Supplementary Fig. 6c) and that HOX motifs were enriched in regions where compartment status changed from B to A (Supplementary Fig. 6d).

Integrating chromatin modifications with dynamic interactions, we demonstrated a strong correlation of H3K4me1 alteration with “rewired” interactions (Fig. 5c and Supplementary Fig. 6e), with H3K4me1 gain often evident at enhancers that formed new interactions and its loss seen at enhancers where interactions were abrogated. Similarly, in the “flipped” compartments, transition from B to A associated strongly with increased H3K4me1 modification and acquisition of primed enhancers, but also with an increase in activated enhancers and chromatin accessibility (Fig. 5d and Supplementary Fig. 6f,g). Conversely, compartments that transitioned from A to B were enriched for decommissioned and deactivated enhancers and loss of chromatin accessibility (Fig. 5d and Supplementary Fig. 6f,g).

Our integrated analysis suggested Pu.1 as an important network regulator in Flt3-ITD/Npm1c leukemia. We therefore assessed the regulation of its gene (Spi1) during leukemia induction and could demonstrate coordinated changes in histone modification, chromatin accessibility and promoter-enhancer interactions across states (Fig. 5e). All WT and mutant HSPCs demonstrated H3K4me1 priming and some H3K27ac at the well described Spi1 upstream regulatory element (URE)30-33. However, at this locus, accessibility was increased upon Flt3-ITD expression and H3K27ac increased in DM HSPCs. These changes were associated with a “rewiring” of its URE in DM HSPCs to allow communication of the activating signals from the enhancer and correlated with the upregulation of Pu.1 expression only in the rewired DM leukemic HSPCs.

Definitively linking promoters to enhancer regions using pCHiC also allowed us to identify novel long-range regulatory interactions. An exemplar of this was seen for the Hoxa locus, where a “hardwired” interaction with an H3K4me1 modified region > 1MB upstream (Fig. 5f) could be demonstrated for all WT and mutant HSPCs. However, this region demonstrated a marked increase in H3K27ac modification in both Npm1c and in DM HSPCs. Of note, acquisition of this novel super-enhancer, which we have named the Hoxa-long range super-enhancer (Hoxa-LRSE), one of only three super-enhancers gained in Npm1c mutated HSPCs, led to a marked upregulation of multiple Hoxa genes (Supplementary Fig. 6h), where it appears to collaborate with the increased accessibility across the promoters of Hoxa genes (Supplementary Fig. 7a). In addition, and as a direct link to human AML, a wider region on murine chromosome 6 that spans the entire interaction is syntenic with a homologous region on human chromosome 7. Using the LIFTOVER tool to assess cross-species homology, we could identify a homologous region to the mouse Hoxa-LRSE in the human genome (Supplementary Fig. 7b). Moreover, using published pCHiC data from human CD34+ cells18, we could also demonstrate that the interaction was also present across species, and ChIP-seq data from ourselves34 and others35,36 indicated that this region also contained a SE in human AML cells that overexpress HOXA genes (OCI-AML3 cell line) but not in AML cells that do not (Kasumi-1 cell line) (Supplementary Fig. 7b). The human HOXA-LRSE was also demonstrated to have an open chromatin confirmation in AML patients by DNase hypersensitivity analysis37,38, further demonstrating conservation of this long-range regulation in humans.

Perturbation of critical cis-regulatory hubs and their target genes abrogates leukemia maintenance

Integration of the transcriptional, chromatin modification, accessibility and DNA-interaction data allows identification of putative networks of critical regulatory elements, their trans-regulatory factors and the critical downstream genes that these regulate. As exemplars of putative network nodes we chose the AP-1 complex members c-Fos and c-Jun, the transcription factors Spi1/Pu.1 and the Hoxa genes and their regulatory loci, and validated their critical role in the maintenance of Flt3-ITD/Npm1c AML using RNAi and CRISPR-editing. As proof of this principle, we could demonstrate that knockdown of two partners of the AP-1 complex, c-Jun and c-Fos, significantly decreased clonogenic capacity in DM cells using in vitro assays (Supplementary Fig. 8a,b). We could further demonstrate that shRNA-mediated depletion of Spi1 could abrogate leukemia proliferation in liquid culture and clonogenic function in colony formation, confirming its requirement to maintain leukemia growth (Supplementary Fig. 8c-e). Utilising ex vivo cultured bone marrow cells harvested from mice carrying Npm1c/Flt3-ITD/Cas939 (termed DM-Cas9 cells), CRISPR-Cas9 mediated genetic excision of the Spi1-URE was achieved using dual guide RNAs to target its 5kb central region (Fig. 6a and Supplementary Fig. 8f,g). Removal of the URE in DM-Cas9 cells resulted in ∼30% reduction of Spi1 expression in bulk cultures (Fig. 6b), which generated a significantly decreased number of colonies in methylcellulose plating assays and decreased proliferation of bulk leukemia cells in liquid culture (Fig. 6c,d and Supplementary Fig. 8h).

Fig. 6:
  • Download figure
  • Open in new tab
Fig. 6: CRISPR-Cas9 mediated disruption of critical cis-regulatory elements perturbs target gene expression and abrogates leukemia maintenance.

a. Experimental design of strategy targeting Spi1-URE for deletion using CRISPR-Cas9 plus dual guide RNAs (gRNA) flanking its ∼5kb central region. Primer 1 and 2 are oligos for PCR confirmation of deletion. Sizes of PCR products from wildtype allele (wt) or modified allele (del) were indicated (see also Supp Fig 8f,g). b. Spi1 mRNA expression detected by RT-qPCR in DM-Cas9 cells expressing Spi1-URE gRNAs relative to control gRNAs. *, p < 0.05. c. Ex vivo cell proliferation in DM-Cas9 cells expressing Spi1-URE gRNAs relative to normalised control gRNAs growth in liquid culture. ns, non-significant; *, p < 0.05; ***, p < 0.001. d. Colony-forming unit (CFU) assay of DM-Cas9 cells expressing Spi1-URE gRNAs or control gRNAs in methylcellulose culture. ***, p < 0.001. e. Experimental design of targeting Hoxa-LRSE for deletion using CRISPR-Cas9 plus dual gRNAs flanking its ∼2.4kb central region (Supp Fig 8i,j). f. mRNA expression of Hoxa genes detected in DM-Cas9 cells expressing Hoxa-LRSE gRNAs relative to control gRNAs. *, p < 0.05; **, p < 0.01. g. Ex vivo cell proliferation in DM-Cas9 cells expressing Hoxa-LRSE gRNAs relative to normalised control gRNAs growth in liquid culture. *, p < 0.05; **, p < 0.01; ***, p < 0.001. h. CFU assay of DM-Cas9 cells expressing Hoxa-LRSE gRNAs or control gRNAs in methylcellulose culture. # indicates three independent single cell derived clones. ***, p < 0.001.

We could also demonstrate the requirement for sustained Hoxa gene expression within our AML network, with knockdown of Hoxa9 or Hoxa10 abrogating leukemia cell proliferation and clonogenicity in liquid culture and plating experiments (Supplementary Fig. 8c-e). We used a similar experimental strategy to examine the role of the Hoxa-LRSE in Hoxa gene regulation and the leukemia phenotype, by removing its ∼2.4kb central region in DM-Cas9 cells (Fig. 6e and Supplementary Fig. 8i,j). Genetic disruption of Hoxa-LRSE resulted in significant reduced transcription of all Hoxa genes tested (Hoxa3, 5, 6, 7, 9, 10, Fig. 6f). In addition, as with the Spi1-URE, excision of the Hoxa-LRSE also significantly reduced colony-forming capacity and was associated with a significant decline in cell proliferation (Fig. 6g, h and Supplementary Fig. 8k). Taken together, perturbation of these exemplar nodes validate our chromatin profiling and interaction analysis strategy to identify critical network members and their cis-regulatory elements and demonstrated these nodes to include AP-1 complex members, Pu.1 and Hoxa genes that were induced by Npm1c and Flt3-ITD, and are required for the maintenance of the leukemia.

Discussion

By deconstructing an experimental model of a common AML subtype, our study infers the contribution of the commonest individual mutations in AML, where it documents their marked synergy in remodeling the epigenetic landscape, 3D-DNA topology and transcriptional output during the induction of disease. The study demonstrated obvious effects at chromatin, although only very modest changes in gene expression, for both Flt3-ITD and Npm1c. More surprisingly, it identified marked functional differences between the two mutations, with Flt3-ITD able to alter chromatin modification, chromatin accessibility and 3D-DNA topology, but Npm1c only able to alter chromatin accessibility. We could also demonstrate marked synergy for the same processes when the mutations were combined, particularly at the level of transcription, an observation that suggests functional co-selection at the single cell level and likely underlies their very common co-occurrence in AML. While chromatin accessibility, enhancer specification and activation were highly dynamic processes related to the presence of mutations during leukemia induction, mutation-associated alterations in 3D-DNA topology were much less frequent. Moreover, the use of promoter-based capture HiC in our study allowed categorical annotation of physical 3D interactions between promoters and distal enhancer elements rather than their inference through linear proximity, allowing the identification of many novel long-range interactions. Finally, integrating these epigenetic and transcriptional layers together allowed the construction of regulatory networks and the identification of major nodes that defined them, including trans-acting TF and cis-regulatory elements, which we have functionally validated through perturbation experiments at exemplar region and factors.

Our study prospectively confirms that common AML mutations that lack direct epigenetic or transcriptional effects can indirectly alter the epigenetic landscape and transcription to generate convergent leukemia-associated transcriptional programs. Somewhat to our surprise, particularly given the very modest transcriptional changes seen in single mutant states, this was most evident for HSPCs that expressed Flt3-ITD, where significant alterations in chromatin accessibility, 3D-DNA topology and chromatin modifications, particularly the enhancer specifying mark H3K4me1, were evident. Overlap between the Flt3-ITD and DM states suggested Flt3-ITD to have a dominant role in malignant remodeling of the epigenome and 3D genome. Moreover, overt gene expression changes seen in DM cells were often preceded with epigenetic alterations in regulatory elements associated with the same genes in Flt3-ITD HSPCs, observations reminiscent of normal hematopoietic differentiation, where epigenetic changes precede gene expression changes that drive specific differentiated cell types40, suggesting a degree of epigenetic “priming” by AML mutations, particularly Flt3-ITD. Mechanistically, GSEA demonstrated that Flt3-ITD signaling upregulated inflammatory response gene programs, activating and/or mimicked major inflammatory signaling pathways (TNF-α, NF-κB, IFN-α and γ), and that multiple chemokine ligand (Ccl2, 3, 4, 5, 6, 17, 22, 24) and chemokine receptor (Ccr2, 3, 5, 6, 7, 9, Ccr1l1 and Ccrl2) genes were upregulated in DM cells, including multiple cognate chemokine ligand-receptor pairs (including Ccr2-cCc2, Ccr3-Ccl5 or Ccr3-Ccl24 and Ccr5-Ccl3, Ccr5 Ccl-4 or Ccr5-Ccl541,42) that might drive paracrine loops. In addition, de novo motif analysis and RNAi and CRISPR editing experiments demonstrated the role of signal-inducible TFs such as AP143, interferon response factors (particularly IRF4 and -8)43 and signal responsive TFs such as PU.144 and RUNX145 in this regulation. Of note, these findings accord with and extend similar integrated genomic analysis in patient samples from FLT3-ITD mutated AML37,38. By contrast, the major alteration that occurred in HSPC expressing Npm1c was an alteration of chromatin accessibility, with an obvious exemplar the changes at the Hoxa locus demonstrated to drive leukemogenesis. The nature of this alteration in chromatin accessibility is not immediately apparent, however, may relate to the known histone chaperone function of NPM146, to its interaction with ATP-dependent chromatin remodeling complexes22 or to its role as the major protein component of the nucleolus, a structure that regulates genome 3D topology47,48 and further work is warranted to investigate these possibilities.

Our study models the alterations in 3-D genome organization that accompany the induction of hematological malignancy and provides interesting insights into the role of DNA topology in cancer. Of note, the combined mutant state was able to alter gross genome organization at the level of genome compartments. However, although dynamic, only 10% of interactions were rewired secondary to the presence of Flt3-ITD and Npm1c, with these rewired interactions associated with alterations of H3K4me1, as has been previously suggested in correlative and experimental systems49. As expected, many of these rewired interactions were also associated with alterations in gene expression. However, the majority of genes whose expression was altered demonstrated no dynamic alteration and occurred at preformed, hardwired interactions, with the gene expression alterations presumably reflecting alterations of chromatin modification and accessibility at these loci. As these interactions likely reflect at least part of the “potential” functional cistrome of a given tissue, it is interesting to see that enhancers differentially utilized later in neutrophil differentiation were “repurposed” in the DM leukemic state, albeit that more were decommissioned than activated. In addition, for some rewired interactions, such as those at the Spi1/Pu.1 URE (Fig. 5e), the same interactions were also seen in neutrophils, however, others, such as those at Irf8 enhancers (Fig. 4j and Supplementary Fig. 3h), were not repurposed from neutrophil differentiation and appear leukemia-specific. This suggests that malignant transformation can generate novel DNA-topologies and enhancer states but that it may also utilize enhancers and DNA topologies associated with other differentiation states in the tissue from which it is derived, and future work is required to further address this hypothesis.

Our experimental framework has also allowed us to identify exemplar regulatory network nodes, such as the Hoxa genes, AP1 complex members and Pu.1 and IRF transcription factors. Of note, when we compare these same factors with our own and other CRISPR screens performed in human AML cell line models50,51, we see that they constitute similar vulnerabilities across multiple AML genotypes. Regarding Spi1/Pu.1, although one of the first descriptions of it relates to its overexpression in erythroleukemia associated with retroviral insertion of the spleen focus-forming virus (SFFV), PU.1 is classically regarded as a tumour suppressor in AML. Mice with genetically engineered decreased PU.1 expression, via deletion of its upstream regulatory element (URE) or PU.1 knockdown develop AML32,33,52,53. In further support of its role as a TSG, heterozygous deletions of the locus or rare loss-of-function mutations of PU.1 have been described in patients with AML9,11,54. Moreover, oncogenic translocations or point mutations involving transcription factors such as RUNX1 abrogate the activity of PU.155,56 and therapeutically, the anti-leukemic activity of LSD1 inhibitors has been associated with re-activation of a PU.1 driven transcriptional programme57,58. However, in contrast, our study identifies Spi1/PU.1 as an oncogenic transcription factor in Npm1c/Flt3-ITD AML, and is in accord with our previous work and that of others that have demonstrated PU.1 to be a vulnerability identified by CRISPR screens in certain AML subtypes50,51,59. These findings suggest the role of PU.1 to be cell context, genotype and possibly stage-specific in AML and warrant further study. Of note, PU.1’s oncogenic effect appears to occur in combination with its known co-activators, AP-160,61, IRF transcription factors62,63 and Hoxa964,65, with this observation a further demonstration of the synergy between Npm1c and Flt3-ITD in that there is coordinated upregulation of these synergistic TF when the mutations occur together. Finally, our findings not only add mechanistic insight into the association between Hoxa cluster gene expression and the Npm1c mutation, they also describe an entirely novel Hoxa-long range superenhancer (Hoxa-LRSE) that appears to regulate the entire cluster. Importantly, we demonstrate that the Hoxa-LRSE is not species-specific but is conserved within human cells. Further study of the Hoxa-LRSE and the integration of its signal to that of local HOX gene regulatory sequences will further inform the study of these genes that are critical for hematopoiesis and leukemogenesis66,67.

Taken together our work demonstrates the complicated interplay that occurs between synergistic mutations to remodel the epigenetic landscape and rewire the epigenome to induce and maintain leukemogenic transcriptional programs and identifies critical network characteristics that might be targeted for therapeutic gain.

Methods

Mice

C57/BL6 strain mice carrying single mutations (conditional Npm1flox-cA/+;Mx1-Cre+, or constitutive Flt3ITD/+) or the combined mutations (Npm1flox-cA/+;Flt3ITD/+;Mx1-Cre+) have been described previously14,15,16. The activation of the Npm1cA allele was achieved by induction of Cre-mediated recombination in 5- to 6-week-old Npm1flox-cA/+;Mx1-Cre+ mice which were administered six doses of polyinosinic-polycytidylic acid(pIpC). The activation of Npm1cA allele in mice carrying Npm1flox-cA/+;Flt3ITD/+;Mx1-Cre+ does not require administration of pIpC to induce Mx1-Cre for recombination16. For tissue harvest, wildtype (WT) C57/BL6 mice and Flt3ITD/+ mice were sacrificed at 10-12 weeks of age; Npm1flox-cA/+ mice were sacrificed immediately after 6x pIpC treatment (aged 11-12 weeks of age); mice carrying the combined mutations were sacrificed at 5-6 weeks of age. Our animal procedures were regulated under UK Home Office Animals (Scientific Procedures) Act 1986 Amendment Regulations 2012 under project license 80/2564.

Isolation of mouse hematopoietic populations

Femurs and tibias were harvested from humanely euthanized mice as a source of hematopoietic bone marrow (BM) cells. BM cells were collected by flushing femurs and tibias with sterile PBS and were filtered through 70-µm EASYstrainer (Greiner). Red blood cells were then depleted using RBC lysis buffer (5 PRIME). The isolation of lineage negative (Lin-) HSPCs or neutrophils were performed using Lineage Cell depletion kit or Neutrophil Isolation Kit (Miltenyi Biotec) according to the manufacturer’s instructions.

RNA-seq

Procedures of RNA-seq experiments have been described previously68. Briefly, total RNA was extracted from Lin-HSPCs or neutrophils isolated from wildtype or mutant mice, using RNeasy Plus Mini kit (Qiagen) according to the manufacturer’s instructions. RNA quantity and quality was measured using a NanoDrop 1000 Spectrophotometer (Thermo Scientific) and RNA 6000 Nano Kit (Agilent Technologies) on an Agilent 2100 bioanalyzer, respectively. For each experiment, 2-5 µg total RNA was processed for ribosomal RNA depletion and then for library preparation using a Rapid Directional RNA-Seq Kit (NEXTflex). Barcoded libraries were pooled together for paired-end (2× 125 bp) sequencing on a Illumina HiSeq 2500 at the Cancer Research UK (CRUK) Cambridge Institute Genomics Core. Two independent RNA samples of each cellular state were used to generate two biological replicates of RNA-seq libraries.

Gene expression analysis

RNA-seq reads were quality filtered and mapped using STAR69 (version 2.4.0) against the mouse genome (mm10). Uniquely mapped reads were quantified with HTSeq and protein-coding (PC) genes with non-zero read count in wildtype or mutant HSPCs (n = 16,771) were included for downstream analysis. Reads Per Kilobase Million (RPKM) mapped reads for each PC genes were calculated using Bioconductor package DiffBind70. Principal component analysis (PCA) was performed on RPKM values of 16,771 PC genes. Differential expression of PC genes was analysed with these counts using Bioconductor package DESeq271. Significantly differential expression was considered by setting adjusted P value (adjP) < 0.05 and fold change (FC) >= 1.5 between mutants and wildtype HSPCs. Unsupervised clustering was used to generate the heatmap.

Gene set enrichment analysis (GSEA)

GSEA was performed on pre-ranked gene lists generated by differential gene expression analysis using DESeq2 by running GSEAPreranked module of GSEA 3.072. Significantly enriched gene sets with normalised enrichment score (NES) > 2 or < −2 and FDR q value < 0.05 were shown.

ChIP-seq

ChIP-seq was performed as previously described68. Chromatin immunoprecipitation was performed with iDeal ChIP-seq kit for Histones (Diagenode) following manufacturer’s recommendations. In brief, 1 × 106 Lin-HSPCs or neutrophils isolated from wildtype or mutant mice were crosslinked with formaldehyde (Thermo Fisher Scientific) at a final concentration of 1% for 10 minutes and then quenched by addition of Glycine (Thermo Fisher Scientific) for 5 minutes incubation. Lysis buffers iL1 and iL2 were used to prepare the nuclei pellets. Chromatin was sheared in Shearing buffer iS1 using the Bioruptor Plus (Diagenode) for 20 cycles (each cycle for 30 seconds “ON” and 45 seconds “OFF”) at high power setting. Immunoprecipitation, wash, decrosslinking and elution were carried out as per the manufacturer’s protocol, using 1 µg antibody (anti-H3K4me1, #ab8895, Abcam; anti-H3K27ac, #ab4729, Abcam; anti-H3K4me3, #c15410003, Diagenode). 10% of the sheared chromatins were kept aside as input samples. ChIP-seq library preparation of ChIP DNA or input DNA was performed using TruSeq ChIP Sample Prep Kit (Illumina) following standard procedures from manufacturer. The library DNA was quantified using KAPA Library Quantification Kit (Kapa Biosystems) and library average size was determined by Agilent DNA 1000 Kit (Agilent Technologies). Libraries were pooled for single-read (1 × 50 bp) sequencing on an Illumina HiSeq 2500 or HiSeq 4000 platform at CRUK Cambridge Institute Genomics Core. Experiments were performed in duplicate on biologically independent samples.

Analysis of histone modifications

ChIP-seq reads were aligned to the mouse reference genome (mm10) using Bowtie (version 2.1.0)73. Duplicates were removed using PICARD tools (http://broadinstitute.github.io/picard/). As for H3K4me1 or H3K4me3, enriched regions (peaks) were identified using findPeaks program from Homer software25, with the setting of “-size 1000 -minDist 1000”. Peaks overlapped in both replicates of individual cell type were kept. If centers of two consecutive peaks were less than 500 bp, only the peak with higher enrichment signal was counted. H3K4me1 or H3K4me3 peak sets were generated by combining total peaks in all four HPSCs and extending +/- 500 bp. Frequency distribution of H3K4me3 peaks were then plotted based on H3K4me3 enrichment, and peaks with high levels of H4me3 (CPM > 16, Supplementary Fig. 2a) were considered to be active promoters. H3K4me1-based enhancer catalog was created by excluding H3K4me1 peaks that intersect with active promoters, and then merging the overlapped ones. H3K27ac ChIP-seq peaks were called using MACS225 against input sample of each cell type with a P-value cutoff of 1e-9 and with setting “nomodel”. A list of H3K27ac consensus peak set was made using DiffBind. Enhancers that overlap with any H3K27ac peaks were defined as active enhancers. Differential enrichment of chromatin marks (H3k4me1 and H3k27ac) was analysed using edgeR, with significant changes being defined by FDR value < 0.05 and FC >= 1.5 (gain or loss) in the presence of any mutations. Enhancers with gain or loss of H3K4me1 levels in any mutant vs WT HSPCs were defined as dynamic enhancers, likewise, H3K27ac peaks with mutation-induced gain or loss of H3K27ac enrichment were considered with dynamic activity.

Super-enhancers (SEs) analysis

The method of identifying SEs based on H3K27ac ChIP-seq data has been introduced previously74. In brief, promoter H3K27ac peaks (+/- 1kb from TSS) were first removed from H3K27ac consensus peak set, and then stitch the rest of the peaks if within 12.5kb (even across some promoters). The reads at stitched peaks in replicates for each condition were merged accordingly. All stitched peaks with their corresponding reads in each cellular state were concatenated to a full list. Then Rose R script74 was run to determine the threshold for calling SEs. SEs with differential H3K27ac were identified using edgeR, with significant changes being defined by FDR value < 0.05 and FC >= 1.5 (gain or loss) in the presence of any mutations.

Identification of leukemia-specific enhancer changes

Enhancers with DM-induced significantly differential (gained or lost) H3K4me1 were selected for their H3K4me1 enrichment in WT HSPCs, DM HSPCs and WT neutrophils using deepTools. Two subgroups of each gained or lost enhancer groups were formed based on H3K4me1 enrichment patterns by k-means clustering (n = 2). The subgroups demonstrate different changes in DM and NE cells were considered as “leukemia-specific” altered enhancers. Enrichment of H3K27ac at these subgroups was plotted side by side with H3K4me1 signals.

ATAC-seq

Following the original protocol17, ATAC-seq was carried out with some modifications. In brief, 100,000 freshly isolated cells were pelleted and washed with 500 µl ice-cold PBS. The cell pellet was collected by spinning at 500 x g for 5 minutes at 4 °C. Then the pellet was resuspended in 100 µL Lysis buffer and gently pipetted up and down 10 times, followed by spinning down for at 500 x g for 10 minutes at 4 °C. The crude nuclei pellet was preserved for tagmentation using Nextera DNA Sample Preparation Kit (Illumina) for 30 minutes incubation at 37 °C. Afterwards, transposed DNA was purified using with MinElute PCR Purification kit (Qiagen). Eluted DNA was amplified by setting the PCR cycling as below: step 1. 72 °C 5 minutes, 98 °C 30 seconds; step 2. 11-12 cycles: 98 °C 10 seconds, 63 °C 30 seconds, and 72 °C 1 minute; step 3. 72 °C 5 minutes. PCR products were again purified with Qiagen MinElute PCR Purification kit. The quantity and quality of library DNA was measured using KAPA Library Quantification kit (Kapa Biosystems) and Bioanalyzer 2100 system with a High Sensitivity DNA chip (Agilent Technologies) respectively. The pooled libraries were sequenced on Hiseq 4000 at CRUK Cambridge Institute Genomics Core. Two biological replicates were performed independently for each cellular condition.

Chromatin accessibility analysis

Chromatin accessibility probed by ATAC-seq was analysed in like manner as H3K27ac ChIP-seq analysis, including the procedures of trimming, mapping, filtering and peak calling. Briefly, trimmed sequences were mapped against mm10 reference genome using Bowtie2 and only uniquely mapped reads were kept. Peaks were called using MACS2 with the setting “-nomodel-nolambda”. A list of ATAC-seq consensus peak set was made using DiffBind. Differential enrichment was analysed using edgeR, with significant changes being defined by FDR value < 0.05 and FC >= 2 (gain or loss) in the presence of any mutations.

Promoter capture HiC (pCHiC)

pCHiC assays were carried out in all four Lin-HSPCs and wildtype neutrophils according to an established protocol75 using murine promoter bait library. Briefly, the whole experimental procedure covers three stages: 1. generation of HiC DNA ligation pool; 2. generation of HiC library; 3. generation of capture HiC library. At stage 1, to start, 20-25 million cells were crosslinked by 1% formaldehyde for 10 minutes at room temperature and the reaction was then quenched by glycine. Cells were lysed with the help of homogenization. Chromatin digestion was achieved overnight using 1000 units of HindIII (NEB). Digested DNA ends were biotinylated by filling with biotin-14-dATP (Invitrogen) and underwent in-nucleus ligation using 40 units of T4 DNA ligase (Invitrogen) at 16 °C overnight. Crosslink was then reversed by incubation with Proteinase K (Roche) at 65°C overnight. RNA was removed by adding RNase A and incubating at 37°C for 1 hour. DNA was purified twice with Phenol:Chloroform:Isoamyl Alcohol (Sigma-Aldrich) and was quantified using Qubit dsDNA HS assays (Invitrogen). No less than 40 μg DNA was required for the following steps. At stage 2, extra Biotin-14-dATP was first removed using 15 units of T4 DNA polymerase (NEB) and incubating at 20°C for 4 hours. DNA was sheared using Covaris S2 to achieve a target peak around 400 bp, with parameters: Water Level, 12; Intensity, 4; Duty Cycle, 10%; Cycles per burst, 200; Treatment time, 55 seconds. Sheared DNA was subject to end repair and purification, followed by addition of dATP as A-tail. DNA fragments were size selected to enrich at 200-650 bp by double-sided SPRI beads (0.6x followed by 0.9x, Invitrogen). Biotin-labeled fragments were pulled down using Dynabeads MyOne Streptavidin C1 (Invitrogen) and ligated to pre-annealed paired-end (PE) adapter on beads. Eventually, HiC DNA was amplified using PE PCR primers and Phusion DNA polymerase (NEB) for 7-9 cycles. HiC library was purified twice using 1.8x SPRI beads (which also minimized the contamination of primer dimers). At stage 3, 500 ng of HiC library was used for hybridization with biotinylated RNA bait library (Agilent Technologies), which contained 39684 biotinylated RNA probes covering 22047 promoters. Hybridization was performed as per manufacture’s protocol. Dynabeads MyOne Streptavidin T1 beads were utilized to pull down biotin-labeled DNA-RNA hybrids. Final PCR amplification and DNA purification of promoter capture HiC library were performed similarly as for HiC library, with the difference for 4 cycles of PCR. pCHiC were performed in biological duplicates of each cellular states. Per pCHiC library of each replicate was subject to paired-end (2× 125bp) sequencing on a single lane on HiSeq 2500 or HiSeq 4000 platform at CRUK Cambridge Institute Genomics Core.

Promoter-anchored interaction analysis

Paired-end sequences of pCHiC were processed using the HiCUP76 pipeline with default parameters for the following steps: quality control, identification of reads containing HiC junctions, mapping to reference genome mm10, and filtering duplicated HiC di-tags. Output bam files containing valid HiC di-tags were processed by Bioconductor package CHiCAGO77, to call significant promoter-based interactions. CHiCAGO considers distance effect on interaction frequencies by virtue of a convolution background model and a distance-weighted p-value. CHiCAGO scores represent -log weighted p-values, the higher the more likelihood of interaction formed. CHiCAGO scores were calculated for each pCHiC sample, and significant interactions were called when CHiCAGO scores > 5. Significant interactions were computed per HSPCs by merging their replicates, and were combined to form a matrix of total unique interactions. While, interactions present in both replicates per HSPCs were considered with high confidence (HC). Total pCHiC reads at individual promoters were summed to perform differential analysis in mutant vs WT HSPCs using edgeR. Differential total interaction reads were defined by adjP < 0.05 and absolute FC > 1. Rewired interactions were identified by comparing and ranking their CHiCAGO scores in WT and mutant cells. Those HC interactions, absent in WT (score <= 5) but present in mutant (score > 5), with scores ranked in the bottom quartile in WT but in the top quartile in mutant, were considered mutation-associated gained interactions. And vice versa, HC interactions that were present in WT but absent in mutant, ranked in the top quartile in WT but in the bottom quartile in mutant, were considered as lost interactions by mutations. In addition, total unique interaction profiles facilitated the annotation of distal cis-regulatory elements (e.g. SEs) to their target genes in HSPCs.

Chromatin compartment analysis

Sub-nuclear compartmentation represented by self-associating chromatin domains were analysed by means of principal component analysis (PCA) on capture HiC data in a similar way as described on HiC data26, mainly using Homer software. To do so, uniquely mapped reads from HiCUP analysis were used to create “Tag Directory” using makeTagDirectory from Homer. Principal component (PC1) values were calculated by running runHiCpca.pl with default setting (resolution at 50 kb). This led to the separation of chromatin into two compartments, with positive PC1 regions reflecting “active” chromatin and negative PC1 regions indicative of “inactive” chromatin. Regions of continuous positive or negative PC1 values were stitched to be identified as A or B compartments, respectively. Genome-wide correlation of compartment PC1 values between mutant and WT cells was performed by running getHiCcorrDiff.pl from Homer. Flipped compartments (A to B, or B to A) were identified using Homer findHiCCompartments.pl. Genes involved in the flipped compartments were selected to analyse their mRNA expression changes by mutations.

Visualisation of gene expression, chromatin states, and 3-D DNA topology profiles

All sequencing tracks in this study were displayed on WashU Epigenome Browser78. For RNA-seq, ChIP-seq and ATAC-seq, bigwig files which were generated from their uniquely mapped reads were uploaded for visualisation. RNA-seq and ChIP-seq bigwig files were normalised to library size (as CPM), whereas ATAC-seq tracks were normalised to the total reads at ATAC-seq consensus peak set (also as CPM). Promoter-anchored interactions were visualised in the form of arcs, of which the color density was determined by CHiCAGO scores. Visualisation of chromatin compartment utilised bedGraph files which were generated using runHiCpca.pl from Homer. To visualize the enrichment of chromatin states (H3K4me1 or H3K27ac) or ATAC-seq at specific genomic regions, these signals were calculated into a matrix of enrichment scores using computeMatrix from deepTools79. Such a matrix was further illustrated by either a heatmap or a plot of average profile using plotHeatmap or plotProfile, respectively.

Evaluation of human homologous region to mouse Hoxa-LRSE

A human genomic region homologous to mouse Hoxa-LRSE was identified using the LIFTOVER tools (https://genome.ucsc.edu/cgi-bin/hgLiftOver, from GRCm38/mm10 to GRCh38/hg38) from UCSC Genome Browser. Chromatin states at this locus were demonstrated by ChIP-seq on BRD4 and H3K27ac in leukemia cell line OCI-AML3 cells (overexpress HOXA genes), H3K27ac ChIP-seq in Kasumi-1 cell line (with low expression of HOXA genes), and H3K4me1 ChIP-seq in human CD34+ HSPCs34,35. Chromatin accessibility assessed by mapping DNase Hypersenstivity Site (DHS-seq) or ATAC-seq was shown in NPM1 mutant AML patients (two with FLT3-ITD and two without) or in human CD34+ HSPCs, respectively37,38. In addition, promoter associated interactions were illustrated using pCHiC data from human CD34+ cells18, and CTCF ChIP-seq was included to visualize the CTCF occupancy in human CD34+ cells36.

Gene perturbation through shRNA-mediated knock-down

Npm1c/Flt3-ITD leukemia cells (termed DM cells) were derived by isolation of Lin-BM HSPCs from double mutant mice and were cultured in X-Vivo medium (Lonza) in the presence of 10 ng/mL mIL-3, 10 ng/mL hIL-6, and 50 ng/mL mSCF (Peprotech). Lentiviral shRNA plasmids were obtained from Sigma-Aldrich in the form of MISSION pLKO.1-puro vector. Lentiviral particles were produced by co-transfection of shRNA plasmids with psPAX and pMDG.2 in 293T cells using the Trans-IT LT-1 transfection reagent (Mirus). DM cells were infected by shRNA lentivirus twice via spinoculation in the presence of 5 μg/mL polybrene (Sigma-Aldrich). 72 hours post transduction, cells were selected with 2 μg/mL puromycin (Sigma-Aldrich) for 72 hours.

Disruption of cis-regulatory elements through CRISPR-mediated deletion

Mouse leukemia cells carrying Npm1c/Flt3-ITD/Cas9 (termed DM-Cas9 cells) were derived and cultured ex vivo as previously described79. In the presence of Cas9, using dual guide RNAs (gRNAs) to specifically target two genomic loci has been demonstrated previously50. Paired dual gRNAs targeting specific cis-regulatory elements or scramble control dual gRNAs were designed with IDT web tools (https://eu.idtdna.com/site/order/designtool/index/CRISPR_CUSTOM). Oligos of dual gRNAs were cloned into lentivirual vector pKLV2.2-h7SKgRNA5(SapI)-hU6gRNA5(BbsI)-PGKpuroBFP-W (#72666, Addgene). Production of gRNA particles and transduction of DM-Cas9 cells were carried out same as for shRNA. To confirm the target deletion in the bulk transduced cells, PCR primers were designed to amplify the wildtype or modified allele. Size of PCR products was checked by agarose gel electrophoresis. Furthermore, PCR products were subject to TA cloning (kit form Promega). Plasmid DNA was extracted from individual colonies and was confirmed for deletion by Sanger sequencing.

Cell proliferation assay

Leukemia cells expressing shRNAs or Cas9 plus dual gRNAs were cultured in the same liquid condition. After puromycin selection, viable cells were seeded at a concentration of 2×10^5 cells/mL in X-Vivo medium supplemented with cytokines and incubated at 37 °C with 5% CO2. Cells were counted periodically using hemocytometer and trypan blue, and the culture medium was replaced every two (for DM cells) or three days (for DM-Cas9 cells). Three biological replicates were performed independently.

Colony forming unit (CFU) assay

For DM cells with the expression of shRNA, post puromycin selection, cells were plated at concentrations of 1,000 cells per plate in duplicate on methylcellulose medium (MethoCult GF M3434, STEMCELL Technologies). Colonies were scored manually after 8-10 days. Three independent biological replicates were performed. For DM-Cas9 cells with the expression of Cas9 and dual gRNAs, single cell derived clones were generated by plating puromycin-selected bulk cells on M3434 methylcellulose medium for 7 days growth, and then transferring single colonies into liquid culture for another 2-3 weeks growth in the presence of 1 μg/mL puromycin. Single cell derived clones were than plated on methylcellulose medium following the same procedure as DM cells expressing shRNA.

Quantitative real-time PCR (qRT-PCR)

Total RNA was isolated using an AllPrep DNA/RNA Mini kit or RNeasy Plus Micro kit (Qiagen). cDNA was prepared using the SuperScript III Reverse transcription Kit (Invitrogen). qRT-PCR was performed on 1:10 diluted cDNA using either SYBR-containing Brilliant III Ultra-Fast QPCR Master Mix (Agilent Technologies) and gene-specific primers, or TaqMan Universal Master Mix II plus pre-designed assay oligos, and running on an MX3000p qPCR system (Agilent Technologies) with standard cycling setup.

Data availability

All sequencing raw data, normalised bigwig tracks for RNA-seq, ChIP-seq and ATAC-seq will be available at GEO database upon acceptance for peer reviewed publication. All supporting data derived from the sequencing analysis to assist understanding of the results and discussions in the paper were provided as supplementary data.

References

  1. 1.↵
    Maston, G.A., Evans, S.K. & Green, M.R. Transcriptional regulatory elements in the human genome. Annu Rev Genomics Hum Genet 7, 29–59 (2006).
    OpenUrlCrossRefPubMedWeb of Science
  2. 2.↵
    Lagha, M., Bothma, J.P. & Levine, M. Mechanisms of transcriptional precision in animal development. Trends Genet 28, 409–16 (2012).
    OpenUrlCrossRefPubMedWeb of Science
  3. 3.↵
    Spitz, F. & Furlong, E.E. Transcription factors: from enhancer binding to developmental control. Nat Rev Genet 13, 613–26 (2012).
    OpenUrlCrossRefPubMed
  4. 4.↵
    Schoenfelder, S. & Fraser, P. Long-range enhancer-promoter contacts in gene expression control. Nat Rev Genet 20, 437–455 (2019).
    OpenUrlCrossRef
  5. 5.↵
    Laurenti, E. & Gottgens, B. From haematopoietic stem cells to complex differentiation landscapes. Nature 553, 418–426 (2018).
    OpenUrlCrossRefPubMed
  6. 6.↵
    Huang, J. et al. Dynamic Control of Enhancer Repertoires Drives Lineage and Stage-Specific Transcription during Hematopoiesis. Dev Cell 36, 9–23 (2016).
    OpenUrlCrossRefPubMed
  7. 7.
    Orkin, S.H. & Zon, L.I. Hematopoiesis: an evolving paradigm for stem cell biology. Cell 132, 631–44 (2008).
    OpenUrlCrossRefPubMedWeb of Science
  8. 8.↵
    Cedar, H. & Bergman, Y. Epigenetics of haematopoietic cell development. Nat Rev Immunol 11, 478–88 (2011).
    OpenUrlCrossRefPubMed
  9. 9.↵
    Cancer Genome Atlas Research, N. et al. Genomic and epigenomic landscapes of adult de novo acute myeloid leukemia. N Engl J Med 368, 2059–74 (2013).
    OpenUrlCrossRefPubMedWeb of Science
  10. 10.
    Grimwade, D., Ivey, A. & Huntly, B.J. Molecular landscape of acute myeloid leukemia in younger adults and its clinical relevance. Blood 127, 29–41 (2016).
    OpenUrlAbstract/FREE Full Text
  11. 11.↵
    Papaemmanuil, E. et al. Genomic Classification and Prognosis in Acute Myeloid Leukemia. N Engl J Med 374, 2209–2221 (2016).
    OpenUrlCrossRefPubMed
  12. 12.↵
    Dohner, H., Weisdorf, D.J. & Bloomfield, C.D. Acute Myeloid Leukemia. N Engl J Med 373, 1136–52 (2015).
    OpenUrlCrossRefPubMed
  13. 13.↵
    Tenen, D.G. Disruption of differentiation in human cancer: AML shows the way. Nat Rev Cancer 3, 89–101 (2003).
    OpenUrlCrossRefPubMedWeb of Science
  14. 14.↵
    Lee, B.H. et al. FLT3 mutations confer enhanced proliferation and survival properties to multipotent progenitors in a murine model of chronic myelomonocytic leukemia. Cancer Cell 12, 367–80 (2007).
    OpenUrlCrossRefPubMedWeb of Science
  15. 15.↵
    Vassiliou, G.S. et al. Mutant nucleophosmin and cooperating pathways drive leukemia initiation and progression in mice. Nat Genet 43, 470–5 (2011).
    OpenUrlCrossRefPubMed
  16. 16.↵
    Mupo, A. et al. A powerful molecular synergy between mutant Nucleophosmin and Flt3-ITD drives acute myeloid leukemia in mice. Leukemia 27, 1917–20 (2013).
    OpenUrlCrossRefPubMed
  17. 17.↵
    Buenrostro, J.D., Giresi, P.G., Zaba, L.C., Chang, H.Y. & Greenleaf, W.J. Transposition of native chromatin for fast and sensitive epigenomic profiling of open chromatin, DNA-binding proteins and nucleosome position. Nat Methods 10, 1213–8 (2013).
    OpenUrlCrossRefPubMedWeb of Science
  18. 18.↵
    Mifsud, B. et al. Mapping long-range promoter contacts in human cells with high-resolution capture Hi-C. Nat Genet 47, 598–606 (2015).
    OpenUrlCrossRefPubMed
  19. 19.↵
    Calo, E. & Wysocka, J. Modification of enhancer chromatin: what, how, and why? Mol Cell 49, 825–37 (2013).
    OpenUrlCrossRefPubMedWeb of Science
  20. 20.↵
    Loven, J. et al. Selective inhibition of tumor oncogenes by disruption of super-enhancers. Cell 153, 320–34 (2013).
    OpenUrlCrossRefPubMedWeb of Science
  21. 21.↵
    Whyte, W.A. et al. Master transcription factors and mediator establish super-enhancers at key cell identity genes. Cell 153, 307–19 (2013).
    OpenUrlCrossRefPubMedWeb of Science
  22. 22.↵
    Bell, O., Tiwari, V.K., Thoma, N.H. & Schubeler, D. Determinants and dynamics of genome accessibility. Nat Rev Genet 12, 554–64 (2011).
    OpenUrlCrossRefPubMed
  23. 23.↵
    Snow, J.W. et al. Context-dependent function of “GATA switch” sites in vivo. Blood 117, 4769–72 (2011).
    OpenUrlAbstract/FREE Full Text
  24. 24.↵
    Yamazaki, H. et al. A remote GATA2 hematopoietic enhancer drives leukemogenesis in inv(3)(q21;q26) by activating EVI1 expression. Cancer Cell 25, 415–27 (2014).
    OpenUrlCrossRefPubMedWeb of Science
  25. 25.↵
    Heinz, S. et al. Simple combinations of lineage-determining transcription factors prime cis-regulatory elements required for macrophage and B cell identities. Mol Cell 38, 576–89 (2010).
    OpenUrlCrossRefPubMedWeb of Science
  26. 26.↵
    Lieberman-Aiden, E. et al. Comprehensive mapping of long-range interactions reveals folding principles of the human genome. Science 326, 289–93 (2009).
    OpenUrlAbstract/FREE Full Text
  27. 27.↵
    Tischkowitz, M. et al. Bi-allelic silencing of the Fanconi anaemia gene FANCF in acute myeloid leukaemia. Br J Haematol 123, 469–71 (2003).
    OpenUrlCrossRefPubMedWeb of Science
  28. 28.↵
    Freire-Pritchett, P. et al. Global reorganisation of cis-regulatory units upon lineage commitment of human embryonic stem cells. Elife 6(2017).
  29. 29.↵
    Rubin, A.J. et al. Lineage-specific dynamic and pre-established enhancer-promoter contacts cooperate in terminal differentiation. Nat Genet 49, 1522–1528 (2017).
    OpenUrlCrossRefPubMed
  30. 30.↵
    Li, Y. et al. Regulation of the PU.1 gene by distal elements. Blood 98, 2958–65 (2001).
    OpenUrlAbstract/FREE Full Text
  31. 31.
    Okuno, Y. et al. Potential autoregulation of transcription factor PU.1 by an upstream regulatory element. Mol Cell Biol 25, 2832–45 (2005).
    OpenUrlAbstract/FREE Full Text
  32. 32.↵
    Rosenbauer, F. et al. Lymphoid cell growth and transformation are suppressed by a key regulatory element of the gene encoding PU.1. Nat Genet 38, 27–37 (2006).
    OpenUrlCrossRefPubMed
  33. 33.↵
    Will, B. et al. Minimal PU.1 reduction induces a preleukemic state and promotes development of acute myeloid leukemia. Nat Med 21, 1172–81 (2015).
    OpenUrlCrossRef
  34. 34.↵
    Dawson, M.A. et al. Recurrent mutations, including NPM1c, activate a BRD4-dependent core transcriptional program in acute myeloid leukemia. Leukemia 28, 311–20 (2014).
    OpenUrlCrossRefPubMedWeb of Science
  35. 35.↵
    Gerlach, D. et al. The novel BET bromodomain inhibitor BI 894999 represses super-enhancer-associated transcription and synergizes with CDK9 inhibition in AML. Oncogene 37, 2687–2701 (2018).
    OpenUrl
  36. 36.↵
    Hu, G. et al. Regulation of nucleosome landscape and transcription factor targeting at tissue-specific enhancers by BRG1. Genome Res 21, 1650–8 (2011).
    OpenUrlAbstract/FREE Full Text
  37. 37.↵
    Assi, S.A. et al. Subtype-specific regulatory network rewiring in acute myeloid leukemia. Nat Genet 51, 151–162 (2019).
    OpenUrl
  38. 38.↵
    Cauchy, P. et al. Chronic FLT3-ITD Signaling in Acute Myeloid Leukemia Is Connected to a Specific Chromatin Signature. Cell Rep 12, 821–36 (2015).
    OpenUrlCrossRefPubMed
  39. 39.↵
    Dovey, O.M. et al. Molecular synergy underlies the co-occurrence patterns and phenotype of NPM1-mutant acute myeloid leukemia. Blood 130, 1911–1922 (2017).
    OpenUrlAbstract/FREE Full Text
  40. 40.↵
    Lara-Astiaso, D. et al. Immunogenetics. Chromatin state dynamics during blood formation. Science 345, 943–9 (2014).
    OpenUrlAbstract/FREE Full Text
  41. 41.↵
    Daugherty, B.L. et al. Cloning, expression, and characterization of the human eosinophil eotaxin receptor. J Exp Med 183, 2349–54 (1996).
    OpenUrlAbstract/FREE Full Text
  42. 42.↵
    Hughes, C.E. & Nibbs, R.J.B. A guide to chemokines and their receptors. FEBS J 285, 2944–2971 (2018).
    OpenUrlCrossRefPubMed
  43. 43.↵
    Foletta, V.C., Segal, D.H. & Cohen, D.R. Transcriptional regulation in the immune system: all roads lead to AP-1. J Leukoc Biol 63, 139–52 (1998).
    OpenUrlCrossRefPubMedWeb of Science
  44. 44.↵
    Carotta, S., Wu, L. & Nutt, S.L. Surprising new roles for PU.1 in the adaptive immune response. Immunol Rev 238, 63–75 (2010).
    OpenUrlCrossRefPubMed
  45. 45.↵
    Voon, D.C., Hor, Y.T. & Ito, Y. The RUNX complex: reaching beyond haematopoiesis into immunity. Immunology 146, 523–36 (2015).
    OpenUrlCrossRef
  46. 46.↵
    Gadad, S.S. et al. The multifunctional protein nucleophosmin (NPM1) is a human linker histone H1 chaperone. Biochemistry 50, 2780–9 (2011).
    OpenUrlCrossRefPubMedWeb of Science
  47. 47.↵
    Nemeth, A. et al. Initial genomics of the human nucleolus. PLoS Genet 6, e1000889 (2010).
    OpenUrlCrossRefPubMed
  48. 48.↵
    van Koningsbruggen, S. et al. High-resolution whole-genome sequencing reveals that specific chromatin domains from most human chromosomes associate with nucleoli. Mol Biol Cell 21, 3735–48 (2010).
    OpenUrlAbstract/FREE Full Text
  49. 49.↵
    Yan, J. et al. Histone H3 lysine 4 monomethylation modulates long-range chromatin interactions at enhancers. Cell Res 28, 204–220 (2018).
    OpenUrlCrossRefPubMed
  50. 50.↵
    Tzelepis, K. et al. A CRISPR Dropout Screen Identifies Genetic Vulnerabilities and Therapeutic Targets in Acute Myeloid Leukemia. Cell Rep 17, 1193–1205 (2016).
    OpenUrl
  51. 51.↵
    Wang, T. et al. Gene Essentiality Profiling Reveals Gene Networks and Synthetic Lethal Interactions with Oncogenic Ras. Cell 168, 890–903 e15 (2017).
    OpenUrlCrossRefPubMed
  52. 52.↵
    McKenzie, M.D. et al. Interconversion between Tumorigenic and Differentiated States in Acute Myeloid Leukemia. Cell Stem Cell 25, 258–272 e9 (2019).
    OpenUrl
  53. 53.↵
    Rosenbauer, F., Koschmieder, S., Steidl, U. & Tenen, D.G. Effect of transcription-factor concentrations on leukemic stem cells. Blood 106, 1519–24 (2005).
    OpenUrlAbstract/FREE Full Text
  54. 54.↵
    Bonadies, N., Pabst, T. & Mueller, B.U. Heterozygous deletion of the PU.1 locus in human AML. Blood 115, 331–4 (2010).
    OpenUrlAbstract/FREE Full Text
  55. 55.↵
    Huang, G. et al. The ability of MLL to bind RUNX1 and methylate H3K4 at PU.1 regulatory regions is impaired by MDS/AML-associated RUNX1/AML1 mutations. Blood 118, 6544–52 (2011).
    OpenUrlAbstract/FREE Full Text
  56. 56.↵
    Vangala, R.K. et al. The myeloid master regulator transcription factor PU.1 is inactivated by AML1-ETO in t(8;21) myeloid leukemia. Blood 101, 270–7 (2003).
    OpenUrlAbstract/FREE Full Text
  57. 57.↵
    Bell, C.C. et al. Targeting enhancer switching overcomes non-genetic drug resistance in acute myeloid leukaemia. Nat Commun 10, 2723 (2019).
    OpenUrlCrossRef
  58. 58.↵
    Cusan, M. et al. LSD1 inhibition exerts its antileukemic effect by recommissioning PU.1- and C/EBPalpha-dependent enhancers in AML. Blood 131, 1730–1742 (2018).
    OpenUrlAbstract/FREE Full Text
  59. 59.↵
    Gozdecka, M. et al. UTX-mediated enhancer and chromatin remodeling suppresses myeloid leukemogenesis through noncatalytic inverse regulation of ETS and GATA programs. Nat Genet 50, 883–894 (2018).
    OpenUrlCrossRef
  60. 60.↵
    Behre, G. et al. Meropenem monotherapy versus combination therapy with ceftazidime and amikacin for empirical treatment of febrile neutropenic patients. Ann Hematol 76, 73–80 (1998).
    OpenUrlCrossRefPubMedWeb of Science
  61. 61.↵
    Grondin, B. et al. c-Jun homodimers can function as a context-specific coactivator. Mol Cell Biol 27, 2919–33 (2007).
    OpenUrlAbstract/FREE Full Text
  62. 62.↵
    Kurotaki, D. et al. Transcription Factor IRF8 Governs Enhancer Landscape Dynamics in Mononuclear Phagocyte Progenitors. Cell Rep 22, 2628–2641 (2018).
    OpenUrl
  63. 63.↵
    Mancino, A. et al. A dual cis-regulatory code links IRF8 to constitutive and inducible gene expression in macrophages. Genes Dev 29, 394–408 (2015).
    OpenUrlAbstract/FREE Full Text
  64. 64.↵
    Huang, Y. et al. Identification and characterization of Hoxa9 binding sites in hematopoietic cells. Blood 119, 388–98 (2012).
    OpenUrlAbstract/FREE Full Text
  65. 65.↵
    Zhou, J. et al. PU.1 is essential for MLL leukemia partially via crosstalk with the MEIS/HOX pathway. Leukemia 28, 1436–48 (2014).
    OpenUrlCrossRefPubMed
  66. 66.↵
    Neijts, R. & Deschamps, J. At the base of colinear Hox gene expression: cis-features and trans-factors orchestrating the initial phase of Hox cluster activation. Dev Biol 428, 293–299 (2017).
    OpenUrl
  67. 67.↵
    Noordermeer, D. et al. The dynamic architecture of Hox gene clusters. Science 334, 222–5 (2011).
    OpenUrlAbstract/FREE Full Text
  68. 68.↵
    Horton, S.J. et al. Early loss of Crebbp confers malignant stem cell properties on lymphoid progenitors. Nat Cell Biol 19, 1093–1104 (2017).
    OpenUrlCrossRef
  69. 69.↵
    Dobin, A. et al. STAR: ultrafast universal RNA-seq aligner. Bioinformatics 29, 15–21 (2013).
    OpenUrlCrossRefPubMedWeb of Science
  70. 70.↵
    Ross-Innes, C.S. et al. Differential oestrogen receptor binding is associated with clinical outcome in breast cancer. Nature 481, 389–93 (2012).
    OpenUrlCrossRefPubMedWeb of Science
  71. 71.↵
    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).
    OpenUrlCrossRefPubMed
  72. 72.↵
    Subramanian, A. et al. Gene set enrichment analysis: a knowledge-based approach for interpreting genome-wide expression profiles. Proc Natl Acad Sci U S A 102, 15545–50 (2005).
    OpenUrlAbstract/FREE Full Text
  73. 73.↵
    Langmead, B., Trapnell, C., Pop, M. & Salzberg, S.L. Ultrafast and memory-efficient alignment of short DNA sequences to the human genome. Genome Biol 10, R25 (2009).
    OpenUrlCrossRefPubMed
  74. 74.↵
    Huang, J. et al. Dissecting super-enhancer hierarchy based on chromatin interactions. Nat Commun 9, 943 (2018).
    OpenUrlCrossRef
  75. 75.↵
    Schoenfelder, S. et al. The pluripotent regulatory circuitry connecting promoters to their long-range interacting elements. Genome Res 25, 582–97 (2015).
    OpenUrlAbstract/FREE Full Text
  76. 76.↵
    Wingett, S. et al. HiCUP: pipeline for mapping and processing Hi-C data. F1000Res 4, 1310 (2015).
    OpenUrl
  77. 77.↵
    Cairns, J. et al. CHiCAGO: robust detection of DNA looping interactions in Capture Hi-C data. Genome Biol 17, 127 (2016).
    OpenUrlCrossRef
  78. 78.↵
    Zhang, Y. et al. Model-based analysis of ChIP-Seq (MACS). Genome Biol 9, R137 (2008).
    OpenUrlCrossRefPubMed
  79. 79.↵
    Ramirez, F. et al. deepTools2: a next generation web server for deep-sequencing data analysis. Nucleic Acids Res 44, W160–5 (2016).
    OpenUrlCrossRefPubMed
Back to top
PreviousNext
Posted March 20, 2020.
Download PDF

Supplementary Material

Email

Thank you for your interest in spreading the word about bioRxiv.

NOTE: Your email address is requested solely to identify you as the sender of this article.

Enter multiple addresses on separate lines or separate them with commas.
Mutational synergy coordinately remodels chromatin accessibility, enhancer landscape and 3-Dimensional DNA topology to alter gene expression during leukemia induction
(Your Name) has forwarded a page to you from bioRxiv
(Your Name) thought you would like to see this page from the bioRxiv website.
CAPTCHA
This question is for testing whether or not you are a human visitor and to prevent automated spam submissions.
Share
Mutational synergy coordinately remodels chromatin accessibility, enhancer landscape and 3-Dimensional DNA topology to alter gene expression during leukemia induction
Haiyang Yun, Shabana Vohra, Annalisa Mupo, George Giotopoulos, Daniel Sasca, Sarah J. Horton, Shuchi Agrawal-Singh, Eshwar Meduri, Faisal Basheer, Ludovica Marando, Malgorzata Gozdecka, Oliver M. Dovey, Aracely Castillo-Venzor, Xiaonan Wang, Paolo Gallipoli, Carsten Müller-Tidow, Cameron S. Osborne, George S. Vassiliou, Brian J. P. Huntly
bioRxiv 2020.03.20.000059; doi: https://doi.org/10.1101/2020.03.20.000059
Digg logo Reddit logo Twitter logo Facebook logo Google logo LinkedIn logo Mendeley logo
Citation Tools
Mutational synergy coordinately remodels chromatin accessibility, enhancer landscape and 3-Dimensional DNA topology to alter gene expression during leukemia induction
Haiyang Yun, Shabana Vohra, Annalisa Mupo, George Giotopoulos, Daniel Sasca, Sarah J. Horton, Shuchi Agrawal-Singh, Eshwar Meduri, Faisal Basheer, Ludovica Marando, Malgorzata Gozdecka, Oliver M. Dovey, Aracely Castillo-Venzor, Xiaonan Wang, Paolo Gallipoli, Carsten Müller-Tidow, Cameron S. Osborne, George S. Vassiliou, Brian J. P. Huntly
bioRxiv 2020.03.20.000059; doi: https://doi.org/10.1101/2020.03.20.000059

Citation Manager Formats

  • BibTeX
  • Bookends
  • EasyBib
  • EndNote (tagged)
  • EndNote 8 (xml)
  • Medlars
  • Mendeley
  • Papers
  • RefWorks Tagged
  • Ref Manager
  • RIS
  • Zotero
  • Tweet Widget
  • Facebook Like
  • Google Plus One

Subject Area

  • Cancer Biology
Subject Areas
All Articles
  • Animal Behavior and Cognition (3592)
  • Biochemistry (7562)
  • Bioengineering (5508)
  • Bioinformatics (20762)
  • Biophysics (10309)
  • Cancer Biology (7967)
  • Cell Biology (11627)
  • Clinical Trials (138)
  • Developmental Biology (6602)
  • Ecology (10190)
  • Epidemiology (2065)
  • Evolutionary Biology (13594)
  • Genetics (9532)
  • Genomics (12834)
  • Immunology (7917)
  • Microbiology (19525)
  • Molecular Biology (7651)
  • Neuroscience (42027)
  • Paleontology (307)
  • Pathology (1254)
  • Pharmacology and Toxicology (2196)
  • Physiology (3263)
  • Plant Biology (7029)
  • Scientific Communication and Education (1294)
  • Synthetic Biology (1949)
  • Systems Biology (5422)
  • Zoology (1114)