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:

Lineage-specific and single-cell chromatin accessibility charts human hematopoiesis and leukemia evolution

Abstract

We define the chromatin accessibility and transcriptional landscapes in 13 human primary blood cell types that span the hematopoietic hierarchy. Exploiting the finding that the enhancer landscape better reflects cell identity than mRNA levels, we enable 'enhancer cytometry' for enumeration of pure cell types from complex populations. We identify regulators governing hematopoietic differentiation and further show the lineage ontogeny of genetic elements linked to diverse human diseases. In acute myeloid leukemia (AML), chromatin accessibility uncovers unique regulatory evolution in cancer cells with a progressively increasing mutation burden. Single AML cells exhibit distinctive mixed regulome profiles corresponding to disparate developmental stages. A method to account for this regulatory heterogeneity identified cancer-specific deviations and implicated HOX factors as key regulators of preleukemic hematopoietic stem cell characteristics. Thus, regulome dynamics can provide diverse insights into hematopoietic development and disease.

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

Figure 1: Interrogation of chromatin landscapes in primary blood cells.
Figure 2: Distal regulatory elements enable accurate classification of the hematopoietic hierarchy.
Figure 3: Enhancer cytometry allows for deconvolution of the hematopoietic hierarchy.
Figure 4: Integrative analysis of the hematopoietic regulome refines the transcriptional circuitry driving cell specification and enriches understanding of human disease.
Figure 5: Acute myeloid leukemia regulomes show cooption of normal myelopoiesis.
Figure 6: Enhancer cytometry and single-cell regulomes support a model of regulatory heterogeneity and allow for deconvolution of AML-specific biology.
Figure 7: Early chromatin accessibility alterations in pHSCs cause defects in differentiation that correlate with adverse patient outcomes.

Similar content being viewed by others

Accession codes

Primary accessions

Gene Expression Omnibus

Referenced accessions

Gene Expression Omnibus

Sequence Read Archive

References

  1. Quesenberry, P.J. & Colvin, G.A. in Williams Hematology 153–174 (McGraw-Hill, 2005).

  2. Ji, H. et al. Comprehensive methylome map of lineage commitment from haematopoietic progenitors. Nature 467, 338–342 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  3. Lara-Astiaso, D. et al. Chromatin state dynamics during blood formation. Science 345, 943–949 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  4. Chen, L. et al. Transcriptional diversity during lineage commitment of human blood progenitors. Science 345, 1251033 (2014).

    Article  PubMed  PubMed Central  Google Scholar 

  5. Novershtern, N. et al. Densely interconnected transcriptional circuits control cell states in human hematopoiesis. Cell 144, 296–309 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  6. 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, 1213–1218 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  7. Jin, W. et al. Genome-wide detection of DNase I hypersensitive sites in single cells and FFPE tissue samples. Nature 528, 142–146 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  8. Shih, A.H. et al. The role of mutations in epigenetic regulators in myeloid malignancies. Nat. Rev. Cancer 12, 599–612 (2012).

    Article  CAS  PubMed  Google Scholar 

  9. Genovese, G. et al. Clonal hematopoiesis and blood-cancer risk inferred from blood DNA sequence. N. Engl. J. Med. 371, 2477–2487 (2014).

    Article  PubMed  PubMed Central  Google Scholar 

  10. Jaiswal, S. et al. Age-related clonal hematopoiesis associated with adverse outcomes. N. Engl. J. Med. 371, 2488–2498 (2014).

    Article  PubMed  PubMed Central  Google Scholar 

  11. Jan, M. et al. Clonal evolution of preleukemic hematopoietic stem cells precedes human acute myeloid leukemia. Sci. Transl. Med. 4, 149ra118 (2012).

    Article  PubMed  PubMed Central  Google Scholar 

  12. Corces-Zimmerman, M.R., Hong, W.-J., Weissman, I.L., Medeiros, B.C. & Majeti, R. Preleukemic mutations in human acute myeloid leukemia affect epigenetic regulators and persist in remission. Proc. Natl. Acad. Sci. USA 111, 2548–2553 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  13. Shlush, L.I. et al. Identification of pre-leukaemic haematopoietic stem cells in acute leukaemia. Nature 506, 328–333 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  14. Majeti, R., Park, C.Y. & Weissman, I.L. Identification of a hierarchy of multipotent hematopoietic progenitors in human cord blood. Cell Stem Cell 1, 635–645 (2007).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  15. Manz, M.G., Miyamoto, T., Akashi, K. & Weissman, I.L. Prospective isolation of human clonogenic common myeloid progenitors. Proc. Natl. Acad. Sci. USA 99, 11872–11877 (2002).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  16. Kohn, L.A. et al. Lymphoid priming in human bone marrow begins before expression of CD10 with upregulation of L-selectin. Nat. Immunol. 13, 963–971 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  17. Seita, J. & Weissman, I.L. Hematopoietic stem cell: self-renewal versus differentiation. Wiley Interdiscip. Rev. Syst. Biol. Med. 2, 640–653 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  18. Roadmap Epigenetics Consortium. et al. Integrative analysis of 111 reference human epigenomes. Nature 518, 317–330 (2015).

  19. Manning, C.D., Raghavan, P. & Schutze, H. Introduction to Information Retrieval (Cambridge University Press, 2008).

  20. Cauchy, P. et al. Chronic FLT3–ITD signaling in acute myeloid leukemia is connected to a specific chromatin signature. Cell Rep. 12, 821–836 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  21. 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, 576–589 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  22. Newman, A.M. et al. Robust enumeration of cell subsets from tissue expression profiles. Nat. Methods 12, 453–457 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  23. Buenrostro, J.D. et al. Single-cell chromatin accessibility reveals principles of regulatory variation. Nature 523, 486–490 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  24. Weiss, M.J. & Orkin, S.H. GATA transcription factors: key regulators of hematopoiesis. Exp. Hematol. 23, 99–107 (1995).

    CAS  PubMed  Google Scholar 

  25. Burns, C.E., Traver, D., Mayhall, E., Shepard, J.L. & Zon, L.I. Hematopoietic stem cell fate is established by the Notch–Runx pathway. Genes Dev. 19, 2331–2342 (2005).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  26. Nerlov, C. & Graf, T. PU.1 induces myeloid lineage commitment in multipotent hematopoietic progenitors. Genes Dev. 12, 2403–2412 (1998).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  27. Neph, S. et al. An expansive human regulatory lexicon encoded in transcription factor footprints. Nature 489, 83–90 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  28. Gjoneska, E. et al. Conserved epigenomic signals in mice and humans reveal immune basis of Alzheimer's disease. Nature 518, 365–369 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  29. Farh, K.K. et al. Genetic and epigenetic fine mapping of causal autoimmune disease variants. Nature 518, 337–343 (2015).

    Article  CAS  PubMed  Google Scholar 

  30. Maurano, M.T. et al. Systematic localization of common disease-associated variation in regulatory DNA. Science 337, 1190–1195 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

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

    Article  PubMed  Google Scholar 

  32. Bonnet, D. & Dick, J.E. Human acute myeloid leukemia is organized as a hierarchy that originates from a primitive hematopoietic cell. Nat. Med. 3, 730–737 (1997).

    Article  CAS  PubMed  Google Scholar 

  33. Goardon, N. et al. Coexistence of LMPP-like and GMP-like leukemia stem cells in acute myeloid leukemia. Cancer Cell 19, 138–152 (2011).

    Article  CAS  PubMed  Google Scholar 

  34. Bennett, J.M. et al. Proposals for the classification of the acute leukaemias. French-American-British (FAB) co-operative group. Br. J. Haematol. 33, 451–458 (1976).

    Article  CAS  PubMed  Google Scholar 

  35. van't Veer, M.B. The diagnosis of acute leukemia with undifferentiated or minimally differentiated blasts. Ann. Hematol. 64, 161–165 (1992).

    Article  CAS  PubMed  Google Scholar 

  36. Rangatia, J. et al. Elevated c-Jun expression in acute myeloid leukemias inhibits C/EBPα DNA binding via leucine zipper domain interaction. Oncogene 22, 4760–4764 (2003).

    Article  CAS  PubMed  Google Scholar 

  37. Volk, A. et al. Co-inhibition of NF-κB and JNK is synergistic in TNF-expressing human AML. J. Exp. Med. 211, 1093–1108 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  38. Hartman, A.D. et al. Constitutive c-Jun N-terminal kinase activity in acute myeloid leukemia derives from Flt3 and affects survival and proliferation. Exp. Hematol. 34, 1360–1376 (2006).

    Article  CAS  PubMed  Google Scholar 

  39. Magnusson, M., Brun, A.C.M., Lawrence, H.J. & Karlsson, S. Hoxa9/Hoxb3/Hoxb4 compound null mice display severe hematopoietic defects. Exp. Hematol. 35, 1421–1428 (2007).

    Article  CAS  PubMed  Google Scholar 

  40. Lawrence, H.J. et al. Mice bearing a targeted interruption of the homeobox gene HOXA9 have defects in myeloid, erythroid, and lymphoid hematopoiesis. Blood 89, 1922–1930 (1997).

    CAS  PubMed  Google Scholar 

  41. Thorsteinsdottir, U. et al. Overexpression of the myeloid leukemia–associated Hoxa9 gene in bone marrow cells induces stem cell expansion. Blood 99, 121–129 (2002).

    Article  CAS  PubMed  Google Scholar 

  42. González, A.J., Setty, M. & Leslie, C.S. Early enhancer establishment and regulatory locus complexity shape transcriptional programs in hematopoietic differentiation. Nat. Genet. 47, 1249–1259 (2015).

    Article  PubMed  PubMed Central  Google Scholar 

  43. Whitaker, J.W., Chen, Z. & Wang, W. Predicting the human epigenome from DNA motifs. Nat. Methods 12, 265–272 (2015).

    Article  CAS  PubMed  Google Scholar 

  44. Macedo, A. et al. Characterization of aberrant phenotypes in acute myeloblastic leukemia. Ann. Hematol. 70, 189–194 (1995).

    Article  CAS  PubMed  Google Scholar 

  45. Tiacci, E. et al. PAX5 expression in acute leukemias: higher B-lineage specificity than CD79a and selective association with t(8;21)-acute myelogenous leukemia. Cancer Res. 64, 7399–7404 (2004).

    Article  CAS  PubMed  Google Scholar 

  46. Jan, M. et al. Prospective separation of normal and leukemic stem cells based on differential expression of TIM3, a human acute myeloid leukemia stem cell marker. Proc. Natl. Acad. Sci. USA 108, 5009–5014 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  47. Hansen, K.D., Irizarry, R.A. & Wu, Z. Removing technical variability in RNA-seq data using conditional quantile normalization. Biostatistics 13, 204–216 (2012).

    Article  PubMed  PubMed Central  Google Scholar 

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

    Article  Google Scholar 

  49. Cancer Genome Atlas Research Network. Genomic and epigenomic landscapes of adult de novo acute myeloid leukemia. N. Engl. J. Med. 368, 2059–2074 (2013).

  50. McKenna, A. et al. The Genome Analysis Toolkit: a MapReduce framework for analyzing next-generation DNA sequencing data. Genome Res. 20, 1297–1303 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  51. Koboldt, D.C. et al. VarScan: variant detection in massively parallel sequencing of individual and pooled samples. Bioinformatics 25, 2283–2285 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  52. Newman, A.M. et al. FACTERA: a practical method for the discovery of genomic rearrangements at breakpoint resolution. Bioinformatics 30, 3390–3393 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  53. Ye, K., Schulz, M.H., Long, Q., Apweiler, R. & Ning, Z. Pindel: a pattern growth approach to detect break points of large deletions and medium sized insertions from paired-end short reads. Bioinformatics 25, 2865–2871 (2009).

    CAS  PubMed  PubMed Central  Google Scholar 

  54. Vaquerizas, J.M., Kummerfeld, S.K., Teichmann, S.A. & Luscombe, N.M. A census of human transcription factors: function, expression and evolution. Nat. Rev. Genet. 10, 252–263 (2009).

    Article  CAS  PubMed  Google Scholar 

  55. Weirauch, M.T. et al. Determination and inference of eukaryotic transcription factor sequence specificity. Cell 158, 1431–1443 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

Download references

Acknowledgements

We thank C. Mazumdar and A. Raj for assistance with RNA-seq, A. Newman for expert assistance with CIBERSORT, and our laboratory members for discussion. We thank the Stanford Hematology Division Tissue Bank and the patients for donating their samples. M.R.C. acknowledges NIH training grant R25CA180993 and NIH F31 Predoctoral fellowship F31CA180659. J.D.B. acknowledges National Science Foundation Graduate Research Fellowships and NIH training grant T32HG000044 for support. M.P.S. acknowledges the NIH and NHGRI for funding through 5U54HG00455805. Research was also supported by the NIH (P50HG007735 to H.Y.C., W.J.G., and M.P.S.), UH2AR067676 (H.Y.C.), the Stanford Cancer Center (H.Y.C.), the Howard Hughes Medical Institute (H.Y.C. and J.K.P.), the Stinehart-Reed Foundation (R.M.), the Ludwig Institute (R.M.), and the NIH (R01CA18805 to R.M.). R.M. is a New York Stem Cell Foundation Robertson Investigator.

Author information

Authors and Affiliations

Authors

Contributions

M.R.C., J.D.B., R.M., and H.Y.C. conceived the project. M.R.C. performed all cell sorting, RNA-seq, and CIBERSORT analysis, AML cell culture experiments, and mouse experiments. J.D.B. performed all ATAC-seq data analysis and regulatory network analysis, and oversaw all ATAC-seq library generation and protocol optimization performed by B.W. M.R.C. and J.L.K. performed DNA genotyping for patients with AML. J.D.B., P.G.G., and A.K. performed GWAS correlation analyses. W.J.G., M.P.S., and J.K.P. assisted with sequencing and study design. S.M.C. collected patient follow-up data and performed all survival analyses. M.R.C., J.D.B., R.M., and H.Y.C. wrote the manuscript with input from all authors.

Corresponding authors

Correspondence to Ravindra Majeti or Howard Y Chang.

Ethics declarations

Competing interests

Stanford University has filed a provisional patent application on the methods described, and J.D.B., H.Y.C., and W.J.G. are named as inventors. H.Y.C. and W.J.G. are founders of Epinomics.

Integrated supplementary information

Supplementary Figure 1 Validation of Fast-ATAC protocol and comparison to the original ATAC-seq protocol.

(a) ATAC-seq insert size distribution for three biological replicates of HSCs. (b,c) Enrichment of signal at annotated transcription start sites (TSSs) from Fast-ATAC data compared to DNase-seq (b) and previously published ATAC-seq (c) data using the original ATAC-seq protocol6. (d,e) The fraction of reads mapping to mitochondrial DNA (d) and the number of unique fragments in peaks per cell used in the assay (e), comparing the original ATAC-seq protocol and the Fast-ATAC protocol. Error bars in e, s.e.m. (f) Accessible chromatin landscape surrounding a constitutively accessible region of the genome. Profiles represent the union of all technical and biological replicates for each cell type. Genomic coordinates: chr. 19: 36,102,236–36,277,236. The y-axis scales range from 0–10 in normalized arbitrary units. (g,h) Reproducibility, measured by Pearson correlation of normalized counts per peak (Online Methods), for each cell type (g) and comparing technical and biological replicates (h). Error bars, 1 s.d.

Supplementary Figure 2 Sorting strategy for HSPCs.

Representative examples of sorting strategies and post-sort analysis for the seven CD34+ HSPC populations isolated in this study. Post-sort analyses represent a minimum of 100 events. Events displayed represent single-cell events for lineage-negative, live cells.

Supplementary Figure 3 Sorting strategy for differentiated lymphoid cells.

Representative examples of sorting strategies and post-sort analysis for the differentiated lymphoid cells isolated in this study. Post-sort analyses represent a minimum of 1,000 events. Events displayed represent single-cell events for lineage-negative, live cells.

Supplementary Figure 4 Sorting strategy for differentiated monocytes and erythroblasts.

Representative examples of sorting strategies and post-sort analysis for the differentiated erythroid (left) and myeloid (right) cells isolated in this study. Post-sort analyses represent a minimum of 1,000 events. Events displayed represent single-cell events from lineage-negative, live cells.

Supplementary Figure 5 Data processing pipelines.

(a,b) Data processing pipeline for ATAC-seq (a) and RNA-seq (b) data (see the Online Methods for a full description).

Supplementary Figure 6 Validation of RNA-seq and ATAC-seq using GO term enrichment analysis and motif enrichment analysis.

(a,b) GO term analyses from unique gene expression (a) and accessible peaks (b) from normal hematopoietic cells. (c) Enrichment of developmentally relevant motifs in accessible peaks using the HOMER21 motif enrichment tool.

Supplementary Figure 7 PCA and t-SNE analysis of RNA-seq and ATAC-seq data from primary human hematopoietic cells.

(a,b) PCA of RNA-seq (a) and ATAC-seq (b) data using all genes and peaks, respectively. (c,d) t-SNE analysis of RNA-seq (c) and ATAC-seq (d) data using all genes and peaks, respectively.

Supplementary Figure 8 Enhancer cytometry with CIBERSORT provides robust classification of most human hematopoietic cell types.

(ac) Leave-one-out cross-validation of CIBERSORT, comparing the predicted fraction to the synthesized (ground truth) fraction in MPPs (a), CLPs (b), and monocytes (c). For these in silico validations, one replicate of each normal cell type was excluded from use in generating the signature matrix using CIBERSORT. Random synthetic mixtures of these ‘left-out’ samples were then used to assess the ability of CIBERSORT to correctly predict cellular composition. (d) Overall classification performance of deconvolving synthesized mixtures for each hematopoietic cell type shown in Figure 3d. r squared value is derived from correlation of the CIBERSORT-predicted fraction with the synthesized ground truth fraction of 100 random permutations. (e,f) Correlation of deviation between MPPs and HSCs (e) and MPPs and CMPs (f) derived from Figure 3d. Higher correlation of observed deviations implies that the two cell types are more frequently misclassified for one another using CIBERSORT. (g) Heat map representation of all pairwise correlations for deviation as shown in e and f. Heat map color represents the r2 value from the correlation of deviation between two cell types (Online Methods). (h) Enhancer cytometry performed without manual removal of peaks corresponding to chromosome X, TSS peaks, and cancer-specific peaks shows a lower correlation with ground truth data (r2 = 0.91 as compared to r2 = 0.95 with manual curation of peaks). (i) Enhancer cytometry of DNase I hypersensitivity data using a signature matrix derived from our ATAC-seq data.

Supplementary Figure 9 Trans regulators of hematopoiesis.

(a) Summary of motif usage across hematopoiesis. The values shown represent normalized relative deviation scores. Relative deviation scores for representative transcription factor motifs are shown in Figure 4b,c. Scores are normalized by the maximum and minimum (range) across the 13 cell types assayed, such that the values shown range from 0 to 1. The range for the raw scores is shown above each column. Raw values can be found in Supplementary Table 3. (b) Unsupervised hierarchical clustering of hematopoiesis transcription factor motifs (n = 46; rows) with in vitro–derived motifs (CIS-BP) (n = 806; columns). Values shown represent motif similarity using Pearson correlation (Online Methods). Boxes highlight clustered CIS-BP motifs. (c,d) Examples of individual motif PWMs from clustered motifs as in b for the GATA (c) and MEIS1 (d) hematopoiesis transcription factor motifs. The top row represents a hematopoietic TF motif and subsequent rows show similar CIS-BP motifs from the cluster analysis. (e) Histogram of all correlation values in b. A correlation cutoff of 0.8 was used to assign potential regulators to motifs. Putative hematopoietic regulators are highlighted (n = 255). (f) Correlation of motif deviation scores to gene expression changes across hematopoietic cell types for two developmentally important transcription factors, GATA1 and PAX5. Each point represents a cell type comparison (for example, HSC versus MPP). Error bars, error in the gene expression measurement, as estimated by DESeq2. (g,h) Summary list of putative transcription factor positive (g) and negative (h) regulators of hematopoiesis motifs. Motifs are listed on the left, and the names of the genes predicted to regulate that motif are listed on the right. Values represent correlation coefficients (Pearson) comparing changes in motif deviation scores to the expression of putative regulators, as shown in f and Figure 4e, only the most correlated gene is shown here; a full list of correlation coefficients and P values can be found in Supplementary Table 4.

Supplementary Figure 10 GWAS enrichments across hematopoiesis.

(a) Representative example of GWAS enrichment across tissues. Fold enrichment (see the Online Methods for description of the calculation) is calculated from increasingly stringent GWAS SNP associations within different tissues. Line color represents tissue as shown in b. (b) Hierarchical clustering of row-normalized, rank fold enrichment values (Online Methods), for all GWAS collected (n = 235) across diverse tissues. The colored dendrogram and color bar below denote GWAS enriched in blood cells. (c) Summary of GWAS deviations across hematopoiesis. Scores are normalized by the maximum and minimum (range) across the 13 cell types assayed, such that the values shown range from 0 to 1. The range for the raw scores is shown above each column. Raw values can be found in Supplementary Table 3.

Supplementary Figure 11 Characterization of leukemogenic evolution in the patient cohort.

(a) Representative sorting strategy for HSCs (green), LSCs (yellow), and blasts (orange) from a single patient with AML (SU353). HSCs are characterized immunophenotypically as CD34+CD38CD99TIM3. LSCs are characterized immunophenotypically as CD34+CD38CD99+TIM3+. Blasts are characterized as CD99+TIM3+CD45midSSChigh. Depending on the sample, blasts can be CD34+ or CD34. (b) Preleukemic HSC purification strategies for all patients with AML analyzed by ATAC-seq and RNA-seq. Healthy bone marrow after CD34 magnetic bead enrichment is included as a normal control. (c) Mutation frequencies observed in the full patient cohort (Fig. 5a; n = 39). This cohort is biased toward samples harboring internal tandem duplications in FLT3. This lesion leads to a higher white blood cell count and, therefore, more cells to work with, making these samples more attractive. (d) Variant allele frequencies in HSCs isolated from patients with AML (n = 39) as determined by targeted amplicon sequencing to depth >500 reads. *P < 0.05, **P < 0.01, χ2 test. (e,f) Principal-component analysis of RNA-seq data (e) and ATAC-seq data (f) showing the spread of AML cell types across the myelopoietic continuum. The position of each normal hematopoietic cell type is shown using cartoon cells. These normal hematopoietic cell type positions represent the average for all biological and technical replicates.

Supplementary Figure 12 Validation of enhancer cytometry in AML cell lines and primary cells by single-cell ATAC-seq.

(a) Enhancer cytometry of ATAC-seq data as shown in Figure 6a but ordered by patient instead of by AML cell type. (b) Projection of downsampled bulk hematopoiesis data onto a 2D PCA. Bulk ATAC-seq data were downsampled to give 1,000 fragments, approximately the lower bound of fragments obtained by scATAC-seq. Each dot represents an individual permutation of downsampled bulk ATAC-seq data from HSCs (green), GMPs (yellow), and monocytes (orange). Principal components were derived from bulk ATAC-seq data for normal hematopoietic cells. (c) Projection of downsampled bulk hematopoiesis data onto a one-dimensional myelopoietic differentiation trajectory. (d,e) scATAC-seq data derived from FACS-purified LMPPs (n = 94) (d) and monocytes (n = 88) (e) projected onto a 2D PCA plot. Principal components were derived from bulk ATAC-seq data for normal hematopoietic cells. (f) Enhancer cytometry of bulk ATAC-seq data derived from various blood cell lines demonstrates mixed regulatory contribution from various normal hematopoietic cell types. (g) Projection of scATAC-seq data derived from healthy monocytes, LMPPs, and leukemic cell types from patients SU070 and SU353 onto a one-dimensional myeloid differentiation trajectory. (h) Projection of scATAC-seq data derived from HL60 cells (n = 90) onto a one-dimensional myeloid differentiation trajectory. (i,j) Projection of scATAC-seq data derived from single cells of the common cell lines K562 (n = 227 total) and TF1 (n = 56) (i) and GM12878 (n = 75) (j) onto a one-dimensional erythroid (i) and lymphoid (k) differentiation trajectory. AU, arbitrary units.

Supplementary Figure 13 Analysis of the synthetic normal analog comparisons used for AML cell types.

(a) Pearson correlations derived from comparison of all peaks for AML cell types with the closest normal analog (color) and the enhancer cytometry–derived synthetic normal (gray). (b) Total significant peaks observed after comparison of AML cell types to synthetic normal analogs. Significant peaks were defined as those with a log2 (fold change) value greater than 3 or less than –3 for comparison of AML with the closest normal or synthetic normal analog. (c) Comparison of AML cell types to synthetic normal analogs as shown in Figure 6i except reordered by patient instead of AML cell type. For each sample, the closest normal is shown with the color of the bar. The percentage of the total significant peaks that were removed by comparison to synthetic normal analogs is plotted for each sample. (d) Global epigenetic variance of AML cell types after subtraction of signal corresponding to synthetic normal analogs. Values are normalized to the maximum variance observed in Figure 5d to allow for comparison across plots.

Supplementary Figure 14 Validation of regulatory network analysis in AML cell types

(a) Mean Euclidean distance of centroids, derived by k-means clustering, of differentially accessible peaks in AML samples compared to their synthetic normal, for increasing K from 1 to 20. Mean distances across centroids stabilize after a K of 7.

(b) Expression of JUN in various normal hematopoietic cell types, pHSCs, and blasts. *P <0.05, two-tailed t-test

(c-e) The effect of JNK/ERK inhibition by (a) JNK-IN-8, (b) SP600125, and (c) SCH772984 was determined by IC50 of sorted primary AML blast cells in comparison to CD34+ HSPCs derived from umbilical cord blood. Viability was determined by flow cytometric assessment of Annexin V and DAPI.

(f) Strategy for in vitro differentiation of HSPCs down the myeloid and erythroid lineages. HSPCs are grown in defined culture media for 6 days and then analyzed for cell surface markers of stemness or differentiation. Immature cells at day 6 express CD34 and have not yet upregulated CD33.

(g) Quantitative reverse-transcriptase PCR validation of HOXA9 knockdown via shRNA. Knockdown performed in THP1 cells for 72 hours and validated with two separate primer sets.

(h,i) Fold change in the percent of (h) CD15+ granulocytes or (i) CD71+GPA+ erythroblasts between cord blood-derived CD34+ HSPCs infected with lentiviruses with non-targeting or HOXA9 shRNAs after 6 days of differentiation down the (h) myeloid or (i) erythroid lineage. ***P <0.001, ****P <0.0001 by two-tailed t-test

(j) Fold change in the percent of CD34+ HSPCs after 6 days of culture in stemness retention media (see methods) between cord blood-derived CD34+ HSPCs infected with lentiviruses with non-targeting or HOXA9 shRNAs.

(k) Burden (% of cells with mutation) of mutations in DNMT3A, TET2, IDH1/2, or other genes when detected in pre-leukemic HSC. *P < 0.05, **P < 0.01 by two-tailed t-test

Supplementary information

Supplementary Text and Figures

Supplementary Figures 1–14 and Supplementary Note. (PDF 5363 kb)

Supplementary Table 1

Normal donor information, cell sorting strategies, and sort purities. (XLSX 22 kb)

Supplementary Table 2

Enhancer cytometry signature matrix used for cell type deconvolution via CIBERSORT. (XLSX 222 kb)

Supplementary Table 3

Transcription factor and GWAS deviations across hematopoiesis. (XLSX 33 kb)

Supplementary Table 4

Transcription factor motif–gene association table and correlation values of RNA-seq. (XLSX 301 kb)

Supplementary Table 5

Patient clinical follow-up and genotyping data. (XLSX 64 kb)

Supplementary Table 6

Enhancer cytometry–derived fractional contributions of normal cell regulomes to AML samples. (XLSX 20 kb)

Supplementary Table 7

shRNA sequences used in this study. (XLSX 13 kb)

Supplementary Data: Compiled archive of motif images for hematopoiesis motifs obtained from JASPAR.

The motif images for all transcription factors used in this study are provided for reference. (ZIP 2563 kb)

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Corces, M., Buenrostro, J., Wu, B. et al. Lineage-specific and single-cell chromatin accessibility charts human hematopoiesis and leukemia evolution. Nat Genet 48, 1193–1203 (2016). https://doi.org/10.1038/ng.3646

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/ng.3646

This article is cited by

Search

Quick links

Nature Briefing: Translational Research

Sign up for the Nature Briefing: Translational Research newsletter — top stories in biotechnology, drug discovery and pharma.

Get what matters in translational research, free to your inbox weekly. Sign up for Nature Briefing: Translational Research