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:

Epigenetic programming defines haematopoietic stem cell fate restriction

Abstract

Haematopoietic stem cells (HSCs) are multipotent, but individual HSCs can show restricted lineage output in vivo. Currently, the molecular mechanisms and physiological role of HSC fate restriction remain unknown. Here we show that lymphoid fate is epigenetically but not transcriptionally primed in HSCs. In multi-lineage HSCs that produce lymphocytes, lymphoid-specific upstream regulatory elements (LymUREs) but not promoters are preferentially accessible compared with platelet-biased HSCs that do not produce lymphoid cell types, providing transcriptionally silent lymphoid lineage priming. Runx3 is preferentially expressed in multi-lineage HSCs, and reinstating Runx3 expression increases LymURE accessibility and lymphoid-primed multipotent progenitor 4 (MPP4) output in old, platelet-biased HSCs. In contrast, platelet-biased HSCs show elevated levels of epigenetic platelet-lineage priming and give rise to MPP2 progenitors with molecular platelet bias. These MPP2 progenitors generate platelets with faster kinetics and through a more direct cellular pathway compared with MPP2s derived from multi-lineage HSCs. Epigenetic programming therefore predicts both fate restriction and differentiation kinetics in HSCs.

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: Fate-restricted HSC clones are internally coherent and molecularly distinct.
Fig. 2: Identification of transcriptional and epigenetic programs of fate-restricted HSCs.
Fig. 3: HSC fate is predicted by epigenetic priming.
Fig. 4: Lymphoid gene expression during HSC commitment is driven by primed enhancers.
Fig. 5: Epigenetic priming identifies an increased abundance of PLT HSCs in aged mouse bone marrow.
Fig. 6: Ectopic expression of Runx3 rescues epigenetic lymphoid priming and corrects Meg bias in aged HSCs.
Fig. 7: Distinct MPP2 pathways arise from PLT and MUL HSCs.
Fig. 8: Accelerated platelet production by PLT HSC-derived MPP2s.

Similar content being viewed by others

Data availability

All of the raw and processed high-throughput sequencing data generated in this study have been deposited in the Gene Expression Omnibus database under accession number GSE188119. Public datasets used in this study include the JASPAR 2020 motif database52, RNA-seq of HSCs/MPPs21, available in the ArrayExpress database (http://www.ebi.ac.uk/arrayexpress) under accession number E-MTAB-2262, and haematopoietic progenitors24, available in the Haemospere database (https://www.haemosphere.org/). All other data generated in this study are available upon reasonable request. Requests for materials and manuscript correspondence should be directed to the corresponding author. Source data are provided with this paper.

References

  1. Muller-Sieburg, C. E., Cho, R. H., Thoman, M., Adkins, B. & Sieburg, H. B. Deterministic regulation of hematopoietic stem cell self-renewal and differentiation. Blood 100, 1302–1309 (2002).

    Article  CAS  PubMed  Google Scholar 

  2. Dykstra, B. et al. Long-term propagation of distinct hematopoietic differentiation programs in vivo. Cell Stem Cell 1, 218–229 (2007).

    Article  CAS  PubMed  Google Scholar 

  3. Yamamoto, R. et al. Clonal analysis unveils self-renewing lineage-restricted progenitors generated directly from hematopoietic stem cells. Cell 154, 1112–1126 (2013).

    Article  CAS  PubMed  Google Scholar 

  4. Carrelha, J. et al. Hierarchically related lineage-restricted fates of multipotent haematopoietic stem cells. Nature 554, 106–111 (2018).

    Article  CAS  PubMed  Google Scholar 

  5. Pei, W. et al. Resolving fates and single-cell transcriptomes of hematopoietic stem cell clones by PolyloxExpress barcoding. Cell Stem Cell 27, 383–395.e8 (2020).

    Article  CAS  PubMed  Google Scholar 

  6. Rodriguez-Fraticelli, A. E. et al. Single-cell lineage tracing unveils a role for TCF15 in haematopoiesis. Nature 583, 585–589 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  7. Jacobsen, S. E. W. & Nerlov, C. Haematopoiesis in the era of advanced single-cell technologies. Nat. Cell Biol. 21, 2–8 (2019).

    Article  CAS  PubMed  Google Scholar 

  8. Yu, V. W. C. et al. Epigenetic memory underlies cell-autonomous heterogeneous behavior of hematopoietic stem cells. Cell 167, 1310–1322 (2016).

    Article  CAS  PubMed  Google Scholar 

  9. Picelli, S. et al. Full-length RNA-seq from single cells using Smart-seq2. Nat. Protoc. 9, 171–181 (2014).

    Article  CAS  PubMed  Google Scholar 

  10. Subramanian, A. et al. Gene set enrichment analysis: a knowledge-based approach for interpreting genome-wide expression profiles. Proc. Natl Acad. Sci. USA 102, 15545–15550 (2005).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  11. Wilson, N. K. et al. Combined single-cell functional and gene expression analysis resolves heterogeneity within stem cell populations. Cell Stem Cell 16, 712–724 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  12. Pietras, E. M. et al. Re-entry into quiescence protects hematopoietic stem cells from the killing effect of chronic exposure to type I interferons. J. Exp. Med. 211, 245–262 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  13. Cabezas-Wallscheid, N. et al. Vitamin A–retinoic acid signaling regulates hematopoietic stem cell dormancy. Cell 169, 807–823.e19 (2017).

    Article  CAS  PubMed  Google Scholar 

  14. Lauridsen, F. K. B. et al. Differences in cell cycle status underlie transcriptional heterogeneity in the HSC compartment. Cell Rep. 24, 766–780 (2018).

    Article  CAS  PubMed  Google Scholar 

  15. Tong, J. et al. Hematopoietic stem cell heterogeneity is linked to the initiation and therapeutic response of myeloproliferative neoplasms. Cell Stem Cell 28, 502–513.e6 (2021).

    Article  CAS  PubMed  Google Scholar 

  16. Schep, A. N., Wu, B., Buenrostro, J. D. & Greenleaf, W. J. chromVAR: inferring transcription-factor-associated accessibility from single-cell epigenomic data. Nat. Methods 14, 975–978 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  17. Mansson, R. et al. Molecular evidence for hierarchical transcriptional lineage priming in fetal and adult stem cells and multipotent progenitors. Immunity 26, 407–419 (2007).

    Article  PubMed  Google Scholar 

  18. Grover, A. et al. Single-cell RNA sequencing reveals molecular and functional platelet bias of aged haematopoietic stem cells. Nat. Commun. 7, 11075 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  19. Bereshchenko, O. et al. Hematopoietic stem cell expansion precedes the generation of committed myeloid leukemia-initiating cells in C/EBPα mutant AML. Cancer Cell 16, 390–400 (2009).

    Article  CAS  PubMed  Google Scholar 

  20. Stojnic, R. & Diez, D. PWMEnrich: PWM enrichment analysis. R package version 4.30.0 https://doi.org/10.18129/B9.bioc.PWMEnrich (2021).

  21. Cabezas-Wallscheid, N. et al. Identification of regulatory networks in HSCs and their immediate progeny via integrated proteome, transcriptome, and DNA methylome analysis. Cell Stem Cell 15, 507–522 (2014).

    Article  CAS  PubMed  Google Scholar 

  22. Pietras, E. M. et al. Functionally distinct subsets of lineage-biased multipotent progenitors control blood production in normal and regenerative conditions. Cell Stem Cell 17, 35–46 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  23. Adolfsson, J. et al. Identification of Flt3+ lympho-myeloid stem cells lacking erythro-megakaryocytic potential: a revised road map for adult blood lineage commitment. Cell 121, 295–306 (2005).

    Article  CAS  PubMed  Google Scholar 

  24. Choi, J. et al. Haemopedia RNA-seq: a database of gene expression during haematopoiesis in mice and humans. Nucleic Acids Res. 47, D780–D785 (2019).

    Article  CAS  PubMed  Google Scholar 

  25. Benz, C. et al. Hematopoietic stem cell subtypes expand differentially during development and display distinct lymphopoietic programs. Cell Stem Cell 10, 273–283 (2012).

    Article  CAS  PubMed  Google Scholar 

  26. Newman, A. M. et al. Determining cell type abundance and expression from bulk tissues with digital cytometry. Nat. Biotechnol. 37, 773–782 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  27. Young, K. et al. Progressive alterations in multipotent hematopoietic progenitors underlie lymphoid cell loss in aging. J. Exp. Med. 213, 2259–2267 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  28. Wong, W. F. et al. Over-expression of Runx1 transcription factor impairs the development of thymocytes from the double-negative to double-positive stages. Immunology 130, 243–253 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  29. Menezes, A. C. et al. RUNX3 overexpression inhibits normal human erythroid development. Sci. Rep. 12, 1243 (2022).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  30. Drissen, R. et al. Distinct myeloid progenitor-differentiation pathways identified through single-cell RNA sequencing. Nat. Immunol. 17, 666–676 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  31. Madisen, L. et al. A robust and high-throughput Cre reporting and characterization system for the whole mouse brain. Nat. Neurosci. 13, 133–140 (2010).

    Article  CAS  PubMed  Google Scholar 

  32. Fujii, Y. et al. A novel mechanism of thrombocytopenia by PS exposure through TMEM16F in sphingomyelin synthase 1 deficiency. Blood Adv. 5, 4265–4277 (2021).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  33. Morcos, M. N. F. et al. Fate mapping of hematopoietic stem cells reveals two pathways of native thrombopoiesis. Nat. Commun. 13, 4504 (2022).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  34. Noah, T. K., Donahue, B. & Shroyer, N. F. Intestinal development and differentiation. Exp. Cell. Res. 317, 2702–2710 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  35. Taupin, P. & Gage, F. H. Adult neurogenesis and neural stem cells of the central nervous system in mammals. J. Neurosci. Res. 69, 745–749 (2002).

    Article  CAS  PubMed  Google Scholar 

  36. Blanpain, C. & Fuchs, E. Epidermal homeostasis: a balancing act of stem cells in the skin. Nat. Rev. Mol. Cell Biol. 10, 207–217 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  37. Kim, D. et al. TopHat2: accurate alignment of transcriptomes in the presence of insertions, deletions and gene fusions. Genome Biol. 14, R36 (2013).

    Article  PubMed  PubMed Central  Google Scholar 

  38. Liao, Y., Smyth, G. K. & Shi, W. featureCounts: an efficient general purpose program for assigning sequence reads to genomic features. Bioinformatics 30, 923–930 (2014).

    Article  CAS  PubMed  Google Scholar 

  39. Rodriguez-Meira, A. et al. Unravelling intratumoral heterogeneity through high-sensitivity single-cell mutational analysis and parallel RNA sequencing. Mol. Cell 73, 1292–1305.e8 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  40. Corces, M. R. et al. An improved ATAC-seq protocol reduces background and enables interrogation of frozen tissues. Nat. Methods 14, 959–962 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  41. Bolger, A. M., Lohse, M. & Usadel, B. Trimmomatic: a flexible trimmer for Illumina sequence data. Bioinformatics 30, 2114–2120 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

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

    Article  CAS  PubMed  Google Scholar 

  43. Ewels, P., Magnusson, M., Lundin, S. & Kaller, M. MultiQC: summarize analysis results for multiple tools and samples in a single report. Bioinformatics 32, 3047–3048 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

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

    Article  PubMed  PubMed Central  Google Scholar 

  45. Subramanian, A., Kuehn, H., Gould, J., Tamayo, P. & Mesirov, J. P. GSEA-P: a desktop application for gene set enrichment analysis. Bioinformatics 23, 3251–3253 (2007).

    Article  CAS  PubMed  Google Scholar 

  46. Barbie, D. A. et al. Systematic RNA interference reveals that oncogenic KRAS-driven cancers require TBK1. Nature 462, 108–112 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  47. 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 

  48. Li, H. et al. The Sequence Alignment/Map format and SAMtools. Bioinformatics 25, 2078–2079 (2009).

    Article  PubMed  PubMed Central  Google Scholar 

  49. Ou, J. H. et al. ATACseqQC: a Bioconductor package for post-alignment quality assessment of ATAC-seq data. BMC Genomics 19, 169 (2018).

    Article  PubMed  PubMed Central  Google Scholar 

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

    Article  PubMed  PubMed Central  Google Scholar 

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

    Article  CAS  PubMed  Google Scholar 

  52. Fornes, O. et al. JASPAR 2020: update of the open-access database of transcription factor binding profiles. Nucleic Acids Res. 48, D87–D92 (2020).

    CAS  PubMed  Google Scholar 

  53. Bentsen, M. et al. ATAC-seq footprinting unravels kinetics of transcription factor binding during zygotic genome activation. Nat. Commun. 11, 4267 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  54. Chen, X., Miragaia, R. J., Natarajan, K. N. & Teichmann, S. A. A rapid and robust method for single cell chromatin accessibility profiling. Nat. Commun. 9, 5345 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  55. Di Genua, C. et al. C/EBPα and GATA-2 mutations induce bilineage acute erythroid leukemia through transformation of a neomorphic neutrophil-erythroid progenitor. Cancer Cell 37, 690–704.e8 (2020).

    Article  PubMed  PubMed Central  Google Scholar 

  56. Valletta, S. et al. Micro-environmental sensing by bone marrow stroma identifies IL-6 and TGFβ1 as regulators of haematopoietic ageing. Nat. Commun. 11, 4075 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

Download references

Acknowledgements

We thank the Weatherall Institute of Molecular Medicine (WIMM) FACS facility for assistance with cell sorting and the WIMM Computational Biology Research Group for computational support. This research was funded in whole or in part by Biotechnology and Biological Sciences Research Council project grant BB/V002198/1 and Medical Research Council (MRC) Unit grants MC_UU_12009/7 and MC_UU_00029/9 (to C.N.) and MRC Unit grant MC_UU_12009/5 and Swedish Research Council grant 538-2013-8995 (to S.E.J.). For the purpose of open access, the authors have applied a CC BY public copyright licence to any author accepted manuscript version arising from this submission. The WIMM FACS Core Facility is supported by the MRC Human Immunology Unit, MRC Molecular Haematology Unit (MC_UU_12009), National Institute for Health Research Oxford Biomedical Research Centre and John Fell Fund (131/030 and 101/517), EPA Cephalosporin fund (CF182 and CF170) and WIMM Strategic Alliance awards (G0902418 and MC_UU_12025).

Author information

Authors and Affiliations

Authors

Contributions

C.N. and S.E.J. conceived of the study and supervised the experiments. Y.M., J.C., R.D., X.R., B.Z., A.G. and S.V. performed the experiments. Y.M., J.C. and C.N. analysed the data. Y.M., B.Z., S.T. and C.N. performed the bioinformatics analysis. C.N. and Y.M. wrote the manuscript.

Corresponding author

Correspondence to Claus Nerlov.

Ethics declarations

Competing interests

The authors declare no competing interests.

Peer review

Peer review information

Nature Cell Biology thanks John Dick, Alejo Rodriguez-Fraticelli and the other, anonymous, reviewer(s) for their contribution to the peer review of this work. Peer reviewer reports are available.

Additional information

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

Extended data

Extended Data Fig. 1 Sorting of single HSCs for transplantation and molecular profiling.

a) Flow cytometry plots showing the gating strategy used for sorting of VT+ HSCs for single cell transplantation and single cell RNAseq. b) Flow cytometry plots showing the gating strategy used for sorting HSCs from single-cell transplanted mice for scRNAseq.

Source data

Extended Data Fig. 2 Expression of random forest genes in fate-restricted single HSCs.

UMAP plots showing the expression of the top 10 genes identified by random forest importance score in single HSCs from Fig. 1d.

Source data

Extended Data Fig. 3 Lineage-specific chromatin signatures.

a) Heatmap of promoter accessibility signatures specific for MkPs (Meg signature), CFU-Es (Ery), NMPs (Myl) and proB/DN3 cells (Lym). Each signature contains the top 1500 features identified as selectively accessible in each of the 4 cell populations/combined cell populations using DESeq2 (Padj<0.05 and |log2(fold change)| >1). N = 3 biological replicates for all populations. Scale shows row z-score of transformed gene expression levels. b) Heatmap of URE accessibility signatures specific for MkPs (Meg signature), CFU-Es (Ery), NMPs (Myl) and proB/DN3 cells (Lym). Each signature contains the top 1500 features identified as selectively accessible in each of the 4 cell populations/combined cell populations using DESeq2 (Padj<0.05 and |log2(fold change)| >1). N = 3 biological replicates for all populations. Scale shows row z-score of transformed chromatin accessibility levels. c) Enrichment of TF motifs differentially accessible in HSC subtypes (from Fig. 3a) in lymphoid-specific promoter peaks calculated using PWMEnrich-based sequence analysis. These values are those plotted along the Fig. 4a x-axis. d) Enrichment analysis of HSC-subtype TF motifs (from Fig. 2a) in Lym URE peaks calculated using PWMEnrich, corresponding to Fig. 4a y-axis.

Source data

Extended Data Fig. 4 Expression of lymphoid TFs in HSC subtypes and HSC signatures in old HSCs.

a) The expression of genes encoding B-cell specific (Ebf1 and Pax5) and T-cell specific (Lef1 and Tcf7) TFs in PLT-, PEM- and MUL-HSC. RPKM values from bulk RNAseq data are shown. Statistical analysis was performed using two-tailed Welch’s t-test. Biological replicates: PLT-HSC, N = 3; PEM-HSC, N = 6; MUL-HSC, N = 4. Mean values with SD are shown. NA indicates that all values in the comparison are zero. b) GSEA analysis comparing young and old HSCs using bulk RNAseq. Plots show the enrichment of transcriptional PLT-HSC and MUL-HSC signatures (from Supplementary Table 2) in old HSCs. P-values and normalised enrichment scores (NES) are shown. c) GSEA analysis comparing young and old HSCs using bulk ATAC. Plots show the enrichment of epigenetic PLT-HSC and MUL-HSC signatures (from Supplementary Table 3) in old HSCs. P-values and normalised enrichment scores (NES) are shown.

Source data

Extended Data Fig. 5 The role of RUNX factors in epigenetic lymphoid priming.

a) The expression of Runx1 and Runx2 in young and old HSCs. RPKM values from bulk RNAseq data are shown. Statistical analysis was performed using two-tailed Welch’s t-test (biological replciates: young, N = 4; old, N = 3). Mean values with SD are shown. b) The expression of Runx1 and Runx2 in PLT-, PEM- and MUL-HSC. RPKM values from bulk RNAseq data are shown. Statistical analysis was performed using two-tailed Welch’s t-test. Biological replicates: PLT-HSC, N = 3; PEM-HSC, N = 6; MUL-HSC, N = 4. Mean values with SD are shown. c) Schematic diagram of the Runx3-expressing lentiviral vector used for overexpressing RUNX3 in old HSCs and transplantation. The human CD2 (hCD2) surface marker was used to identify successfully transduced cells. d) Experimental workflow of transplantation of old HSCs transduced with mock or RUNX3-expressing lentivirus. e) Flow cytometry plots showing the gating strategy for identifying HSC, MPP2, MPP3 and MPP4 derived by donor HSCs transduced with mock or RUNX3-expressing lentivirus for cell sorting and analysis.

Source data

Extended Data Fig. 6 Platelet kinetics in Gata1-CreERT2 transgenic mice.

a) Schematic diagram of the Gata1-CreERT2 (GCET2) transgene used for lineage tracing of GE+ progenitors. b) Representative flow cytometry plots showing EGFP and tdTomato labelling of BM LK and LSK populations in control (No Tamoxifen) GCET2/Ai9 mice (top panels) and in GCET2/Ai9 mice 2 days after tamoxifen administration (bottom panels). c) Representative flow cytometry plots showing the gating strategy for the identification of peripheral blood platelets. d) Representative flow cytometry plots showing the extent of tdTomato labelling of platelets in control GE/GCET2/Ai9 mice (No Tamoxifen), and GE/GCET2/Ai9 mice 7 and 20 days after tamoxifen administration. e) Representative flow cytometry plot showing the identification of donor-derived Gata1-EGFP+/Rosa26-tdTomato+ (GE+RT+) and Gata1-EGFP+/Rosa26-tdTomato– (GE+RT–), as well as recipient-and competitor-derived Gata1-EGFP– (GE–) platelets in peripheral blood of a recipient mouse transplanted with a single GE/GCET2/Ai9 MUL-HSC at day 16 after Tamoxifen administration. The fraction of lineage traced platelets was calculated as GE+RT+/((GE+RT+)+(GE+RT–)).

Source data

Supplementary information

Reporting Summary

Peer Review File

Supplementary Tables

Supplementary Table 1. Cluster distribution of clonal single-HSC transcriptomes. For each clone, the number of cells associated with each cluster is shown, as in the total cell number, the number of cells in the most populated cluster and the probability of the distribution being random, using a binomial distribution test. Supplementary Table 2. Genes differentially regulated between HSC subtypes. Differential genes in bulk RNA-seq (Fig. 2a) were identified using the DEseq2 package (P < 0.05; log2[fold change] > 0.5). Regulons are defined by pairwise comparing of HSC subtypes and identification of genes selectively upregulating in a single subtype (for example, the PLT regulon is defined by PLT > PEM and PLT > MUL) or upregulated in two subtypes compared with the third subtype (for example, the PLT + PEM regulon is defined by PLT > MUL and PEM > MUL). Supplementary Table 3. Chromatin peaks differentially accessible between HSC subtypes. Differentially accessible peaks (P < 0.05; log2[fold change] > 0.5 using DEseq2) for each HSC subtype identified by bulk ATAC-seq (Fig. 2b). For each peak, the genomic coordinates and cell type are shown. Supplementary Table 4. Transcription factor motifs differentially enriched in HSC subtypes. Motifs significantly upregulated (P < 0.05 using chromVAR) in each of the indicated regulons from Fig. 3a are shown. Supplementary Table 5. Chromatin peaks differentially accessible in committed progenitors. Each signature contains the top 2,000 features identified as selectively accessible in each of the four individual or combined cell populations. For each peak, the genomic coordinates and cell type are shown. Supplementary Table 6. Promoter and URE peaks differentially accessible in committed progenitors. Each signature contains the top 1,500 promoter- and URE-associated features identified as selectively accessible in each of the four individual or combined cell populations. For each peak, the genomic coordinates, cell type and element type are shown. Supplementary Table 7. Lymphoid leading-edge genes. Nearest-neighbour genes for the LymURE leading-edge peaks from the comparison of PLT and MUL HSC ATAC-seq. Supplementary Table 8. List of antibodies used for flow cytometry analysis and cell sorting. For each antibody, the antigen, conjugated fluorophore, clone, supplier, catalogue number and application are shown. Supplementary Table 9. Metadata of single-cell-engrafted mice used for ATAC- and RNA-seq. For clones used for single-cell and bulk sequencing, the reconstitution pattern, engraftment level in the LSK CD150+CD48 HSC compartment, time point of HSC re-isolation after transplantation, cellular phenotype re-isolated and number of cells used for each type of analysis are shown. Supplementary Table 10. LymURE leading-edge genomic regions. Leading-edge genomic regions from the GSEA comparing the accessibility of LymURE peaks in PLT versus MUL HSCs (from Fig. 4c), young versus old HSCs (from Fig. 5d) and Runx3 versus control HSCs (from Fig. 6c). These were used for respective leading-edge analyses. For each peak, the genomic coordinates and comparison are shown.

Source data

Rights and permissions

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

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Meng, Y., Carrelha, J., Drissen, R. et al. Epigenetic programming defines haematopoietic stem cell fate restriction. Nat Cell Biol 25, 812–822 (2023). https://doi.org/10.1038/s41556-023-01137-5

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41556-023-01137-5

This article is cited by

Search

Quick links

Nature Briefing

Sign up for the Nature Briefing newsletter — what matters in science, free to your inbox daily.

Get the most important science stories of the day, free in your inbox. Sign up for Nature Briefing