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:

Cryptic activation of an Irf8 enhancer governs cDC1 fate specification

Abstract

Induction of the transcription factor Irf8 in the common dendritic cell progenitor (CDP) is required for classical type 1 dendritic cell (cDC1) fate specification, but the mechanisms controlling this induction are unclear. In the present study Irf8 enhancers were identified via chromatin profiling of dendritic cells and CRISPR/Cas9 genome editing was used to assess their roles in Irf8 regulation. An enhancer 32 kilobases (kb) downstream of the Irf8 transcriptional start site (+32-kb Irf8) that was active in mature cDC1s was required for the development of this lineage, but not for its specification. Instead, a +41-kb Irf8 enhancer, previously thought to be active only in plasmacytoid dendritic cells, was found to also be transiently accessible in cDC1 progenitors, and deleting this enhancer prevented the induction of Irf8 in CDPs and abolished cDC1 specification. Thus, cryptic activation of the +41-kb Irf8 enhancer in dendritic cell progenitors is responsible for cDC1 fate specification.

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: The +32-kb Irf8 enhancer is required for cDC1 development.
Fig. 2: The +32-kb Irf8 enhancer is required for compensatory cDC1 development but not for cDC1 fate specification.
Fig. 3: The 50-kb Irf8 enhancer is not required for DC development.
Fig. 4: The 50-kb Irf8 enhancer controls Irf8 expression in monocytes and macrophages.
Fig. 5: ATAC-seq identifies the +41-kb Irf8 enhancer as transiently active in cDC1 progenitors.
Fig. 6: The +41-kb Irf8 enhancer is required for cDC1 fate specification and development.
Fig. 7: E-proteins are involved in the development of cDC1s and pDCs.

Similar content being viewed by others

Data availability.

The sequencing and microarray data generated during the course of this study have been deposited and are available on the Gene Expression Omnibus database. The ChIP-seq data of DC progenitors used in Fig. 3 can be accessed with the following accession number: GSE132239. The ATAC-seq data of DC progenitors used in Figs. 5 and 6 can be accessed with the following accession number: GSE132240. The microarrays utilized in Supplementary Fig. 4 can be accessed with the following accession numbers: GSE123747, GSE132767 and GSE132768.

All other primary data and materials that support the findings of this study are available from the corresponding author upon request.

References

  1. Graf, T. & Enver, T. Forcing cells to change lineages. Nature 462, 587–594 (2009).

    Article  CAS  Google Scholar 

  2. Orkin, S. H. & Zon, L. I. Hematopoiesis: an evolving paradigm for stem cell biology. Cell 132, 631–644 (2008).

    Article  CAS  Google Scholar 

  3. Miyamoto, T. et al. Myeloid or lymphoid promiscuity as a critical step in hematopoietic lineage commitment. Dev. Cell 3, 137–147 (2002).

    Article  CAS  Google Scholar 

  4. Hoppe, P. S. et al. Early myeloid lineage choice is not initiated by random PU.1 to GATA1 protein ratios. Nature 535, 299–302 (2016).

    Article  CAS  Google Scholar 

  5. Long, H. K., Prescott, S. L. & Wysocka, J. Ever-changing landscapes: transcriptional enhancers in development and evolution. Cell 167, 1170–1187 (2016).

    Article  CAS  Google Scholar 

  6. Sagai, T. et al. Elimination of a long-range cis-regulatory module causes complete loss of limb-specific Shh expression and truncation of the mouse limb. Development 132, 797–803 (2005).

    Article  CAS  Google Scholar 

  7. Shim, S. et al. Cis-regulatory control of corticospinal system development and evolution. Nature 486, 74–79 (2012).

    Article  CAS  Google Scholar 

  8. Frankel, N. et al. Phenotypic robustness conferred by apparently redundant transcriptional enhancers. Nature 466, 490–493 (2010).

    Article  CAS  Google Scholar 

  9. Osterwalder, M. et al. Enhancer redundancy provides phenotypic robustness in mammalian development. Nature 554, 239–243 (2018).

    Article  CAS  Google Scholar 

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

    Article  CAS  Google Scholar 

  11. Steinman, R. M. & Cohn, Z. A. Identification of a novel cell type in peripheral lymphoid organs of mice. I. Morphology, quantitation, tissue distribution. J. Exp. Med. 137, 1142–1162 (1973).

    Article  CAS  Google Scholar 

  12. Cella, M. et al. Plasmacytoid monocytes migrate to inflamed lymph nodes and produce large amounts of type I interferon. Nat. Med. 5, 919–923 (1999).

    Article  CAS  Google Scholar 

  13. Durai, V. & Murphy, K. M. Functions of murine dendritic cells. Immunity 45, 719–736 (2016).

    Article  CAS  Google Scholar 

  14. Hildner, K. et al. Batf3 deficiency reveals a critical role for CD8alpha+ dendritic cells in cytotoxic T cell immunity. Science 322, 1097–1100 (2008).

    Article  CAS  Google Scholar 

  15. Gubin, M. M. et al. Checkpoint blockade cancer immunotherapy targets tumour-specific mutant antigens. Nature 515, 577–581 (2014).

    Article  CAS  Google Scholar 

  16. Salmon, H. et al. Expansion and activation of CD103+ dendritic cell progenitors at the tumor site enhances tumor responses to therapeutic PD-L1 and BRAF inhibition. Immunity 44, 924–938 (2016).

    Article  CAS  Google Scholar 

  17. Saxena, M. & Bhardwaj, N. Re-emergence of dendritic cell vaccines for cancer treatment. Trends Cancer 4, 119–137 (2018).

    Article  CAS  Google Scholar 

  18. Fogg, D. K. et al. A clonogenic bone marrow progenitor specific for macrophages and dendritic cells. Science 311, 83–87 (2006).

    Article  CAS  Google Scholar 

  19. Naik, S. H. et al. Development of plasmacytoid and conventional dendritic cell subtypes from single precursor cells derived in vitro and in vivo. Nat. Immunol. 8, 1217–1226 (2007).

    Article  CAS  Google Scholar 

  20. Onai, N. et al. Identification of clonogenic common Flt3+ M-CSFR+ plasmacytoid and conventional dendritic cell progenitors in mouse bone marrow. Nat. Immunol. 8, 1207–1216 (2007).

    Article  CAS  Google Scholar 

  21. Grajales-Reyes, G. E. et al. Batf3 maintains autoactivation of Irf8 for commitment of a CD8alpha+ conventional DC clonogenic progenitor. Nat. Immunol. 16, 708–717 (2015).

    Article  CAS  Google Scholar 

  22. Schlitzer, A. et al. Identification of cDC1- and cDC2-committed DC progenitors reveals early lineage priming at the common DC progenitor stage in the bone marrow. Nat. Immunol. 16, 718–728 (2015).

    Article  CAS  Google Scholar 

  23. Schiavoni, G. et al. ICSBP is essential for the development of mouse type I interferon-producing cells and for the generation and activation of CD8alpha+ dendritic cells. J. Exp. Med. 196, 1415–1425 (2002).

    Article  CAS  Google Scholar 

  24. Kashiwada, M. et al. NFIL3/E4BP4 is a key transcription factor for CD8{alpha}+ dendritic cell development. Blood 117, 6193–6197 (2011).

    Article  CAS  Google Scholar 

  25. Hacker, C. et al. Transcriptional profiling identifies Id2 function in dendritic cell development. Nat. Immunol. 4, 380–386 (2003).

    Article  CAS  Google Scholar 

  26. Tussiwand, R. et al. Compensatory dendritic cell development mediated by BATF–IRF interactions. Nature 490, 502–507 (2012).

    Article  CAS  Google Scholar 

  27. Seillet, C. et al. CD8alpha+ DCs can be induced in the absence of transcription factors Id2, Nfil3, and Batf3. Blood 121, 1574–1583 (2013).

    Article  CAS  Google Scholar 

  28. Cisse, B. et al. Transcription factor E2-2 is an essential and specific regulator of plasmacytoid dendritic cell development. Cell 135, 37–48 (2008).

    Article  CAS  Google Scholar 

  29. Schonheit, J. et al. PU.1 level-directed chromatin structure remodeling at the Irf8 gene drives dendritic cell commitment. Cell Rep. 3, 1617–1628 (2013).

    Article  Google Scholar 

  30. Platt, R. J. et al. CRISPR-Cas9 knockin mice for genome editing and cancer modeling. Cell 159, 440–455 (2014).

    Article  CAS  Google Scholar 

  31. Diamond, M. S. et al. Type I interferon is selectively required by dendritic cells for immune rejection of tumors. J. Exp. Med. 208, 1989–2003 (2011).

    Article  CAS  Google Scholar 

  32. Karki, R. et al. IRF8 regulates transcription of Naips for NLRC4 inflammasome activation. Cell 173, 920–933 (2018).

    Article  CAS  Google Scholar 

  33. Buenrostro, J. D. et al. 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  Google Scholar 

  34. 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  Google Scholar 

  35. Heath, V. et al. C/EBPalpha deficiency results in hyperproliferation of hematopoietic progenitor cells and disrupts macrophage development in vitro and in vivo. Blood 104, 1639–1647 (2004).

    Article  CAS  Google Scholar 

  36. Labzin, L. I. et al. ATF3 is a key regulator of macrophage IFN responses. J. Immunol. 195, 4446–4455 (2015).

    Article  CAS  Google Scholar 

  37. Sawai, C. M. et al. Transcription factor Runx2 controls the development and migration of plasmacytoid dendritic cells. J. Exp. Med. 210, 2151–2159 (2013).

    Article  CAS  Google Scholar 

  38. Satpathy, A. T. et al. Runx1 and Cbfbeta regulate the development of Flt3+ dendritic cell progenitors and restrict myeloproliferative disorder. Blood 123, 2968–2977 (2014).

    Article  CAS  Google Scholar 

  39. Theisen, D. J. et al. WDFY4 is required for cross-presentation in response to viral and tumor antigens. Science 362, 694–699 (2018).

    Article  CAS  Google Scholar 

  40. Sichien, D. et al. IRF8 transcription factor controls survival and function of terminally differentiated conventional and plasmacytoid dendritic cells, respectively. Immunity 45, 626–640 (2016).

    Article  CAS  Google Scholar 

  41. Kwon, K. et al. Instructive role of the transcription factor E2A in early B lymphopoiesis and germinal center B cell development. Immunity 28, 751–762 (2008).

    Article  CAS  Google Scholar 

  42. Wohner, M. et al. Molecular functions of the transcription factors E2A and E2-2 in controlling germinal center B cell and plasma cell development. J. Exp. Med. 213, 1201–1221 (2016).

    Article  Google Scholar 

  43. Iwata, A. et al. Quality of TCR signaling determined by differential affinities of enhancers for the composite BATF–IRF4 transcription factor complex. Nat. Immunol. 18, 563–572 (2017).

    Article  CAS  Google Scholar 

  44. Zhang, P. et al. PU.1 inhibits GATA-1 function and erythroid differentiation by blocking GATA-1 DNA binding. Blood 96, 2641–2648 (2000).

    CAS  PubMed  Google Scholar 

  45. Nerlov, C. et al. GATA-1 interacts with the myeloid PU.1 transcription factor and represses PU.1-dependent transcription. Blood 95, 2543–2551 (2000).

    CAS  PubMed  Google Scholar 

  46. Tsai, S. F., Strauss, E. & Orkin, S. H. Functional analysis and in vivo footprinting implicate the erythroid transcription factor GATA-1 as a positive regulator of its own promoter. Genes Devel. 5, 919–931 (1991).

    Article  CAS  Google Scholar 

  47. Chen, H. et al. PU.1 (Spi-1) autoregulates its expression in myeloid cells. Oncogene 11, 1549–1560 (1995).

    CAS  PubMed  Google Scholar 

  48. Heng, T. S. & Painter, M. W., Immunological Genome Project Consortium. The Immunological Genome Project: networks of gene expression in immune cells. Nat. Immunol. 9, 1091–1094 (2008).

    Article  CAS  Google Scholar 

  49. Durai, V. et al. Altered compensatory cytokine signaling underlies the discrepancy between Flt3-/- and Flt3l-/- mice. J. Exp. Med. 215, 1417–1435 (2018).

    Article  CAS  Google Scholar 

  50. Ranganath, S. et al. GATA-3-dependent enhancer activity in IL-4 gene regulation. J. Immunol. 161, 3822–3826 (1998).

    CAS  PubMed  Google Scholar 

  51. Sedy, J. R. et al. B and T lymphocyte attenuator regulates T cell activation through interaction with herpesvirus entry mediator. Nat. Immunol. 6, 90–98 (2005).

    Article  CAS  Google Scholar 

  52. Zhu, H. et al. Unexpected characteristics of the IFN-gamma reporters in nontransformed T cells. J. Immunol. 167, 855–865 (2001).

    Article  CAS  Google Scholar 

  53. McKinney, J. et al. Tightly regulated gene expression system in Salmonella enterica serovar Typhimurium. J. Bacteriol. 184, 6056–6059 (2002).

    Article  CAS  Google Scholar 

  54. Kohwi-Shigematsu, T. et al. SATB1-mediated functional packaging of chromatin into loops. Methods 58, 243–254 (2012).

    Article  CAS  Google Scholar 

  55. Corces, M. R. et al. The chromatin accessibility landscape of primary human cancers. Science 362, 420 (2018).

    Article  CAS  Google Scholar 

  56. Schep, A. N. et al. chromVAR: inferring transcription-factor-associated accessibility from single-cell epigenomic data. Nat. Methods 14, 975–978 (2017).

    Article  CAS  Google Scholar 

  57. 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  Google Scholar 

Download references

Acknowledgements

This work was supported by the Howard Hughes Medical Institute (K.M. and H.C.), the US National Institutes of Health (NIH: F30 DK108498 to V.D.; K08 CA230188 to A.S.; P50 HG007735 to H.C.; R01 AI106352 to B.K.; R01 DK097317 to R.N.), the National Science Foundation (DGE-1745038 to P.B.), the Parker Institute for Cancer Immunotherapy (A.S. and H.C.) and Boehringer Ingelheim (M.W., H.T. and M.B.). A.S. was supported by a Career Award for Medical Scientists from the Burroughs Wellcome Fund. This work benefitted from data assembled by the ImmGen consortium48. We thank the Genome Technology Access Center in the Department of Genetics at Washington University in St Louis School of Medicine for help with genomic analysis. The Center is partially supported by the National Cancer Institute’s Cancer Center Support grant no. P30 CA91842 to the Siteman Cancer Center and by the Institute of Clinical and Translational Sciences/Clinical and Translational Science Award grant no. UL1TR000448 from the National Center for Research Resources (NCRR), a component of the NIH, and NIH Roadmap for Medical Research. This publication is solely the responsibility of the authors and does not necessarily represent the official view of the NCRR or the NIH.

Author information

Authors and Affiliations

Authors

Contributions

V.D., T.L.M. and K.M.M. designed the study. V.D. generated the enhancer knockout mice with advice and reagents from S.J.P. V.D. and P.B. performed experiments related to analysis of immune populations, cell sorting and culture, with advice from J.T.D., R.W., T.-T.L., X.H., C.G.B. and G.E.G.-R. V.D., J.M.G., A.T.S. and H.Y.C. performed ATAC-seq of DC progenitors. J.M.G., A.T.S. and H.Y.C. performed computational analysis of ATAC-seq data. D.H.K. and R.D.N. performed Salmonella infections. A.I. performed ChIP-seq and computational analysis on DC progenitors. M.W., H.T. and M.B. provided ChIP-seq and ATAC-seq data from B cells and advice. B.L.K. provided Tcf3–/– mice and advice. V.D. and K.M.M. wrote the manuscript with advice from all the authors.

Corresponding author

Correspondence to Kenneth M. Murphy.

Ethics declarations

Competing interests

The authors declare no competing interests.

Additional information

Peer review information: Z. Fehervari was the primary editor on this article and managed its editorial process and peer review in collaboration with the rest of the editorial team.

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

Integrated supplementary information

Supplementary Figure 1 AICEs in the +32-kb Irf8 enhancer drive cDC1 development.

a, Shown are the sequences, either wild type (3x AICE1 WT) or mutated (3x AICE1 Mut), of the first 3 AICEs from the +32 kb Irf8 enhancer used in retroviral reporter assays in (b). b, The regions from (a) were placed in a reverse strand retroviral reporter and GFP activity was measured in either cDC1s or cDC2s as previously described21. Gray histograms represent the activity of the empty reporter. Data are representative of three independent experiments with similar results (n = 3 biological replicates per group). c, The full 574 bp +32 kb Irf8 enhancer (+32kb WT), or the 5’ half containing 4 AICE sites (+32 kb 5’) or the 3’ half containing no AICEs (+32 kb 3’), were assayed for activity in cDC1s using retroviral reporters as in (b). Numbers indicate the mean fluorescence intensity (MFI) of GFP in cDC1s. Data are representative of three independent experiments with similar results (n = 3 biological replicates per group). d, CD117hi BM cells from R26Cas9/+ mice were infected with a retroviral vector expressing a control scramble sgRNA (scramble), an sgRNA targeting the Irf8 coding sequence (αIrf8), or an sgRNA targeting the AICE site 4 shown in (a) (αIrf8 +32kb), and then cultured in Flt3L for 8 days. Flow cytometry was then used to assess cDC1 development. Cells are pre-gated as Thy1.1+, to identify infected cells, and as B220MHC-II+CD11c+, to identify DCs. Numbers indicate the percent of cells in the indicated gates. Data are representative of three independent experiments (n = 3 biological replicates per group). e, Shown is the nucleotide sequence of the +32 kb Irf8 enhancer region. The extent of deletion in the two mouse strains is indicated by colored fonts. Pink nucleotides were deleted only in Irf8 +32 5’−/− mice, blue nucleotides were deleted in both Irf8 +32 5’−/− mice and Irf8 +32−/− mice, and red nucleotides were deleted only in Irf8 +32–/– mice. Red lines indicate the sequences of the sgRNAs used for CRISPR/Cas9 induced deletion in mice. Green boxes indicate the AICE motifs. f, PCR of the +32 kb locus in wildtype (WT), Irf8 +32 5’−/−, and Irf8 +32−/− mice demonstrating the deletions produced by CRISPR/Cas9 genome editing. Numbers indicate the size of the PCR product from each strain. Data are representative of two independent experiments with similar results (n = 2 biological replicates per population). g,h, Flow cytometry of live splenocytes from mice of the indicated genotypes was used to identify dendritic cell (DC) subsets. Numbers indicate the percent of cells in the indicated gates (g). Statistical analysis of the frequency and absolute number of splenic cDC1s in mice of the indicated genotypes (h). Small horizontal lines indicate the mean. Data are pooled from three independent experiments (n = 6 mice for WT, n = 5 mice for Irf8 +32 5’–/–, and n = 5 mice for Irf8 +32–/–). ***P<0.001; and ****P<0.0001, unpaired two-tailed Student’s t-test (h).

Supplementary Figure 2 Loss of the +32-kb Irf8 enhancer does not impact non-cDC1 immune lineages.

a,b, Flow cytometry of live splenocytes from mice of the indicated genotypes was used to examine myeloid and lymphoid cells. Numbers indicate the percent of cells in the indicated gates (a). Statistical analysis of the frequency of the indicated populations within spleens of mice of the indicated genotypes (b). Small horizontal lines indicate the mean. Data are pooled from four independent experiments (n = 5 mice for WT, n = 5 mice for Irf8 +32–/–, and n = 3 mice for Irf8–/–). c, Intracellular staining for IRF8 in the indicated populations in mice of the indicated genotypes. Data are representative of four independent experiments with similar results (n = 5 mice for WT, n = 5 mice for Irf8 +32–/–, and n = 3 mice for Irf8–/–). ns, not significant (P>0.05); **P<0.01; ***P<0.001; and ****P<0.0001, ordinary one-way ANOVA (b).

Supplementary Figure 3 The +32-kb Irf8 enhancer is required for all compensatory cDC1 development.

a,b, Flow cytometry of lung cells from mice of the indicated genotypes was used to identify DC subsets. Numbers indicate the percent of cells in the indicated gates (a). Statistical analysis of the frequency of cDC1s among CD45+ cells from the lungs of mice of the indicated genotypes (b). Small horizontal lines indicate the mean. Data are pooled from four independent experiments (n = 5 mice for WT, n = 5 mice for Irf8 +32–/–, n = 3 mice for Batf3–/–, and n = 3 mice for Irf8–/–). c,d, BM cells from mice of the indicated genotypes was transferred into irradiated WT mice and allowed to reconstitute for three weeks. Flow cytometry of live splenocytes from these chimeras was then used to identify DC subsets. Numbers indicate the percent of cells in the indicated gates (c). Statistical analysis of the frequency of splenic cDC1s in the indicated chimeras (d). Small horizontal lines indicate the mean. Data are pooled from two independent experiments (n = 3 for WT into WT chimeras, n = 5 for Irf8 +32–/– into WT chimeras, n = 5 for Batf3–/– into WT chimeras). e, Statistical analysis of the frequency of pre-cDC1s in mice of the indicated genotypes. Small horizontal lines indicate the mean. Data are pooled from three independent experiments (n = 3 mice for WT and n = 3 mice for Irf8 +32–/–). ns, not significant (P>0.05); **P<0.01, ordinary one-way ANOVA (b,d) or unpaired two-tailed Student’s t-test (e).

Supplementary Figure 4 CD226 uniquely identifies pre-cDC1s in the bone marrow.

a, Cd226 gene expression levels in CDPs and pre-cDC1s quantified by gene expression microarrays. Small horizontal lines indicate the mean. Data are pooled from three independent experiments (n = 3 biological replicates per population). b, Flow cytometry of Lin BM cells from WT mice was used to examine marker expression in pre-cDC1s. Data are representative of four independent experiments with similar results (n = 4 mice). c,d, CDPs (LinCD117intCD135+CD115+) and Zbtb46-GFP+ pre-cDC1s (LinCD117intCD135+Zbtb46-GFP+) (c), or CDPs and CD226+ pre-cDC1s (LinCD117intCD135+CD226+) (d) were sorted from mice and cultured for 5 days in Flt3L. Cells were then analyzed by flow cytometry to identify dendritic cell subsets. Data are representative of three independent experiments with similar results (n = 3 biological replicates per population). e, Microarray analysis of cDC2s derived from in vitro culture of pre-cDC1s from Irf8 +32–/– mice compared to cDC1s derived from in vitro culture of WT pre-cDC1s (left) or compared to cDC2s derived from in vitro culture of WT pre-cDC2s (right), presented as M-plots. Green lines indicate the threshold for a two-fold difference in expression between samples. Data was pooled from two independent experiments (n = 2 biological replicates per population). f, Microarray analysis of WT CD24+CD172a+ cDC2s compared to WT CD24CD172+ cDC2s (left) or compared to WT CD24+CD172a cDC1s (right), presented as M-plots. Green lines indicate the threshold for a two-fold difference in expression between samples. Data was pooled from two independent experiments (n = 2 biological replicates per population). g, Irf4 expression levels in the indicated populations quantified by gene expression microarrays. Small horizontal lines indicate the mean. Data was pooled from two independent experiments (n = 2 biological replicates per population). h, Microarray analysis of cDC2s derived from in vitro culture of Batf3–/– pre-cDC1s compared to cDC1s derived from in vitro culture of WT pre-cDC1s (left) or compared to cDC2s derived from in vitro culture of WT pre-cDC2s (right), presented as M-plots. Green lines indicate the threshold for a two-fold difference in expression between samples. Data was pooled from two independent experiments (n = 2 biological replicates per population).

Supplementary Figure 5 Loss of the −50-kb Irf8 enhancer does not impact non-monocyte/macrophage immune lineages.

a, PCR of the -50 kb locus in WT and Irf8 -50–/– mice demonstrating the deletion produced by CRISPR/Cas9 genome editing. Numbers indicate the size of the PCR product from each strain. Data are representative of two independent experiments with similar results (n = 2 biological replicates per population). b, Shown is the nucleotide sequence of the -50 kb Irf8 enhancer region. The extent of deletion in Irf8 -50—/— mice is indicated by colored font. Red nucleotides were deleted in Irf8 -50—/— mice. Red lines indicate the sequences of the sgRNAs used for CRISPR/Cas9 induced deletion in mice. Purple boxes indicate the PU.1 motifs. c, Statistical analysis of the frequency of the indicated populations gated as in Supplementary Fig. 2a in spleens of mice of the indicated genotypes. Small horizontal lines indicate the mean. Data are pooled from four independent experiments. (n = 5 mice for WT, n = 5 mice for Irf8 -50–/–, and n = 4 mice for Irf8–/–). d, Intracellular staining for IRF8 in the indicated populations in mice of the indicated genotypes. Data are representative of four independent experiments with similar results (n = 5 mice for WT, n = 5 mice for Irf8 -50–/–, and n = 4 mice for Irf8–/–). ns, not significant (P>0.05); *P<0.05; **P<0.01; and ****P<0.0001, ordinary one-way ANOVA (c).

Supplementary Figure 6 Loss of the +41-kb Irf8 enhancer does not impact non-DC immune lineages.

a, PCR of the +41 kb locus in WT and Irf8 +41–/– mice demonstrating the deletion produced by CRISPR/Cas9 genome editing. Numbers indicate the size of the PCR product from each strain. Data are representative of two independent experiments with similar results (n = 2 biological replicates per population). b, Shown is the nucleotide sequence of the +41-kb Irf8 enhancer region. The extent of deletion in Irf8 +41–/– mice is indicated by colored font. Red nucleotides were deleted in Irf8 +41–/– mice. Red lines indicate the sequences of the sgRNAs used for CRISPR/Cas9 induced deletion in mice. Blue boxes indicate the E box motifs. c, Statistical analysis of the frequency of the indicated populations gated as in Supplementary Fig. 2a in spleens of mice of the indicated genotypes. Small horizontal lines indicate the mean. Data are pooled from five independent experiments (n = 7 mice for WT, n = 7 mice for Irf8 +41–/–, and n = 4 mice for Irf8–/–). d, Intracellular staining for IRF8 in the indicated populations in mice of the indicated genotypes. Data are representative of four independent experiments with similar results (n = 5 mice for WT, n = 5 mice for Irf8 +41–/–, and n = 3 mice for Irf8–/). ns, not significant (P>0.05); ***P<0.001; and ****P<0.0001, ordinary one-way ANOVA (c).

Supplementary Figure 7 Tcf3–/– BM progenitors have reduced cDC1 and pDC potential.

a, Flow cytometry of BM cells from mice of the indicated genotypes was used to determine intracellular E2A-GFP levels in the indicated populations. CMPs were gated as LinCD117hiSca-1CD16/32+CD34+, CLPs were gated as LinCD127+CD135+CD117loB220CD11c, and pre-B cells were gated as LinCD19+CD25+. Numbers indicate the MFI of E2A-GFP levels. Data are representative of two independent experiments with similar results (n = 3 mice for Tcf3fl/+ and n = 3 mice for Tcf3fl/+VavCretg). b,c, Flow cytometry of live splenocytes from mice of the indicated genotypes was used to identify DC subsets. Numbers indicate the percent of cells in the indicated gates (b). Statistical analysis of the frequency of splenic cDC1s and pDCs in mice of the indicated genotypes (c). Small horizontal lines indicate the mean. Data are pooled from three independent experiments. (n = 4 mice for WT and n = 3 mice for Tcf3–/–). d,e, WT CD45.1 mice were lethally irradiated and reconstituted with Tcf3+/+ CD45.2: Tcf3+/+ CD45.1 BM at a 1:1 ratio or Tcf3+/+ CD45.2: Tcf3–/– CD45.1 BM at a 1:1 ratio. Mice were analyzed after four weeks of reconstitution. Flow cytometry of live splenocytes from the indicated chimeras was then used to identify DC subsets gated as the following: cDC2s, CD11c+MHCII+CD172a+; cDC1s, CD11c+MHCII+CD24+CD172a; pDCs, B220+CD317+. Numbers indicate the percent of cells in the indicated gates (d). Summary data of the chimerism ratio of splenic cDC1s (left) or pDCs (right) presented as the ratio of CD45.1+ to CD45.2+ cells of each type normalized to the ratio of CD45.1+ to CD45.2+ cDC2s within the same mouse (e). Small horizontal lines indicate the mean. Data are pooled from two independent experiments (n = 6 mice for Tcf3+/+:Tcf3+/+ chimeras and n = 4 mice for Tcf3+/+:Tcf3–/– chimeras). ns, not significant (P>0.05); *P<0.01; and ****P<0.0001, unpaired two-tailed Student’s t-test (c,e).

Supplementary information

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Durai, V., Bagadia, P., Granja, J.M. et al. Cryptic activation of an Irf8 enhancer governs cDC1 fate specification. Nat Immunol 20, 1161–1173 (2019). https://doi.org/10.1038/s41590-019-0450-x

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41590-019-0450-x

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