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:

Functional reprogramming of regulatory T cells in the absence of Foxp3

Abstract

Regulatory T cells (Treg cells) deficient in the transcription factor Foxp3 lack suppressor function and manifest an effector T (Teff) cell–like phenotype. We demonstrate that Foxp3 deficiency dysregulates metabolic checkpoint kinase mammalian target of rapamycin (mTOR) complex 2 (mTORC2) signaling and gives rise to augmented aerobic glycolysis and oxidative phosphorylation. Specific deletion of the mTORC2 adaptor gene Rictor in Foxp3-deficient Treg cells ameliorated disease in a Foxo1 transcription factor–dependent manner. Rictor deficiency re-established a subset of Treg cell genetic circuits and suppressed the Teff cell–like glycolytic and respiratory programs, which contributed to immune dysregulation. Treatment of Treg cells from patients with FOXP3 deficiency with mTOR inhibitors similarly antagonized their Teff cell–like program and restored suppressive function. Thus, regulatory function can be re-established in Foxp3-deficient Treg cells by targeting their metabolic pathways, providing opportunities to restore tolerance in Treg cell disorders.

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: Inactivation of mTORC2, but not mTORC1, in ∆Treg cells mitigates Foxp3 deficiency.
Fig. 2: mTORC2 blockade endows ∆Treg cells with regulatory function.
Fig. 3: Cell-intrinsic and -extrinsic determinants of the Teff cell–like phenotype of ∆Treg cells.
Fig. 4: mTORC2-dependent and -independent gene expression profiles in ∆Treg cells.
Fig. 5: Contribution of the AKT–Foxo1 axis to the rictor-dependent ∆Treg cell phenotype.
Fig. 6: mTORC2 promotes aerobic glycolysis and OXPHOS in ∆Treg cells.
Fig. 7: Blockade of glycolysis improves the scurfy phenotype of Foxp3ΔEGFPiCre mice.
Fig. 8: mTOR inhibition augments the suppressive function of human FOXP3 mutant Treg cells.

Similar content being viewed by others

Data availability

The data presented in the manuscript, including de-identified patient results, will be made available to investigators following request to the corresponding author. Any data and materials to be shared will be released via a material transfer agreement. RNA sequencing datasets have been deposited in the Gene Expression Omnibus with the accession code GSE129472.

References

  1. Josefowicz, S. Z., Lu, L. F. & Rudensky, A. Y. Regulatory T cells: mechanisms of differentiation and function. Annu. Rev. Immunol. 30, 531–564 (2012).

    Article  CAS  Google Scholar 

  2. Nutsch, K. M. & Hsieh, C. S. T cell tolerance and immunity to commensal bacteria. Curr. Opin. Immunol. 24, 385–391 (2012).

    Article  CAS  Google Scholar 

  3. Lin, W. et al. Regulatory T cell development in the absence of functional Foxp3. Nat. Immunol. 8, 359–368 (2007).

    Article  CAS  Google Scholar 

  4. Gavin, M. A. et al. Foxp3-dependent programme of regulatory T-cell differentiation. Nature 445, 771–775 (2007).

    Article  CAS  Google Scholar 

  5. Ohkura, N. et al. T cell receptor stimulation-induced epigenetic changes and Foxp3 expression are independent and complementary events required for Treg cell development. Immunity 37, 785–799 (2012).

    Article  CAS  Google Scholar 

  6. Kaech, S. M. & Cui, W. Transcriptional control of effector and memory CD8+ T cell differentiation. Nat. Rev. Immunol. 12, 749–761 (2012).

    Article  CAS  Google Scholar 

  7. Rudensky, A. Y. Regulatory T cells and Foxp3. Immunol. Rev. 241, 260–268 (2011).

    Article  CAS  Google Scholar 

  8. Buck, M. D., Sowell, R. T., Kaech, S. M. & Pearce, E. L. Metabolic instruction of immunity. Cell 169, 570–586 (2017).

    Article  CAS  Google Scholar 

  9. Beier, U. H. et al. Essential role of mitochondrial energy metabolism in Foxp3+ T-regulatory cell function and allograft survival. FASEB J. 29, 2315–2326 (2015).

    Article  CAS  Google Scholar 

  10. Gerriets, V. A. et al. Metabolic programming and PDHK1 control CD4+ T cell subsets and inflammation. J. Clin. Invest. 125, 194–207 (2015).

    Article  Google Scholar 

  11. Michalek, R. D. et al. Cutting edge: distinct glycolytic and lipid oxidative metabolic programs are essential for effector and regulatory CD4+ T cell subsets. J. Immunol. 186, 3299–3303 (2011).

    Article  CAS  Google Scholar 

  12. Shi, L. Z. et al. HIF1α-dependent glycolytic pathway orchestrates a metabolic checkpoint for the differentiation of TH17 and Treg cells. J. Exp. Med. 208, 1367–1376 (2011).

    Article  CAS  Google Scholar 

  13. Angelin, A. et al. Foxp3 reprograms T cell metabolism to function in low-glucose, high-lactate environments. Cell Metab. 25, 1282–1293.e7 (2017).

    Article  CAS  Google Scholar 

  14. Gerriets, V. A. et al. Foxp3 and Toll-like receptor signaling balance Treg cell anabolic metabolism for suppression. Nat. Immunol. 17, 1459–1466 (2016).

    Article  CAS  Google Scholar 

  15. Pollizzi, K. N. & Powell, J. D. Regulation of T cells by mTOR: the known knowns and the known unknowns. Trends Immunol. 36, 13–20 (2015).

    Article  CAS  Google Scholar 

  16. Chapman, N. M. & Chi, H. mTOR links environmental signals to T cell fate decisions. Front. Immunol. 5, 686 (2014).

    PubMed  Google Scholar 

  17. Valmori, D. et al. Rapamycin-mediated enrichment of T cells with regulatory activity in stimulated CD4+ T cell cultures is not due to the selective expansion of naturally occurring regulatory T cells but to the induction of regulatory functions in conventional CD4+ T cells. J. Immunol. 177, 944–949 (2006).

    Article  CAS  Google Scholar 

  18. Liu, C., Chapman, N. M., Karmaus, P. W., Zeng, H. & Chi, H. mTOR and metabolic regulation of conventional and regulatory T cells. J. Leukoc. Biol. 97, 837–847 (2015).

    Article  CAS  Google Scholar 

  19. Delgoffe, G. M. et al. The mTOR kinase differentially regulates effector and regulatory T cell lineage commitment. Immunity 30, 832–844 (2009).

    Article  CAS  Google Scholar 

  20. Zeng, H. et al. mTORC1 couples immune signals and metabolic programming to establish Treg-cell function. Nature 499, 485–490 (2013).

    Article  CAS  Google Scholar 

  21. Sarbassov, D. D., Guertin, D. A., Ali, S. M. & Sabatini, D. M. Phosphorylation and regulation of Akt/PKB by the rictor–mTOR complex. Science 307, 1098–1101 (2005).

    Article  CAS  Google Scholar 

  22. Manning, B. D. & Toker, A. AKT/PKB signaling: navigating the network. Cell 169, 381–405 (2017).

    Article  CAS  Google Scholar 

  23. Patterson, S. J. et al. Cutting edge: PHLPP regulates the development, function, and molecular signaling pathways of regulatory T cells. J. Immunol. 186, 5533–5537 (2011).

    Article  CAS  Google Scholar 

  24. Shrestha, S. et al. Treg cells require the phosphatase PTEN to restrain TH1 and TFH cell responses. Nat. Immunol. 16, 178–187 (2015).

    Article  CAS  Google Scholar 

  25. Williams, B. L. et al. Phosphorylation of Tyr319 in ZAP-70 is required for T-cell antigen receptor-dependent phospholipase C-gamma1 and ras activation. EMBO J. 18, 1832–1844 (1999).

    Article  CAS  Google Scholar 

  26. Lin, W. et al. Allergic dysregulation and hyperimmunoglobulinemia E in Foxp3 mutant mice. J Allergy Clin. Immunol. 116, 1106–1115 (2005).

    Article  CAS  Google Scholar 

  27. Leach, M. W., Bean, A. G., Mauze, S., Coffman, R. L. & Powrie, F. Inflammatory bowel disease in C.B-17 scid mice reconstituted with the CD45RBhigh subset of CD4+ T cells. Am. J. Pathol. 148, 1503–1515 (1996).

    CAS  PubMed  PubMed Central  Google Scholar 

  28. Charbonnier, L. M., Wang, S., Georgiev, P., Sefik, E. & Chatila, T. A. Control of peripheral tolerance by regulatory T cell-intrinsic notch signaling. Nat. Immunol. 16, 1162–1173 (2015).

    Article  CAS  Google Scholar 

  29. Wan, Y. Y. & Flavell, R. A. Identifying Foxp3-expressing suppressor T cells with a bicistronic reporter. Proc. Natl Acad. Sci. USA 102, 5126–5131 (2005).

    Article  CAS  Google Scholar 

  30. Gan, X. et al. PRR5L degradation promotes mTORC2-mediated PKC-δ phosphorylation and cell migration downstream of Gα12. Nat. Cell Biol. 14, 686–696 (2012).

    Article  CAS  Google Scholar 

  31. Neumann, C. et al. Role of Blimp-1 in programing Th effector cells into IL-10 producers. J. Exp. Med. 211, 1807–1819 (2014).

    Article  CAS  Google Scholar 

  32. Plas, D. R. & Thompson, C. B. Akt activation promotes degradation of tuberin and FOXO3a via the proteasome. J. Biol. Chem. 278, 12361–12366 (2003).

    Article  CAS  Google Scholar 

  33. Ouyang, W. et al. Novel Foxo1-dependent transcriptional programs control Treg cell function. Nature 491, 554–559 (2012).

    Article  CAS  Google Scholar 

  34. Kerdiles, Y. M. et al. Foxo transcription factors control regulatory T cell development and function. Immunity 33, 890–904 (2010).

    Article  CAS  Google Scholar 

  35. Hagiwara, A. et al. Hepatic mTORC2 activates glycolysis and lipogenesis through Akt, glucokinase, and SREBP1c. Cell Metab. 15, 725–738 (2012).

    Article  CAS  Google Scholar 

  36. Wilhelm, K. et al. FOXO1 couples metabolic activity and growth state in the vascular endothelium. Nature 529, 216–220 (2016).

    Article  CAS  Google Scholar 

  37. Pearce, E. L., Poffenberger, M. C., Chang, C. H. & Jones, R. G. Fueling immunity: insights into metabolism and lymphocyte function. Science 342, 1242454 (2013).

    Article  Google Scholar 

  38. Huynh, A. et al. Control of PI(3) kinase in Treg cells maintains homeostasis and lineage stability. Nat. Immunol. 16, 188–196 (2015).

    Article  CAS  Google Scholar 

  39. Wang, R. et al. The transcription factor Myc controls metabolic reprogramming upon T lymphocyte activation. Immunity 35, 871–882 (2011).

    Article  CAS  Google Scholar 

  40. Longo, N., Frigeni, M. & Pasquali, M. Carnitine transport and fatty acid oxidation. Biochim. Biophys. Acta 1863, 2422–2435 (2016).

    Article  CAS  Google Scholar 

  41. De Bock, K. et al. Role of PFKFB3-driven glycolysis in vessel sprouting. Cell 154, 651–663 (2013).

    Article  CAS  Google Scholar 

  42. Chatila, T. A. et al. JM2, encoding a fork head-related protein, is mutated in X-linked autoimmunity-allergic disregulation syndrome. J. Clin. Invest. 106, R75–R81 (2000).

    Article  CAS  Google Scholar 

  43. Bacchetta, R. et al. Defective regulatory and effector T cell functions in patients with FOXP3 mutations. J. Clin. Invest. 116, 1713–1722 (2006).

    Article  CAS  Google Scholar 

  44. Gavin, M. A. et al. Single-cell analysis of normal and FOXP3-mutant human T cells: FOXP3 expression without regulatory T cell development. Proc. Natl Acad. Sci. USA 103, 6659–6664 (2006).

    Article  CAS  Google Scholar 

  45. Bacchetta, R., Barzaghi, F. & Roncarolo, M. G. From IPEX syndrome to FOXP3 mutation: a lesson on immune dysregulation. Ann. NY Acad. Sci. 1417, 5–22 (2018).

    Article  CAS  Google Scholar 

  46. Liu, W. et al. CD127 expression inversely correlates with FoxP3 and suppressive function of human CD4+ T reg cells. J. Exp. Med. 203, 1701–1711 (2006).

    Article  CAS  Google Scholar 

  47. Seddiki, N. et al. Expression of interleukin (IL)-2 and IL-7 receptors discriminates between human regulatory and activated T cells. J. Exp. Med. 203, 1693–1700 (2006).

    Article  CAS  Google Scholar 

  48. Zhang, L. et al. Mammalian target of rapamycin complex 2 controls CD8 T cell memory differentiation in a Foxo1-dependent manner. Cell Rep. 14, 1206–1217 (2016).

    Article  CAS  Google Scholar 

  49. Hayatsu, N. et al. Analyses of a mutant Foxp3 allele reveal BATF as a critical transcription factor in the differentiation and accumulation of tissue regulatory T cells. Immunity 47, 268–283.e9 (2017).

    Article  CAS  Google Scholar 

  50. Bin Dhuban, K. et al. Suppression by human FOXP3+ regulatory T cells requires FOXP3–TIP60 interactions. Sci. Immunol. 2, eaai9297 (2017).

    Article  Google Scholar 

  51. Bennett, C. L. et al. A rare polyadenylation signal mutation of the FOXP3 gene (AAUAAA→AAUGAA) leads to the IPEX syndrome. Immunogenetics 53, 435–439 (2001).

    Article  CAS  Google Scholar 

  52. Raymond, C. S. & Soriano, P. High-efficiency FLP and PhiC31 site-specific recombination in mammalian cells. PLoS ONE 2, e162 (2007).

    Article  Google Scholar 

  53. Rivas, M. N. et al. MyD88 is critically involved in immune tolerance breakdown at environmental interfaces of Foxp3-deficient mice. J. Clin. Invest. 122, 1933–1947 (2012).

    Article  CAS  Google Scholar 

  54. Schmitt, E. G. et al. IL-10 produced by induced regulatory T cells (iTregs) controls colitis and pathogenic ex-iTregs during immunotherapy. J. Immunol. 189, 5638–5648 (2012).

    Article  CAS  Google Scholar 

Download references

Acknowledgements

This work was supported by National Institutes of Health grants 2R01 AI065617 (to T.A.C.) and RO1 AI102888-01A1 (to M.O.L.).

Author information

Authors and Affiliations

Authors

Contributions

L.-M.C. and T.A.C. designed the experiments and evaluated the data. L.-M.C., Y.C., E.V., H.H. and D.L. performed the experiments. D.L. and M.P. performed the gene expression profiling studies. J.J.B., M.I.G.-L., K.C., A.O., M.O.L. and P.C. provided critical material. L.-M.C. and T.A.C. conceived the project and directed the research. L.-M.C. and T.A.C. wrote the manuscript.

Corresponding author

Correspondence to Talal A. Chatila.

Ethics declarations

Competing interests

The authors declare no competing interests.

Additional information

Peer review information: Zoltan 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 Derivation and characterization of Foxp3∆EGFPiCre mice.

(a) Targeting strategy. An IRES-EGFP-iCre-SV40 poly-A cassette was inserted into an SSPI restriction site in the Foxp3 3’-untranslated sequence in exon 11 immediately downstream of the Foxp3 stop codon and upstream of the poly-A adenylation signal. The PGK-neo cassette was later removed by crossing the founder mice with mice transgenic for FLP recombinase. Neo, Frt PGK-neo cassette ; DT, diphtheria toxin gene ; RV, EcoRV ; S, SspI ; H ; HindIII ; K ; KpnI (b) Southern blot analysis of EcoRV-digested and EcoRI DNA derived targeted embryonic stem (ES) cells revealed by 5’ and 3’ probes respectively. (c) PCR products obtained with cDNA after reverse transcription of mRNA or gDNA isolated from CD4+ T cells of WT or Foxp3∆EGFPiCre mice using forward and reverse primers located respectively on Exon 9 and Exon 10 of the Foxp3 locus. (d) Sanger sequencing of the PCR product obtained with cDNA in (c) from Foxp3∆EGFPiCre CD4+ T cells reveal an aberrant spicing with frt and intron 9 sequences. * represents a premature stop codon. (e) EGFP and YFP or Foxp3 (N-terminal domain) expression of CD4+ T cells from Foxp3EGFPCreR26YFP and Foxp3∆EGFPiCreR26YFP mice. (f) Cropped immunoblots of Foxp3 and Actin in GFP and GFP+ CD4+ T cells from Foxp3EGFPCre and Foxp3∆EGFPiCre mice. (g) Foxp3 expression of EGFP and EGFP+ CD4+ T cells from Foxp3EGFPCre (n=7) and Foxp3∆EGFPiCre mice (n=4), normalized on Actin expression and expressed as fold change compared to EGFP+ CD4+ T cells from Foxp3EGFPCre mice. Results are expressed as mean ± SEM and statistical significance was determined by a two-way ANOVA with Sidak’s multiple comparisons (P values as indicated) (h) Methylation status (as defined by the colored bar) of individual or global CpG motifs located at different positions on the Foxp3 Conserved non-coding sequence 2 (CNS2) and the global methylation status of CNS2 CpG motifs (total) of EGFP and EGFP+ CD4+ T cells from Foxp3EGFPCre and Foxp3∆EGFPicre mice (n=3 mice per group). (i) Foxp3 and YFP expression in CD4+CD8+ double positive (DP), CD4CD8 double negative (DN), CD4+ or CD8+ single-positive (CD4 SP, CD8 SP respectively) thymocytes or CD4+ or CD8+ splenocytes of Foxp3∆EGFPiCre/+R26YFP mice.

Supplementary Figure 2 Characterization of Rictor- and Raptor-deficient ∆Treg cells.

(a,b) Representative flow cytometric analysis (a) and mean fluorescence intensity (MFI) (b) of phosphorylated 4EBP (p4EBP) and phosphorylated AKT at the Thr308 residue (pT308AKT) expression in unstimulated (US) or anti-CD3 mAb (α-CD3) -stimulated Treg cells from Foxp3EGFPCre and Foxp3∆EGFPiCre mice (n=4 per group). Results represent one of 2 independent experiments. (c) Representative cropped immunoblots and densitometry analyses of Pten, Phlpp1 and Actin in EGFP and EGFP+ CD4+ T cells from Foxp3EGFPCre and Foxp3∆EGFPiCre mice (n=3 per group). Results represent a pool of 3 independent experiments. (d) Cropped immunoblots and densitometry analyses of pY319-ZAP70 and total ZAP70 in EGFP+ CD4+ T cells from Foxp3EGFPCre and Foxp3∆EGFPiCre mice after 0, 2, 5 and 10 min of anti-CD3 stimulation (1µg/mL) (n=3 per group). Results represent a pool of 3 independent experiments. (e) Rptor gene expression of EGFP and EGFP+ CD4+ T cells isolated from Foxp3EGFPCre, Foxp3∆EGFPiCre and Foxp3∆EGFPicreRptor∆/∆ mice. The values were normalized for Actin expression and expressed as fold change compared to EGFPCD4+ T cells from Foxp3EGFPCre mice (n=3 per group). (f) Rictor gene expression of EGFP and EGFP+ CD4+ T cells isolated from Foxp3EGFPCre (n=8), Foxp3∆EGFPiCre (n=8) and Foxp3∆EGFPiCreRictor∆/∆ mice (n=7), normalized for Actin expression and expressed as fold change compared to EGFPCD4+ T cells from Foxp3EGFPCre mice. (g) Immunoblots of Rictor and Actin in EGFP and EGFP+ CD4+ T cells from Foxp3EGFPCre, Foxp3∆EGFPiCre and Foxp3∆EGFPiCreRictor∆/∆ mice. (n=3 per group). (h) Frequencies of Treg cells (EGFP+CD4+) from Foxp3EGFPCre (n=15), Foxp3∆EGFPicre (n=14), Foxp3∆EGFPicreRictor∆/∆ (n=17) and Foxp3∆EGFPicreRptor∆/∆ (n=5) mice. (P values as indicated) Results represent a pool of 5 independent experiments. (i, j) Representative flow cytometric analysis (i) and scatter plot representation (j) of Foxp3, CD25, CTLA-4, ICOS, Helios, Nrp1, GITR and CD73 MFI in Foxp3EGFPCre Teff cells and in Foxp3EGFPCre (n=4, except for CD73, n=8), Foxp3∆EGFPiCre (n=4, except for CD73, n=11) and Foxp3∆EGFPiCreRictor∆/∆ (n=4, except for CD73, n=6) Treg cells. Results represent one of 2 independent experiments, except for CD73 which represent a pool of 2 independent experiments. Statistical significance was determined by a one-way ANOVA with Tukey’s multiple comparisons (P values as indicated). Statistical significance was determined by a one-way ANOVA with Tukey’s multiple comparisons (h, j) or a two-way ANOVA with Sidak’s multiple comparisons (b, c, d, e, f, g) (P values as indicated).

Supplementary Figure 3 Characterization of the peripheral T cell compartment of Foxp3∆EGFPiCreRictor∆/∆ cells.

(a, b) Representative flow cytometric analysis (a) and frequencies (b) of CD62LhiCD44lo, CD62LhiCD44hi and CD62LloCD44hi CD8+ T cells, T-Bet, Gata-3, Rorγt and IL-17, Ki67 and EGFP expression by CD4+ Treg (YFP+) and Teff (YFP) cells from spleens of Foxp3EGFPCre, Foxp3∆EGFPiCre and Foxp3∆EGFPiCreRictor∆/∆ mice (n=4–8). Results represent one of 3 independent experiments. Statistical significance was determined by a one-way ANOVA with Tukey’s multiple comparisons (P values as indicated). Ex-Treg cells represent the proportion of EGFP among YFP+ cells.

Supplementary Figure 4 Rictor-deficient Foxp3-sufficient Treg cells have superior in vivo suppressive capacities in CD4 T cell transfer-induced colitis model.

(a) Schematic representation of the protocol of lymphopenia-induced colitis by intraperitoneal adoptive transfer of Teff cells (CD45.1+CD4+CD45RBhiEGFP) from CD45.1 Foxp3EGFP mice without or with Treg cells (CD45.2+CD4+YFP+) from Foxp3YFPcre or Foxp3YFPcreRictor∆/∆ mice in Rag1–/– mice. (b) Changes in body weight over time are shown. (n=15 for “No Treg cells” group, n=17 for “Foxp3YFPcre Treg cells” group and n=18 for “Foxp3YFPcre Rictor∆/∆ Treg cells” group) (c) Absolute number of CD45.1+ T cells (Teff cell input) and CD45.2+ T cells (Treg cell input) within the entire colon (n=10 for “No Treg cells” group, n=12 for “Foxp3YFPcre Treg cells” group and n=12 for “Foxp3YFPcre Rictor∆/∆ Treg cells” group). (d, e) Disease severity was evaluated by H&E staining of colon sections (n=19 for “No Treg cells” group, n=17 for “Foxp3YFPcre Treg cells” group and n=17 for “Foxp3YFPcre Rictor∆/∆ Treg cells” group) (d) and end point colon length (n=11 for “No Treg cells” group, n=13 for “Foxp3YFPcre Treg cells” group and n=17 for “Foxp3YFPcre Rictor∆/∆ Treg cells” group) (e). (f, g) Representative flow cytometric analysis (f) and absolute number (g) of IFN-γ and IL-17-producing CD45.1+CD4+ T cells from the Teff cell input was quantified within the entire colon (n=11 for “No Treg cells” group, n=13 for “Foxp3YFPcre Treg cells” group and n=13 for “Foxp3YFPcre Rictor∆/∆ Treg cells” group). Statistical significance was determined by a one-way ANOVA with Tukey’s multiple comparisons (c, e, g) or by a two-way repeted mesure ANOVA with Tukey’s multiple comparisons (b) (P values as indicated). Results represent a pool of 3 independent experiments.

Supplementary Figure 5 Contribution of the AKT/Foxo1 axis to the Rictor-dependent phenotype of ∆Treg cell.

(a-c) Gross appearance (a), body weight (b) and survival (c) of WT (n=8 for body weight), Foxp3∆EGFPiCre (n=10 for body weight), Foxp3∆EGFPiCreRictor∆/∆ (n=6 for body weight), Foxp3∆EGFPiCreRictor∆/∆ Foxo1∆/∆ (n=11 for body weight) and Foxp3∆EGFPiCreR26Foxo1AAA (n=10 for body weight) mice. Results represent a pool of 3 independent experiments (d) Frequencies of T-Bet+ and Gata-3+ by CD4+ Treg and Teff cells in spleens of Foxp3EGFPCreR26YFP (n=3), Foxp3∆EGFPiCreR26YFP(n=4), Foxp3∆EGFPiCreRictor∆/∆R26YFP (n=5) and Foxp3∆EGFPiCreR26Foxo1AAA (n=5) mice. Results represent one of 3 independent experiments (e, f) Representative flow cytometric analysis (e) and frequencies (f) of IL-10+ by CD4+ Treg cells in spleens of Foxp3EGFPCreR26YFP(n=4), Foxp3∆EGFPiCreRictor∆/∆R26YFP (n=3) and Foxp3∆EGFPiCreRictor∆/∆ Foxo1∆/∆ (n=3) mice. Results represent one of 2 independent experiments. (g) Gross appearance of WT, Foxp3∆EGFPiCre, Foxp3∆EGFPiCreRictor∆/∆ and Foxp3∆EGFPiCreR26Foxo1AAA mice. (h) In vitro suppression of the proliferation of WT CD4+ Teff cells (Tresp) by Foxp3EGFPCre, Foxp3∆EGFPiCre and Foxp3∆EGFPiCreR26Foxo1AAA Treg (Tsupp) cells (n=3 per group). Statistical significance was determined by a one-way ANOVA with Bonferroni’s multiple comparisons (b) or with Tukey’s multiple comparisons (d, f), two-way ANOVA with Tukey’s multiple comparisons (h) and log-rank test (c) (P values as indicated).

Supplementary Figure 6 Metabolic profiling of Foxp3-sufficient Treg cells and Rictor-sufficient and -deficient ∆Treg cells.

(a) Quantitative PCR analysis of transcripts encoding the glycolysis and pentose phosphate cycle (PPC) enzymes, normalized to Actin expression and expressed as Log2 Fold change, in Foxp3EGFP, Foxp3∆EGFPiCre and Foxp3∆EGFPiCreRictor∆/∆ Treg cells compared to WT CD4+ Teff cells. (b, c) Extracellular acidification rate (ECAR) under glycolysis stress test conditions and Oxygen consumption rate (OCR) under mitochondrial stress test conditions of Treg/∆Treg cells isolated from Foxp3EGFP/EGFP, Foxp3∆EGFPiCre/+ and Foxp3∆EGFPiCre/+Rictor∆/∆ females. (n=4 per group). Results represent a pool of 2 experiments. (d) Lactate release by in vitro cultured resting (No Stim) or stimulated (α-CD3+IL-2) Foxp3EGFP Teff cells and Foxp3EGFP, Foxp3∆EGFPiCre and Foxp3∆EGFPiCreRictor∆/∆ Treg cells. (n=6 per group). Results represent a pool of 2 experiments. (e) Hif1a and Myc gene expression in EGFP+CD4+ T cells from Foxp3EGFP (n=5), Foxp3∆EGFPiCre (n=3) and Foxp3∆EGFPiCreRictor∆/∆ (n=3) mice, normalized to Actin expression and expressed as fold change compared to EGFP+CD4+ T cells from Foxp3EGFP mice. (f) Metabolic pathway analysis showing differentially expressed metabolites between Foxp3∆EGFPiCre and Foxp3EGFP Treg cells, Foxp3∆EGFPiCreRictor∆/∆ and Foxp3EGFP Treg cells and Foxp3∆EGFPiCreRictor∆/∆ and Foxp3∆EGFPiCre Treg cells (n=5 per group). Statistical significance was determined by a one-way ANOVA with Holm-Sidak’s or Bonferroni’s multiple comparisons (e) or two-way ANOVA with Sidak’s or Tukey’s multiple comparisons (c, d) (P values as indicated).

Supplementary Figure 7 Glycolytic blockade and ∆Treg cell-specific Rictor deletion synergize in controlling Foxp3 deficiency disease.

(a) Representative flow cytometric analysis of CD62LloCD44hi CD4+ and CD8+ Teff cells in PBS or 2DG-treated Foxp3∆EGFPiCre mice (Frequencies in Fig. 7h). (b, c) Representative flow cytometric analysis (b) and frequencies (c) of Foxp3∆EGFPiCreRictor∆/∆ mice treated either with PBS or 2DG (2µg/g) from day 14 to day 45 every other day. (d) Representative flow cytometric analysis and frequencies of IFN-γ+ and IL-4+ CD4+ Teff and Treg cells of PBS or 2DG-treated Foxp3∆EGFPiCre mice (Frequencies in Fig. 7h). (e, f) Representative flow cytometric analysis (e) and frequencies (f) of IFN-γ+ and IL-4+ CD4+ Teff and Treg cells of PBS or 2DG-treated Foxp3∆EGFPiCreRictor∆/∆ mice. (g) Representative flow cytometric analysis and frequencies of T-Bet+ and Gata-3+ CD4+ Teff and Treg cells of PBS or 2DG-treated Foxp3∆EGFPiCre mice (Frequencies in Fig. 7h). (h, i) Representative flow cytometric analysis (h) and frequencies (i) of T-Bet+ and Gata-3+ CD4+ Teff and Treg cells of PBS or 2DG-treated Foxp3∆EGFPiCreRictor∆/∆ mice (n=6 per group) Statistical significance was determined by an unpaired t-test (P values as indicated).

Supplementary Figure 8 Characterization of Treg cells from IPEX patients.

Representative flow cytometric analysis of CD25 and CD127 expression (a), and CD4 and FOXP3 expression (b) among CD4+ PBMCs isolated from healthy control subjects (HC1-5) and IPEX subjects (P1-5). (c) Expression of FOXP3, CD25, CD127, CTLA-4 and HELIOS in CD4+CD25loCD127hi T cells of HC (Teff cells HC1-5) and CD4+CD25hiCD127lo population from HC (Treg cells HC1-5) and IPEX patients (Treg cells P1-5).

Supplementary information

Supplementary Information

Supplementary Figs. 1–8 and Tables 1–4

Reporting Summary

Supplementary Dataset 1

Differentially expressed genes in WT (Foxp3RFP) versus mutant (Foxp3∆EGFPiCre) Treg cells derived from Foxp3∆EGFPiCre/RFP female mice (n = 4 per group).

Supplementary Dataset 2

Differentially expressed genes in single-mutant (Foxp3∆EGFPiCre) versus double-mutant (Foxp3∆EGFPiCreRictor∆/∆) Treg cells derived, respectively, from Foxp3∆EGFPiCre/RFP and Foxp3∆EGFPiCre/+Rictor∆/∆ female mice (n = 4 per group).

Source Data

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Charbonnier, LM., Cui, Y., Stephen-Victor, E. et al. Functional reprogramming of regulatory T cells in the absence of Foxp3. Nat Immunol 20, 1208–1219 (2019). https://doi.org/10.1038/s41590-019-0442-x

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41590-019-0442-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