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:

Nutrition impact on ILC3 maintenance and function centers on a cell-intrinsic CD71–iron axis

Abstract

Iron metabolism is pivotal for cell fitness in the mammalian host; however, its role in group 3 innate lymphoid cells (ILC3s) is unknown. Here we show that transferrin receptor CD71 (encoded by Tfrc)-mediated iron metabolism cell-intrinsically controls ILC3 proliferation and host protection against Citrobacterrodentium infection and metabolically affects mitochondrial respiration by switching of oxidative phosphorylation toward glycolysis. Iron deprivation or Tfrc ablation in ILC3s reduces the expression and/or activity of the aryl hydrocarbon receptor (Ahr), a key ILC3 regulator. Genetic ablation or activation of Ahr in ILC3s leads to CD71 upregulation or downregulation, respectively, suggesting Ahr-mediated suppression of CD71. Mechanistically, Ahr directly binds to the Tfrc promoter to inhibit transcription. Iron overload partially restores the defective ILC3 compartment in the small intestine of Ahr-deficient mice, consistent with the compensatory upregulation of CD71. These data collectively demonstrate an under-appreciated role of the Ahr-CD71–iron axis in the regulation of ILC3 maintenance and function.

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: Iron-depleted diet impairs gut ILC3 maintenance and function.
Fig. 2: Iron chelation represses ILC3 functions ex vivo and in vitro.
Fig. 3: Iron depletion impairs mitochondrial function and switches OXPHOS to glycolysis in MNK-3 cells.
Fig. 4: CD71 deficiency impairs ILC3 maintenance in a cell-intrinsic manner.
Fig. 5: CD71 ablation in ILC3s impairs gut immunity to C.rodentium.
Fig. 6: Ahr cell-intrinsically inhibits CD71 transcription in ILC3s.
Fig. 7: Ahr regulates CD71 transcription by directly binding to the Tfrc locus.

Similar content being viewed by others

Data availability

All data supporting the findings of this study are available within the article and its supplementary materials. The M.musculus reference genome (GRCm38/mm10 assembly) https://www.ncbi.nlm.nih.gov/assembly/GCF_000001635.20/ was used for RNA-seq and ChIP-seq read alignment. The Gene Expression Omnibus accession code for the sequencing data in this paper is GSE222621. Source data are provided with this paper.

References

  1. Lieu, P. T., Heiskala, M., Peterson, P. A. & Yang, Y. The roles of iron in health and disease. Mol. Asp. Med. 22, 1–87 (2001).

    CAS  Google Scholar 

  2. Hentze, M. W., Muckenthaler, M. U., Galy, B. & Camaschella, C. Two to tango: regulation of mammalian iron metabolism. Cell 142, 24–38 (2010).

    CAS  PubMed  Google Scholar 

  3. Donovan, A. et al. Positional cloning of zebrafish ferroportin1 identifies a conserved vertebrate iron exporter. Nature 403, 776–781 (2000).

    CAS  PubMed  Google Scholar 

  4. Abboud, S. & Haile, D. J. A novel mammalian iron-regulated protein involved in intracellular iron metabolism. J. Biol. Chem. 275, 19906–19912 (2000).

    CAS  PubMed  Google Scholar 

  5. McKie, A. T. et al. A novel duodenal iron-regulated transporter, ireg1, implicated in the basolateral transfer of iron to the circulation. Mol. Cell 5, 299–309 (2000).

    CAS  PubMed  Google Scholar 

  6. Dev, S. & Babitt, J. L. Overview of iron metabolism in health and disease. Hemodial. Int. 21, S6–S20 (2017).

    PubMed  PubMed Central  Google Scholar 

  7. Aydemir, T. B. & Cousins, R. J. The multiple faces of the metal transporter zip14 (slc39a14). J. Nutr. 148, 174–184 (2018).

    PubMed  PubMed Central  Google Scholar 

  8. Jenkitkasemwong, S. et al. Slc39a14 is required for the development of hepatocellular iron overload in murine models of hereditary hemochromatosis. Cell Metab. 22, 138–150 (2015).

    CAS  PubMed  PubMed Central  Google Scholar 

  9. Gammella, E., Buratti, P., Cairo, G. & Recalcati, S. The transferrin receptor: the cellular iron gate. Metallomics 9, 1367–1375 (2017).

    CAS  PubMed  Google Scholar 

  10. Lill, R. Function and biogenesis of iron-sulphur proteins. Nature 460, 831–838 (2009).

    CAS  PubMed  Google Scholar 

  11. Li, H. et al. Iron regulatory protein 2 modulates the switch from aerobic glycolysis to oxidative phosphorylation in mouse embryonic fibroblasts. Proc. Natl Acad. Sci. USA 116, 9871–9876 (2019).

    CAS  PubMed  PubMed Central  Google Scholar 

  12. Ye, H. et al. Glutaredoxin 5 deficiency causes sideroblastic anemia by specifically impairing heme biosynthesis and depleting cytosolic iron in human erythroblasts. J. Clin. Invest. 120, 1749–1761 (2010).

    CAS  PubMed  PubMed Central  Google Scholar 

  13. Cronin, S. J. F., Woolf, C. J., Weiss, G. & Penninger, J. M. The role of iron regulation in immunometabolism and immune-related disease. Front. Mol. Biosci. 6, 116 (2019).

    CAS  PubMed  PubMed Central  Google Scholar 

  14. Chen, A. C., Donovan, A., Ned-Sykes, R. & Andrews, N. C. Noncanonical role of transferrin receptor 1 is essential for intestinal homeostasis. Proc. Natl Acad. Sci. USA 112, 11714–11719 (2015).

    CAS  PubMed  PubMed Central  Google Scholar 

  15. Barrientos, T. et al. Metabolic catastrophe in mice lacking transferrin receptor in muscle. eBioMedicine 2, 1705–1717 (2015).

    PubMed  PubMed Central  Google Scholar 

  16. Artis, D. & Spits, H. The biology of innate lymphoid cells. Nature 517, 293–301 (2015).

    CAS  PubMed  Google Scholar 

  17. Diefenbach, A., Colonna, M. & Koyasu, S. Development, differentiation, and diversity of innate lymphoid cells. Immunity 41, 354–365 (2014).

    CAS  PubMed  PubMed Central  Google Scholar 

  18. Abt, M. C. et al. Innate immune defenses mediated by two ilc subsets are critical for protection against acute Clostridium difficile infection. Cell Host Microbe 18, 27–37 (2015).

    CAS  PubMed  PubMed Central  Google Scholar 

  19. Song, C. et al. Unique and redundant functions of nkp46+ ilc3s in models of intestinal inflammation. J. Exp. Med. 212, 1869–1882 (2015).

    CAS  PubMed  PubMed Central  Google Scholar 

  20. Xiong, L. et al. Group 3 innate lymphoid cell pyroptosis represents a host defence mechanism against salmonella infection. Nat. Microbiol 7, 1087–1099 (2022).

    CAS  PubMed  PubMed Central  Google Scholar 

  21. McKenzie, A. N. J., Spits, H. & Eberl, G. Innate lymphoid cells in inflammation and immunity. Immunity 41, 366–374 (2014).

    CAS  PubMed  Google Scholar 

  22. Silberger, D. J., Zindl, C. L. & Weaver, C. T. Citrobacter rodentium: a model enteropathogen for understanding the interplay of innate and adaptive components of type 3 immunity. Mucosal Immunol. 10, 1108–1117 (2017).

    CAS  PubMed  PubMed Central  Google Scholar 

  23. Di Luccia, B., Gilfillan, S., Cella, M., Colonna, M. & Huang, S. C. Ilc3s integrate glycolysis and mitochondrial production of reactive oxygen species to fulfill activation demands. J. Exp. Med. 216, 2231–2241 (2019).

    CAS  PubMed  PubMed Central  Google Scholar 

  24. Fu, Z. et al. Mitochondrial transcription factor a in rorγt(+) lymphocytes regulate small intestine homeostasis and metabolism. Nat. Commun. 12, 4462 (2021).

    CAS  PubMed  PubMed Central  Google Scholar 

  25. Rothhammer, V. & Quintana, F. J. The aryl hydrocarbon receptor: an environmental sensor integrating immune responses in health and disease. Nat. Rev. Immunol. 19, 184–197 (2019).

    CAS  PubMed  Google Scholar 

  26. Zhou, L. Ahr function in lymphocytes: emerging concepts. Trends Immunol. 37, 17–31 (2016).

    PubMed  Google Scholar 

  27. Stockinger, B., Di Meglio, P., Gialitakis, M. & Duarte, J. H. The aryl hydrocarbon receptor: multitasking in the immune system. Annu. Rev. Immunol. 32, 403–432 (2014).

    CAS  PubMed  Google Scholar 

  28. Qiu, J. et al. The aryl hydrocarbon receptor regulates gut immunity through modulation of innate lymphoid cells. Immunity 36, 92–104 (2012).

    CAS  PubMed  Google Scholar 

  29. Lee, J. S. et al. Ahr drives the development of gut ilc22 cells and postnatal lymphoid tissues via pathways dependent on and independent of notch. Nat. Immunol. 13, 144–151 (2011).

    PubMed  PubMed Central  Google Scholar 

  30. Kiss, E. A. et al. Natural aryl hydrocarbon receptor ligands control organogenesis of intestinal lymphoid follicles. Science 334, 1561–1565 (2011).

    CAS  PubMed  Google Scholar 

  31. Quast, C. et al. The silva ribosomal RNA gene database project: Improved data processing and web-based tools. Nucleic Acids Res. 41, D590–D596 (2013).

    CAS  PubMed  Google Scholar 

  32. Robinette, M. L. et al. Transcriptional programs define molecular characteristics of innate lymphoid cell classes and subsets. Nat. Immunol. 16, 306–317 (2015).

    CAS  PubMed  PubMed Central  Google Scholar 

  33. Wang, Z. et al. Iron drives T helper cell pathogenicity by promoting rna-binding protein pcbp1-mediated proinflammatory cytokine production. Immunity 49, 80–92 (2018).

    CAS  PubMed  Google Scholar 

  34. Wang, W. et al. Cd8(+) T cells regulate tumour ferroptosis during cancer immunotherapy. Nature 569, 270–274 (2019).

    CAS  PubMed  PubMed Central  Google Scholar 

  35. Muckenthaler, M. U., Galy, B. & Hentze, M. W. Systemic iron homeostasis and the iron-responsive element/iron-regulatory protein (ire/irp) regulatory network. Annu. Rev. Nutr. 28, 197–213 (2008).

    CAS  PubMed  Google Scholar 

  36. Allan, D. S. et al. An in vitro model of innate lymphoid cell function and differentiation. Mucosal Immunol. 8, 340–351 (2015).

    CAS  PubMed  Google Scholar 

  37. Xu, W. et al. Lethal cardiomyopathy in mice lacking transferrin receptor in the heart. Cell Rep. 13, 533–545 (2015).

    PubMed  PubMed Central  Google Scholar 

  38. Xiong, L. et al. Ahr-foxp3-rorγt axis controls gut homing of cd4(+) T cells by regulating gpr15. Sci. Immunol. 5, eaaz7277 (2020).

    CAS  PubMed  PubMed Central  Google Scholar 

  39. Li, S. et al. Ikaros inhibits group 3 innate lymphoid cell development and function by suppressing the aryl hydrocarbon receptor pathway. Immunity 45, 185–197 (2016).

    PubMed  PubMed Central  Google Scholar 

  40. 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).

    CAS  PubMed  PubMed Central  Google Scholar 

  41. Bostick, J. W. et al. Dichotomous regulation of group 3 innate lymphoid cells by nongastric helicobacter species. Proc. Natl Acad. Sci. USA 116, 24760–24769 (2019).

    CAS  PubMed  PubMed Central  Google Scholar 

  42. Zelante, T. et al. Tryptophan catabolites from microbiota engage aryl hydrocarbon receptor and balance mucosal reactivity via interleukin-22. Immunity 39, 372–385 (2013).

    CAS  PubMed  Google Scholar 

  43. Cervantes-Barragan, L. et al. Lactobacillus reuteri induces gut intraepithelial cd4(+)cd8αα(+) t cells. Science 357, 806–810 (2017).

    CAS  PubMed  PubMed Central  Google Scholar 

  44. Ye, J. et al. The aryl hydrocarbon receptor preferentially marks and promotes gut regulatory T cells. Cell Rep. 21, 2277–2290 (2017).

    CAS  PubMed  PubMed Central  Google Scholar 

  45. Li, S. et al. Aryl hydrocarbon receptor signaling cell intrinsically inhibits intestinal group 2 innate lymphoid cell function. Immunity 49, 915–928 (2018).

    CAS  PubMed  PubMed Central  Google Scholar 

  46. Motamedi, M., Xu, L. & Elahi, S. Correlation of transferrin receptor (CD71) with ki67 expression on stimulated human and mouse T cells: the kinetics of expression of T cell activation markers. J. Immunol. Methods 437, 43–52 (2016).

    CAS  PubMed  Google Scholar 

  47. Weinberg, E. D. Nutritional immunity. Host’s attempt to withold iron from microbial invaders. JAMA 231, 39–41 (1975).

    CAS  PubMed  Google Scholar 

  48. Sakamoto, K. et al. IL-22 controls iron-dependent nutritional immunity against systemic bacterial infections. Sci. Immunol. 2, eaai8371 (2017).

    PubMed  PubMed Central  Google Scholar 

  49. Kortman, G. A. et al. Low dietary iron intake restrains the intestinal inflammatory response and pathology of enteric infection by food-borne bacterial pathogens. Eur. J. Immunol. 45, 2553–2567 (2015).

    CAS  PubMed  PubMed Central  Google Scholar 

  50. Müllner, E. W., Neupert, B. & Kühn, L. C. A specific mRNA binding factor regulates the iron-dependent stability of cytoplasmic transferrin receptor mRNA. Cell 58, 373–382 (1989).

    PubMed  Google Scholar 

  51. Koeller, D. M. et al. A cytosolic protein binds to structural elements within the iron regulatory region of the transferrin receptor mrna. Proc. Natl Acad. Sci. USA 86, 3574–3578 (1989).

    CAS  PubMed  PubMed Central  Google Scholar 

  52. Jabara, H. H. et al. A missense mutation in tfrc, encoding transferrin receptor 1, causes combined immunodeficiency. Nat. Genet. 48, 74–78 (2016).

    CAS  PubMed  Google Scholar 

  53. Fernandez-Salguero, P. et al. Immune system impairment and hepatic fibrosis in mice lacking the dioxin-binding ah receptor. Science 268, 722–726 (1995).

    CAS  PubMed  Google Scholar 

  54. Eberl, G. et al. An essential function for the nuclear receptor rorgamma(t) in the generation of fetal lymphoid tissue inducer cells. Nat. Immunol. 5, 64–73 (2004).

    CAS  PubMed  Google Scholar 

  55. Aydemir, T. B. et al. Intestine-specific deletion of metal transporter zip14 (slc39a14) causes brain manganese overload and locomotor defects of manganism. Am. J. Physiol. Gastrointest. Liver Physiol. 318, G673–G681 (2020).

    CAS  PubMed  PubMed Central  Google Scholar 

  56. Guo, X. et al. Induction of innate lymphoid cell-derived interleukin-22 by the transcription factor stat3 mediates protection against intestinal infection. Immunity 40, 25–39 (2014).

    CAS  PubMed  PubMed Central  Google Scholar 

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

    CAS  PubMed  Google Scholar 

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

    CAS  Google Scholar 

  59. 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).

    PubMed  PubMed Central  Google Scholar 

  60. Langmead, B. & Salzberg, S. L. Fast gapped-read alignment with bowtie 2. Nat. Methods 9, 357–359 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  61. Li, H. et al. The sequence alignment/map format and samtools. Bioinformatics 25, 2078–2079 (2009).

    PubMed  PubMed Central  Google Scholar 

  62. Quinlan, A. R. & Hall, I. M. Bedtools: a flexible suite of utilities for comparing genomic features. Bioinformatics 26, 841–842 (2010).

    CAS  PubMed  PubMed Central  Google Scholar 

  63. Meyer, L. R. et al. The UCSC genome browser database: extensions and updates 2013. Nucleic Acids Res. 41, D64–D69 (2013).

    CAS  PubMed  Google Scholar 

  64. Patel, R. K. & Jain, M. NGS QC toolkit: a toolkit for quality control of next generation sequencing data. PLoS ONE 7, e30619 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  65. 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).

    CAS  PubMed  PubMed Central  Google Scholar 

  66. Lakhal-Littleton, S. et al. Cardiac ferroportin regulates cellular iron homeostasis and is important for cardiac function. Proc. Natl Acad. Sci. USA 112, 3164–3169 (2015).

    CAS  PubMed  PubMed Central  Google Scholar 

  67. Kanamori, Y. et al. Identification of novel lymphoid tissues in murine intestinal mucosa where clusters of c-kit+ il-7r+ thy1+ lympho-hemopoietic progenitors develop. J. Exp. Med. 184, 1449–1459 (1996).

    CAS  PubMed  Google Scholar 

Download references

Acknowledgements

We thank the Zhou laboratory for help and suggestions; R. Merkel and J. Kyu for mice genotyping; S. G. Swarts from the Department of Radiation Oncology of the University of Florida for assistance of HESKA; and Y. Sun from the Molecular Pathology Core, Department of Pathology of the University of Florida for tissue section and immunofluorescence staining support. Confocal microscopy was performed at the Center for Immunology & Transplantation. The Genomics Facility (University of Chicago) provided sequencing services and assistance. L.Z. is an Investigator in the Pathogenesis of Infectious Disease and is supported by the Burroughs Wellcome Fund. This work was supported by the National Institutes of Health grants AI132391 and AI157109 (to L.Z.).

Author information

Authors and Affiliations

Authors

Contributions

L.X. and L.Z. designed the study. L.X. performed experiments. E.Y.H. and N.S. assisted with part of the experiments. E.Y.H. and J.W.D. performed sequencing data analysis. F.R.J.R. provided reagents. L.Z. supervised the study. L.X. and L.Z. wrote the manuscript with the input from other authors.

Corresponding author

Correspondence to Liang Zhou.

Ethics declarations

Competing interests

The authors declare no competing interests.

Peer review

Peer review information

Nature Immunology thanks the anonymous reviewers for their contribution to the peer review of this work. Primary Handling Editor: S. Houston, in collaboration with the Nature Immunology team.

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 Iron-depleted diet impairs gut ILC3 maintenance and function (Related to Fig. 1).

3-week-old C57BL/6 wild-type (WT) mice were fed an iron-deficient (ID) or control (Ctrl) diet for 3 weeks before analysis. a and b, Hemoglobin concentrations (a) and MCV (b) in the indicated mice. Data are compiled from three independent experiments (n = 8 mice per group). c, Principal component analysis (PCA) plot of the RNA-seq for SI ILC3s in the indicated mice. d, Scatter-plot of log2 (FPKM) gene expression of FACS-sorted SI ILC3s from the indicated mice in RNA-seq analysis. DEGs (ID diet vs Ctrl diet, fold change ≥ 1.5, q value ≤ 0.05) are highlighted in blue (upregulated) and orange (downregulated), respectively (n = 3 mice per group). e, Heat map of the indicated gene expression in glycolysis pathway revealed by RNA-seq analysis. f, FACS analysis of Ahr expression in ILC3s of the mice with indicated treatments. Ahr–/– ILC3s were used as a negative control. Representative FACS plot of three independent experiments. g, Representative FACS gating/sorting strategies used for identification of different immune cell populations in this study.

Source data

Extended Data Fig. 2 Iron chelation represses ILC3 functions ex vivo and in vitro (Related to Fig. 2).

al, SI LPLs from WT mice were treated ex vivo with or without DFO for 16 h. Percentages of IFN-γ+ in ILC3s compiled from two independent experiments (n = 6 mice per group) (a), gMFI of MitoTracker Green FM (b), TMRE (c) and MitoSox (d) in ILC3s (n = 6 mice per group from two independent experiments). FACS analysis of CD71 expression in ILC2s and representative FACS plot of three independent experiments (e). Representative FACS plots (f), percentages of IL-5+ (g) and IL-13+ (h) cells in ILC2s compiled from two independent experiments (n = 6 mice per group) are shown. FACS analysis of CD71 expression in ILC1s and representative FACS plot of three independent experiments (i). Percentages of IFN-γ+ cells in ILC1s (j), IL-22+ (k) and IL-17A+ (l) cells in CD4+ T cells compiled from three independent experiments (n = 6 mice per group). ms, LI LPLs from WT mice were treated ex vivo with or without DFO for 16 h. FACS analysis of CD71 (m) and RORγt (o) expression in ILC3s and representative FACS plots of three independent experiments are shown. gMFI of CD71 (n) and RORγt (p) expression in ILC3s from three independent experiments (n = 3 mice per group) are shown. FACS analysis of IL-22 and IL-17A expression in ILC3s and representative FACS plot of three independent experiments (q). Percentages of IL-22+ (r) and IL-17A+ (s) in ILC3s from three independent experiments (n = 3 mice per group). t and u, MNK-3 cells were treated with or without DFO for 16 h in vitro. gMFI of Ki-67 expression in MNK-3 cells (t) and percentages of MNK-3 cell viability (u) (n = 8 biologically independent replicates per group from three independent experiments).

Source data

Extended Data Fig. 3 Iron depletion impairs mitochondrial function and switches OXPHOS to glycolysis in ILC3s (Related to Fig. 3).

a, PCA of RNA-seq analysis of MNK-3 cells treated with or without DFO for 16 h in vitro (n = 3 biologically independent replicates per group). bg, Mitochondrial mass and activity in MNK-3 cells. MitoTracker Green FM (b), TMRE (d) and MitoSox (f) intensity. Representative FACS plots of three independent experiments. gMFI of MitoTracker Green FM (c), TMRE (e) and MitoSox (g) (n = 6 biologically independent replicates per group). Data are compiled from three independent experiments. hj, Mitochondrial mass and activity of ILC3s from WT mice fed an ID or Ctrl diet. gMFI of MitoTracker Green FM (h), TMRE (i) and MitoSox (j) in ILC3s (control diet: n = 7; ID diet: n = 5). Data are compiled from three independent experiments. k, FACS analysis of 2-NBDG uptake in MNK-3 cells. Representative FACS plot of three independent experiments. ln, Seahorse metabolic flux analysis. Real-time ECAR (l) are representative of two independent experiments, and compiled data on quantification of basal (m) and maximal (n) ECAR of one experiment are shown (n = 3 biologically independent replicates per group). o and p, FACS analysis of IL-22 and IL-17A expression in MNK-3 cells transduced with the indicated retroviral constructs. The cells were treated with FICZ on day 2. Representative FACS plot (o) and percentages of IL-22+ cells in MNK-3 cells from two independent experiments (n = 6 biologically independent replicates per group) (p). q and r, FACS analysis of IL-22 and IL-17A expression in MNK-3 cells transduced with the indicated retroviral constructs. Representative FACS plot (q) and percentages of IL-22+ cells in MNK-3 cells compiled from two independent experiments (n = 3 biologically independent replicates per group) (r).

Source data

Extended Data Fig. 4 CD71 deficiency impairs ILC3 maintenance in a cell-intrinsic manner (Related to Fig. 4).

a, Tfrc mRNA determined by RT-qPCR in SI ILC3s. Data are compiled from three independent experiments (n = 3 mice per group). b and c, FACS analysis of intracellular Fe2+ in ILC3s. Representative FACS plot (b) and gMFI of Fe2+ in ILC3s (n = 3 mice per group) (c). d, Percentages of TH17 cells (CD4+TCRβ+Foxp3GATA3RORγt+) in CD4+ T cells from three independent experiments (Tfrcf/f: n = 6; Tfrcf/f Rorc-cre: n = 7). ek, FACS analysis of ILC3 subsets. NKp46 and RORγt expression in Lin (e), CD4 and RORγt expression in LinNKp46 (g), NKp46 and CCR6 expression in LinRORγt+ (j) LPLs. Percentages of NKp46+ ILC3s (f), CD4+ ILC3s (h) and CD4 ILC3s in Lin (i), and CCR6+ in ILC3s (k) from three independent experiments (Tfrcf/f: n = 9; Tfrcf/f Rorc-cre: n = 7). ln, FACS analysis of ILC3 cytokines. Representative FACS plot (l), percentages of IL-22+ (m) and IL-17A+ (n) cells in ILC3s from three independent experiments (Tfrcf/f: n = 9; Tfrcf/f Rorc-cre: n = 7). o and p, Percentages of Aqua+ in ILC3s from three independent experiments (Tfrcf/f: n = 7 per age; Tfrcf/f Rorc-cre: n = 6 (1-week-old and 4-week-old) or 5 (6-week-old)). q, FACS analysis of CD71 expression in CD4+ T cells and ILC3s. Representative FACS plots of three independent experiments. r and s, FACS analysis of RORγt expression in Lin LPLs. Representative FACS plot (r) and percentages of ILC3s from three independent experiments (Rag1−/− Tfrcf/f: n = 7; Rag1−/− Tfrcf/f Rorc-cre: n = 6) (s). t and u, FACS analysis of RORγt expression in Lin LPLs. Representative FACS plot (t) and percentages of ILC3s from three independent experiments (Tfrcf/f: n = 8; Tfrcf/f Cd4-cre: n = 6) (u).

Source data

Extended Data Fig. 5 CD71 ablation in ILC3s impairs gut immunity to C. rodentium infection (Related to Fig. 5).

Body weight (a) and survival (b) of C. rodentium-infected mice (Tfrcf/f: n = 8; Tfrcf/f Rorc-cre: n = 5). Statistical analysis by the log-rank test. Data are compiled from three independent experiments. c, TH17 percentages in CD4+ T cells in C. rodentium-infected mice from two independent experiments (for each time point: Tfrcf/f: n = 8 and 6; Tfrcf/f Rorc-cre: n = 4). d and e, FACS analysis of CD71 in ILC3s of WT (Tfrcf/f) mice treated with or without iron dextran. Representative FACS plot (d) and percentages of CD71+ cells in ILC3s from three independent experiments (n = 5 for untreated; n = 7 for treated) (e). f, Heat map of OXPHOS pathway and ILC3 signature genes by RNA-seq analysis of SI ILC3s (n = 3 mice per group). g, GSEA showing glycolysis gene pathway enrichment in the indicated ILC3s. h, Representative FACS plot of pS6 in ILC3s of three independent experiments. ik, FACS analysis of mitochondrial mass and activity. gMFI of MitoTracker Green FM (i) (n = 3 mice per group), TMRE (j) (Tfrcf/f: n = 6; Tfrcf/f Rorc-cre: n = 4), MitoSox (k) (Tfrcf/f: n = 6; Tfrcf/f Rorc-cre: n = 5) in ILC3s from three independent experiments. lq, RNA-seq analysis of SI ILC3s from mice with or without iron dextran treatment (n = 3 mice per group). Heat map of the DEGs by RNA-seq analysis in the indicated ILC3s shown in Fig. 5h (l). GSEA showing ILC3 signature gene enrichment (m). FPKM of Ahrr, Gem and Osgin1 in the indicated ILC3s (np) (n = 3 mice per group). GO analysis showing enriched top 10 pathways of the shared 235 DEGs (cKO vs WT, with or without iron dextran) shown in Fig. 5h (q).

Source data

Extended Data Fig. 6 Zip14 is a NTBI transporter in ILC3s (Related to Fig. 5).

a, FACS analysis of RORγt expression in Lin LPLs of the indicated mice, with or without iron dextran treatment. Representative FACS plot of three independent experiments are shown. b, Schematic depiction of experimental design for iron dextran pre-treatment, followed by C. rodentium infection. c, FPKM of literature-reported iron-related potential transporters in RNA-seq of SI ILC3s from WT mice (n = 3 mice per group). d, FPKM of RNA-seq of MNK-3 cells treated with or without DFO (100 µM) for 16 h in vitro (n = 3 biologically independent replicates per group). e, Expression of Slc39a14 and Slc39a8 mRNA determined by RT-qPCR in ILC3s from SI LPLs of Zipf/f and Zip14f/f Rorc-cre littermate mice. Data are representative of two independent experiments (n = 3 mice per group). f, FACS analysis of RORγt expression in Lin LPLs of the indicated mice. Representative FACS plot of three independent experiments. g, FACS analysis of RORγt expression in Lin LPLs of the indicated mice with or without iron dextran treatment. Representative FACS plot of three independent experiments. h and i, Expression of Slc39a8 and Slc11a2 mRNA determined by RT-qPCR in ILC3s from the indicated mice treated with iron dextran (n = 3 mice per group). Data are compiled from two independent experiments.

Source data

Extended Data Fig. 7 Microbiota inhibit CD71 expression in an Ahr-dependent manner (Related to Fig. 6).

ad, Fecal microbiome analysis of the indicated mice. Bacterial (Class) abundance (a) and Shannon Index measuring bacterial diversity (b) revealed by 16 S rRNA gene sequencing in the indicated cohoused mice (Tfrcf/f: n = 7; Tfrcf/f Rorc-cre: n = 6). Data are displayed as box and whisker plots showing the median and the 25th and 75th percentiles and two whiskers at the minimum and maximum Shannon Entropy Index. Statistical significance was determined using the Kruskal-Wallis test. H. apodemus (c) and H. typhlonius (d) bacterial abundance measured by qPCR in the indicated cohoused mice (Tfrcf/f: n = 7; Tfrcf/f Rorc-cre: n = 6). eg, FACS analysis of CD71 expression in ILC3s and TH17 cells in WT SPF and GF mice. Representative FACS plot (e) and percentages of CD71+ cells in ILC3s from three independent experiments (n = 7 mice per group) are shown (f). Percentages of CD71+ cells in TH17 cells from three independent experiments (n = 7 mice per group) are shown (g). Representative FACS plot of CD71 (h) and percentages of CD71+ cells in ILC3s from LI LPLs of the indicated mice treated with broad-spectrum antibiotics (abx) or control water. Data are compiled from five independent experiments (n = 10 for controls; n = 8 for abx-treated) (i). Representative FACS plot (j) and percentages of IL-22+ cells in ILC3s from four independent experiments (Ahr+/+: n = 8 and 7 for control and abx-treated groups; Ahr−/−: n = 7 and 5 for control and abx-treated groups) (k). ln, Il22 (l), Ahrr (m) and Cyp1a1 (n) mRNA determined by RT-qPCR in LI tissues of the indicated mice treated with abx or control water (n = 3 mice per group compiled from two independent experiments).

Source data

Extended Data Fig. 8 Ahr cell-intrinsically inhibits CD71 transcription in ILC3s (Related to Fig. 6).

af, CD71+ percentages in the indicated cells. Data are compiled from three independent experiments (ad: Ahr+/+: n = 4 and 8 for SI and LI; Ahr−/−: n = 4 and 6 for SI and LI); (e and f: n = 4 mice per group). gk, FACS analysis of CD71 in the indicated cells. Representative FACS plots (g and i), and CD71+ percentages in ILC1s (h), ILC2s (j), TH17 cells (k) from three independent experiments (Ahr+/+: n = 6; AhrdCAIR/+: n = 7; AhrdCAIR/dCAIR: n = 8). lq, FACS analysis of CD71 in the indicated ILC3s with or without IL-23 and IL-1β treatment for 6 hours ex vivo. Representative FACS plot (l) and CD71+ percentages in ILC3s compiled from three independent experiments (Ahr+/+: n = 10 per group; Ahr−/−: n = 8 per group) (m). Tfrc mRNA (n), gMFI of the Ahr (o), Ahr (p) and Ahrr (q) mRNAs determined by RT-qPCR in the indicated SI ILC3s (n = 6 mice per group from three independent experiments). rt, Mixed bone marrow (BM) experiment. Data are compiled from two independent experiments (n = 3 mice per group). A schematic depiction of experimental design (note: donor cells were age and sex matched; recipients and donors were sex matched) (r), and CD71+ percentages in the indicated ILC3s (s) and ILC2s (t). u and v, FACS analysis of CD71 in ILC3s. Representative FACS plot (u) and CD71+ percentages in ILC3s from two independent experiments (n = 3 mice per group) (v). w and x, FACS analysis of IL-17A and IL-22 in ILC3s. Absolute numbers of IL-17A+ (w) and IL-22+ (x) ILC3s in the indicated mice (n = 5 mice per group from two independent experiments).

Source data

Extended Data Fig. 9 Iron treatment alleviates ILC3 defects in Ahr-deficient mice (Related to Fig. 6).

a and b, LI-ILC3 percentages in Lin LPLs (a) and absolute numbers (b) in the indicated mice with or without iron dextran treatment. Data are compiled from three independent experiments (Ahr+/+: n = 7 and 8 for untreated and treated; Ahr−/−: n = 5 and 6 for untreated and treated). ce, FACS analysis of IL-22 and IL-17A expression in SI ILC3s of the indicated mice. Representative FACS plot (c), IL-22+ (d) and IL-17A+ (e) percentages in the indicated ILC3s from three independent experiments (Ahr+/+: n = 7 and 8 for untreated and treated; Ahr−/−: n = 5 and 6 for untreated and treated). f and g, FACS analysis of RORγt in Lin LPLs of the mice with indicated treatment. Representative FACS plot (f) and LI-ILC3 percentages in Lin LPLs from three independent experiments (g) (Ahrf/f: n = 6 and 5 for untreated and treated; Ahrf/f Rorc-cre: n = 6 mice per group). h and i, Numbers of CCR6+ ILC3s (h) and NKp46+ ILC3s (i) in SI LPLs of the indicated mice from three independent experiments (Ahrf/f: n = 6 and 5 for untreated and treated; Ahrf/f Rorc-cre: n = 6 mice per group). j, Cryptopatches (CPs)/intestinal lymphoid follicles (ILFs) in representative SI sections of the indicated mice were stained with RORγt (red), CD3 (green), and DAPI (blue) and analyzed by fluorescence microscopy. Scale bar, 100 µm. Data are representative of two independent experiments. k, Numbers of CPs/ILFs in the SI of the indicated mice (n = 3 mice per group compiled from two independent experiments). l, Percentages of Aqua+ in SI ILC3s from three independent experiments (Ahrf/f: n = 6 and 5 for untreated and treated; Ahrf/f Rorc-cre: n = 6 mice per group).

Source data

Extended Data Fig. 10 Ahr regulates CD71 transcription by directly binding to the Tfrc locus (Related to Fig. 7).

ae, RNA-seq analysis of ILC3s from SI and LI LPLs of Rag1−/−Ahr+/+ (WT) and Rag1−/−Ahr−/− (KO) littermate mice (n = 2 mice per group). PCA plot revealed by RNA-seq analysis of ILC3s from the indicated mice (a). Heat maps of the indicated mitochondria-encoding gene expression revealed by RNA-seq analysis (b). GSEA showing enrichment of ILC3 signature genes in SI (c) and LI (d) ILC3s in RNA-seq analysis. Heat maps of the ILC3 signature gene expression revealed by RNA-seq analysis (e). f, Schematic depiction of the Ahr motifs (highlighted in red) in the Ahr ChIP-seq peak at the Tfrc promoter. g, The firefly luciferase pTfrc-pGL2B or pTfrcΔXRE-pGL2B constructs and pRL-TK-Renilla luciferase construct were co-transfected into HEK293T cells, together with Ahr-expression or control plasmids. Cells were stimulated with or without FICZ and the firefly luciferase activity was normalized to Renilla luciferase activity. Representative data of two independent experiments (n = 3 biologically independent replicates per group from one experiment) are shown.

Source data

Supplementary information

Source data

Source Data Fig. 1

Statistical Source Data.

Source Data Fig. 2

Statistical Source Data.

Source Data Fig. 3

Statistical Source Data.

Source Data Fig. 4

Statistical Source Data.

Source Data Fig. 5

Statistical Source Data.

Source Data Fig. 6

Statistical Source Data.

Source Data Fig. 7

Statistical Source Data.

Source Data Extended Data Fig. 1

Statistical Source Data.

Source Data Extended Data Fig. 2

Statistical Source Data.

Source Data Extended Data Fig. 3

Statistical Source Data.

Source Data Extended Data Fig. 4

Statistical Source Data.

Source Data Extended Data Fig. 5

Statistical Source Data.

Source Data Extended Data Fig. 6

Statistical Source Data.

Source Data Extended Data Fig. 7

Statistical Source Data.

Source Data Extended Data Fig. 8

Statistical Source Data.

Source Data Extended Data Fig. 9

Statistical Source Data.

Source Data Extended Data Fig. 10

Statistical 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

Xiong, L., Helm, E.Y., Dean, J.W. et al. Nutrition impact on ILC3 maintenance and function centers on a cell-intrinsic CD71–iron axis. Nat Immunol 24, 1671–1684 (2023). https://doi.org/10.1038/s41590-023-01612-z

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41590-023-01612-z

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