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:

Identification of human exTreg cells as CD16+CD56+ cytotoxic CD4+ T cells

Abstract

In atherosclerosis, some regulatory T (Treg) cells become exTreg cells. We crossed inducible Treg and exTreg cell lineage-tracker mice (FoxP3eGFP−Cre-ERT2ROSA26CAG-fl-stop-fl-tdTomato) to atherosclerosis-prone Apoe/ mice, sorted Treg cells and exTreg cells and determined their transcriptomes by bulk RNA sequencing (RNA-seq). Genes that were differentially expressed between mouse Treg cells and exTreg cells and filtered for their presence in a human single-cell RNA-sequencing (scRNA-seq) panel identified exTreg cell signature genes as CST7, NKG7, GZMA, PRF1, TBX21 and CCL4. Projecting these genes onto the human scRNA-seq with CITE-seq data identified human exTreg cells as CD3+CD4+CD16+CD56+, which was validated by flow cytometry. Bulk RNA-seq of sorted human exTreg cells identified them as inflammatory and cytotoxic CD4+T cells that were significantly distinct from both natural killer and Treg cells. DNA sequencing for T cell receptor-β showed clonal expansion of Treg cell CDR3 sequences in exTreg cells. Cytotoxicity was functionally demonstrated in cell killing and CD107a degranulation assays, which identifies human exTreg cells as cytotoxic CD4+T cells.

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: Deep transcriptomes from mouse exTreg cells and Treg cells identify differentially expressed candidate genes.
Fig. 2: Mouse exTreg cell classifier genes identify human exTreg cell candidate genes and surface markers in a human scRNA-seq and CITE-seq dataset of CD4+T cells.
Fig. 3: Deep transcriptomes from sorted human CD3+CD4+CD16+CD56+ exTreg cells contrasted with Treg cells and NK cells.
Fig. 4: Oligoclonal human exTreg cells are clonally expanded from proliferating Treg cells.
Fig. 5: Human exTreg cells are not suppressive but are cytotoxic.
Fig. 6: Human exTreg cells express cytotoxic proteins, inflammatory cytokines, chemokines and chemokine receptors.
Fig. 7: Inflammatory and cytotoxic human exTreg genes overexpressed in individuals with coronary artery disease.

Similar content being viewed by others

Data availability

RNA-seq data have been uploaded to the NCBI GEO and are accessible under accession numbers GSE217010 (mouse and human bulk RNA-seq) and GSE190570 (human scRNA-seq data). Human TCR-seq data were generated and processed by Adaptive Biotechnologies. Details of productive TCR sequences, accessed through their immunoSEQ Analyzer portal, are provided in Supplementary Table 14. Source data are provided with this paper.

Code availability

No new algorithms were generated for this study.

References

  1. Tse, K. et al. Atheroprotective vaccination with MHC-II restricted peptides from ApoB-100. Front. Immunol. 4, 493 (2013).

    Article  PubMed  PubMed Central  Google Scholar 

  2. Kimura, T. et al. Atheroprotective vaccination with MHC-II-restricted ApoB peptides induces peritoneal IL-10-producing CD4 T cells. Am. J. Physiol. 312, H781–H790 (2017).

    Google Scholar 

  3. Kimura, T. et al. Regulatory CD4+ T cells recognize major histocompatibility complex class II molecule-restricted peptide epitopes of apolipoprotein B. Circulation 138, 1130–1143 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  4. Wolf, D. et al. Pathogenic autoimmunity in atherosclerosis evolves from initially protective apolipoprotein B 100 –reactive CD4+ T-regulatory cells. Circulation 142, 1279–1293 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  5. Marchini, T., Hansen, S. & Wolf, D. ApoB-specific CD4+ T cells in mouse and human atherosclerosis. Cells 10, 446 (2021).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  6. Roy, P. et al. Immunodominant MHC-II (major histocompatibility complex II) restricted epitopes in human apolipoprotein B. Circ. Res. 131, 258–276 (2022).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  7. Saigusa, R. et al. Single-cell transcriptomics and TCR reconstruction reveal CD4 T cell response to MHC-II-restricted APOB epitope in human cardiovascular disease. Nat. Cardiovasc. Res. 1, 462–475 (2022).

    Article  PubMed  PubMed Central  Google Scholar 

  8. Ait-Oufella, H., Lavillegrand, J.-R. & Tedgui, A. Regulatory T cell-enhancing therapies to treat atherosclerosis. Cells 10, 723 (2021).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  9. Roy, P., Orecchioni, M. & Ley, K. How the immune system shapes atherosclerosis: roles of innate and adaptive immunity. Nat. Rev. Immunol. 22, 251–265 (2022).

    Article  CAS  PubMed  Google Scholar 

  10. Li, J. et al. CCR5+T-bet+FoxP3+ effector CD4 T cells drive atherosclerosis. Circ. Res. 118, 1540–1552 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  11. Butcher, M. J. et al. Atherosclerosis-driven Treg plasticity results in formation of a dysfunctional subset of plastic IFNγ+ Th1/Tregs. Circ. Res. 119, 1190–1203 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  12. Qiu, R. et al. Regulatory T cell plasticity and stability and autoimmune diseases. Clin. Rev. Allergy Immunol. 58, 52–70 (2020).

    Article  CAS  PubMed  Google Scholar 

  13. Ali, A. J., Makings, J. & Ley, K. Regulatory T cell stability and plasticity in atherosclerosis. Cells 9, 2665 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  14. Zhou, X. et al. Instability of the transcription factor Foxp3 leads to the generation of pathogenic memory T cells in vivo. Nat. Immunol. 10, 1000–1007 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  15. Bailey-Bucktrout, S. L. et al. Self-antigen-driven activation induces instability of regulatory T cells during an inflammatory autoimmune response. Immunity 39, 949–962 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  16. Svensson, M. N. D. et al. Reduced expression of phosphatase PTPN2 promotes pathogenic conversion of Tregs in autoimmunity. J. Clin. Invest. 129, 1193–1210 (2019).

    Article  PubMed  PubMed Central  Google Scholar 

  17. Gaddis, D. E. et al. Apolipoprotein AI prevents regulatory to follicular helper T cell switching during atherosclerosis. Nat. Commun. 9, 1095 (2018).

    Article  PubMed  PubMed Central  Google Scholar 

  18. Hsieh, W.-C. et al. PTPN2 links colonic and joint inflammation in experimental autoimmune arthritis. JCI Insight 5, e141868 (2020).

    Article  PubMed  PubMed Central  Google Scholar 

  19. Hua, J. et al. Pathological conversion of regulatory T cells is associated with loss of allotolerance. Sci. Rep. 8, 7059 (2018).

    Article  PubMed  PubMed Central  Google Scholar 

  20. Rubtsov, Y. P. et al. Stability of the regulatory T cell lineage in vivo. Science 329, 1667–1671 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  21. Saigusa, R. et al. Sex differences in coronary artery disease and diabetes revealed by scRNA-seq and CITE-seq of human CD4+ T cells. Int. J. Mol. Sci. 23, 9875 (2022).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  22. Glanville, J. et al. Identifying specificity groups in the T cell receptor repertoire. Nature 547, 94–98 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  23. Dogra, P. et al. Tissue determinants of human NK cell development, function, and residence. Cell 180, 749–763 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  24. Ferraro, A. et al. Interindividual variation in human T regulatory cells. Proc. Natl Acad. Sci. USA 111, E1111–E1120 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  25. Sokol, C. L. & Luster, A. D. The chemokine system in innate immunity. Cold Spring Harb. Perspect. Biol. 7, a016303 (2015).

    Article  PubMed  PubMed Central  Google Scholar 

  26. Mackay, C. R. CXCR3+CCR5+ T cells and autoimmune diseases: guilty as charged? J. Clin. Invest. 124, 3682–3684 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  27. Khaw, Y. M. et al. Astrocytes lure CXCR2-expressing CD4+ T cells to gray matter via TAK1-mediated chemokine production in a mouse model of multiple sclerosis. Proc. Natl Acad. Sci. USA 118, e2017213118 (2021).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  28. Weiskopf, D. et al. Dengue virus infection elicits highly polarized CX3CR1+ cytotoxic CD4+ T cells associated with protective immunity. Proc. Natl Acad. Sci. USA 112, E4256–E4263 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  29. Stolla, M. et al. Fractalkine is expressed in early and advanced atherosclerotic lesions and supports monocyte recruitment via CX3CR1. PLoS ONE 7, e43572 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  30. Lesnik, P., Haskell, C. A. & Charo, I. F. Decreased atherosclerosis in CX3CR1−/− mice reveals a role for fractalkine in atherogenesis. J. Clin. Invest. 111, 333–340 (2003).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  31. Abi-Younes, S. et al. The stromal cell-derived factor-1 chemokine is a potent platelet agonist highly expressed in atherosclerotic plaques. Circ. Res. 86, 131–138 (2000).

    Article  CAS  PubMed  Google Scholar 

  32. Shevach, E. M. Foxp3+ T regulatory cells: still many unanswered questions–a perspective after 20 years of study. Front. Immunol. 9, 1048 (2018).

    Article  PubMed  PubMed Central  Google Scholar 

  33. Klein, L., Robey, E. A. & Hsieh, C.-S. Central CD4+ T cell tolerance: deletion versus regulatory T cell differentiation. Nat. Rev. Immunol. 19, 7–18 (2019).

    Article  CAS  PubMed  Google Scholar 

  34. Cording, S. et al. The intestinal micro-environment imprints stromal cells to promote efficient Treg induction in gut-draining lymph nodes. Mucosal Immunol. 7, 359–368 (2014).

    Article  CAS  PubMed  Google Scholar 

  35. Pezoldt, J. et al. Neonatally imprinted stromal cell subsets induce tolerogenic dendritic cells in mesenteric lymph nodes. Nat. Commun. 9, 3903 (2018).

    Article  PubMed  PubMed Central  Google Scholar 

  36. Bettelli, E. et al. Reciprocal developmental pathways for the generation of pathogenic effector TH17 and regulatory T cells. Nature 441, 235–238 (2006).

    Article  CAS  PubMed  Google Scholar 

  37. Korn, T. et al. IL-6 controls Th17 immunity in vivo by inhibiting the conversion of conventional T cells into Foxp3+ regulatory T cells. Proc. Natl Acad. Sci. USA 105, 18460–18465 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  38. 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  PubMed  PubMed Central  Google Scholar 

  39. Josefowicz, S. Z. et al. Extrathymically generated regulatory T cells control mucosal TH2 inflammation. Nature 482, 395–399 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  40. Chaudhry, A. et al. CD4+ regulatory T cells control TH17 responses in a Stat3-dependent manner. Science 326, 986–991 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  41. Zheng, Y. et al. Regulatory T-cell suppressor program co-opts transcription factor IRF4 to control TH2 responses. Nature 458, 351–356 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  42. Koch, M. A. et al. T-bet+ Treg cells undergo abortive Th1 cell differentiation due to impaired expression of IL-12 receptor β2. Immunity 37, 501–510 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  43. Wang, Z. et al. Pairing of single-cell RNA analysis and T cell antigen receptor profiling indicates breakdown of T cell tolerance checkpoints in atherosclerosis. Nat. Cardiovasc. Res. 2, 290–306 (2023).

    Article  PubMed  PubMed Central  Google Scholar 

  44. Depuydt, M. A. C. et al. Single-cell T cell receptor sequencing of paired human atherosclerotic plaques and blood reveals autoimmune-like features of expanded effector T cells. Nat. Cardiovasc. Res. 2, 112–125 (2023).

    Article  Google Scholar 

  45. Fisson, S. et al. Continuous activation of autoreactive CD4+ CD25+ regulatory T cells in the steady state. J. Exp. Med. 198, 737–746 (2003).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  46. Chowdhury, R. R. et al. Human coronary plaque T cells are clonal and cross-react to virus and self. Circ. Res. 130, 1510–1530 (2022).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  47. Thome, J. J. C. & Farber, D. L. Emerging concepts in tissue-resident T cells: lessons from humans. Trends Immunol. 36, 428–435 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  48. Van Acker, H. H., Capsomidis, A., Smits, E. L. & Van Tendeloo, V. F. CD56 in the immune system: more than a marker for cytotoxicity? Front. Immunol. 8, 892 (2017).

    Article  PubMed  PubMed Central  Google Scholar 

  49. Sharma, M. D. et al. An inherently bifunctional subset of Foxp3+ T helper cells is controlled by the transcription factor Eos. Immunity 38, 998–1012 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  50. Sharma, M. D. et al. Reprogrammed foxp3+ regulatory T cells provide essential help to support cross-presentation and CD8+ T cell priming in naive mice. Immunity 33, 942–954 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  51. Rosales, S. L. et al. A sensitive and integrated approach to profile messenger RNA from samples with low cell numbers. Methods Mol. Biol. 1799, 275–302 (2018).

    Article  CAS  PubMed  Google Scholar 

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

    Article  CAS  PubMed  Google Scholar 

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

    Article  PubMed  PubMed Central  Google Scholar 

  54. Pedregosa, F. et al. Scikit-learn: machine learning in Python. J. Mach. Learn. Res. 12, 2825–2830 (2011).

    Google Scholar 

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

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  56. Huang, H., Wang, C., Rubelt, F., Scriba, T. J. & Davis, M. M. Analyzing the Mycobacterium tuberculosis immune response by T cell receptor clustering with GLIPH2 and genome-wide antigen screening. Nat. Biotechnol. 38, 1194–1202 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

Download references

Acknowledgements

We thank A. Rudensky at Memorial Sloan Kettering Cancer Center for providing lineage-tracker mice. We also thank members of the clinical core and flow cytometry core at LJI. We thank Z. Mikulski and S. McArdle, microscopy core, LJI, for capturing images of exTreg cells and Treg cells in mouse arteries. We thank H. Cheroutre and N. Thiault who kindly provided the P815 cell line.

Author information

Authors and Affiliations

Authors

Contributions

K.L., A.F., P.R. and S.S.A. designed the experimental and analytical workflow. A.F. and P.R. conducted most experiments, analyzed and assembled data and drafted parts of the paper. S.S.A., M.O., A.K., J.M., Y.G. and R.G. conducted bioinformatic data analysis. S.K. conducted flow cytometry. M. Orecchioni conducted RT–qPCR and mouse proliferation experiments. A.J.A. and H.W. generated the mouse strain by breeding. A.J.A. generated most of the mouse data. Q.L. assisted in data analysis. E.W. prepared libraries for human bulk RNA-seq experiments. C.D. prepared human single-cell and mouse bulk RNA-seq libraries. F.N. assisted in mouse strain maintenance and provided expertise in proliferation experiments. K.L. conceived and supervised the study, provided funding and wrote the paper. We acknowledge the support from funding agencies. Our work is supported by the National Institutes of Health awards P01 HL136275 and R35 HL145241 to K.L., the American Heart Association’s Career Development Award (941152) to M. Orecchioni, a fellowship from Neven-DuMont Foundation to H.W. and a Deutsche Forschungsgemeinschaft fellowship (NE 2574/1-1) to F.N.

Corresponding author

Correspondence to Klaus Ley.

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: L. A. Dempsey, in collaboration with the Nature Immunology team. Peer reviewer reports are available.

Additional information

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

Extended data

Extended Data Fig. 1 Experimental controls for lineage-tracker atherosclerotic mouse model and differentially expressed genes in mouse Treg cells vs exTregs.

(a,b) Eight week-old female Foxp3eGFP-Cre-ERT2ROSA26fl-STOP-fl-tdTomato Apoe−/− mice were injected with Tamoxifen twice for 5 days each, at week 1 and 6, then fed Western diet (WD) for 12 weeks. a) gating strategy, b) representative plots and quantification of exTreg and Treg cells among CD4+T cells (black circles) in lymph nodes (LNs) and spleen, harvested after 12 weeks of WD from 3 independent mice. Non-T cells (orange open squares) and CD4T cells (green open circles) are negative controls. Frequencies of exTregs and Treg cells among the parent subsets were plotted as mean ± SEM. Statistical comparisons were done using 2-way ANOVA with Dunnett’s multiple comparison test. **** p < 0.0001. c) Volcano plot representing significantly differentially expressed genes between mouse Treg cells and exTregs from 20-week old FoxP3eGFP-Cre-ERT2ROSA26CAG-fl-stop-fl-tdTomato Apoe−/− mice (lymph nodes and spleen pooled). Left, up in Treg cells (blue). Right, up in exTregs (red). Horizontal dotted line is at -log10 (p adjusted) = 1.3 (padj = 0.05). The top 60 exTreg and 60 Treg classifying genes from the SVM model are annotated. Canonical Treg genes Il2ra and Foxp3 are shown in black boxes. Statistical analyses of DE genes using two-tailed Wald test with Benjamini-Hochberg correction for multiple comparisons.

Source data

Extended Data Fig. 2 Expression of exTreg candidate genes in scRNAseq data and validation by qRT-PCR from sorted human cells.

(a) Feature maps showing the gene expression of the single gene markers CST7, NKG7, GZMA, PRF1, TBX21 and CCL4 in the human single-cell dataset for all CD4 T cell clusters. (b) Combinations 1-4 and 6 of exTreg candidate genes are highlighted in red on UMAP embeddings of CD4 T clusters from the scRNA-Seq. (c) UMAP embeddings of CD4 T clusters. Black outline marks cluster CD4T_7; cells that express either CD56 (left) or CD16 (right) are shown as red dots. (d) Gating strategy to identify exTreg and Treg cells in human PBMCs. Dump channel: CD14, CD19. (e) Gene expression analysis of CST7, NKG7, GZMA, PRF1, TBX21 and CCL4 in sorted human Treg cells (blue circles) and exTregs (red circles) by qRT-PCR. Gene-specific Ct values were normalized (ΔCt) based on actin (ACTB). Relative expression was calculated by the 1/ΔCt method. n = 7. 33.33% male, 66.67% female donors; age: 21-54 yrs. Data shown as mean ± SEM. Each dot represents a biological replicate from an independent donor. Statistical comparisons by two-tailed Mann Whitney U test. **p = 0.0012,***p = 0.0006.

Source data

Extended Data Fig. 3 Human bulk RNAseq.

(a) Gating strategy used to sort human exTregs and Treg cells to perform bulk RNA-seq. (b) gene set enrichment analysis (GSEA) of bulk RNA-seq transcriptomes of sorted human exTreg cells against CD4T_7 (left) and all other clusters (right). Normalized enrichment score (NES) and FDR q values are indicated. (c) Significantly (adjusted p < 0.05) enriched pathways in human exTreg cells, based on genes expressed at significantly higher levels in human exTreg than in Treg cells. Analysis by Bioplanet2019 from the EnrichR suite. Dotted line indicates adjusted p = 0.05 (-log10 padj=1.3). Statistical comparisons with two-tailed Fisher’s exact test and Benjamini- Hochberg adjustment of p-values. (d) Gating strategy to identify exTregs and NK cells in human PBMCs. Dump channel: CD14, CD19.

Source data

Extended Data Fig. 4 Mouse bulk RNAseq.

(a) Comparative gene signature analysis between mouse exTregs and Treg cells. Genes were filtered for significant differential expression in mouse and human dataset. Gene expression shown here is from FoxP3eGFP-Cre-ERT2ROSA26CAG-fl-stop-fl-tdTomato Apoe−/− mice. Low-expressed genes (<7 raw reads in all samples) in our dataset were filtered out. Technical replicates were averaged, biological replicates shown as columns. Analysis of differentially expressed (DE) genes was done using DESeq2. Curated list of significant DE (log2FC ± 1, adjusted p < 0.05) genes are shown on normalized heatmaps, scaled by row (z scores). (b) Gene set enrichment analysis (GSEA) of mouse exTreg genes from bulk RNA-seq transcriptomes against human exTreg (left) and Treg cells (right) from the human bulk RNA-seq data set. Mouse orthologs of human genes, filtered for those present in the human scRNA-Seq targeted gene panel, were used to calculate enrichment for mouse bulk RNA-seq dataset. (c) Comparative gene signature analysis between mouse exTreg and NK cells. An external dataset was used for mouse NK cells (3 samples): GSE122597, GSE116177, and GSE52043. EdgeR was used to normalize the counts by applying the trimmed mean of M-values (TMM) method and counts per million (CPM) conversion. All other data processing and filtering steps were same as in a. Curated list of significant DE (log2FC ± 1, adjusted p < 0.05) genes are shown on normalized heatmaps, scaled by row (z scores). Statistical analyses of DE genes (a,c) using two-tailed Wald test with Benjamini-Hochberg correction for p-value adjustment. All data from independent biological replicates.

Source data

Extended Data Fig. 5 Assessment of proliferation in mouse Treg cells and exTregs.

(a) Eight week-old female Foxp3eGFP-Cre-ERT2ROSA26fl-STOP-fl-tdTomato Apoe−/− mice were injected with Tamoxifen twice for 5 days each, at week 1 and 6, then fed Western diet (WD) for 12 weeks. BrDU (0.8 mg/mL) was incorporated in the drinking water for the last 9 days of WD (n = 6). (b) Gating scheme for CD4+T cells. (c) Ki67 FMO control. (d) Representative plots and quantification of proliferating Treg cells (blue circles, %Ki67+BrDU+CD4+Foxp3+RFP+) and exTregs (red circles, %Ki67+BrDU+CD4+Foxp3RFP+) in the spleen (n = 6), as identified by anti-BrDU and anti-Ki-67 Abs. Data shown as mean ± SEM. Each animal is an independent biological replicate. Gates were set by FMO for Ki67 and by no BrdU controls for BrdU. Background from “No BrDU” control was subtracted for normalization. The percentage of proliferating cells was compared by two-tailed Mann-Whitney U test, **p = 0.0087.

Source data

Extended Data Fig. 6 Treg marker expression on human exTregs.

Representative contour FACS plots showing the expressions of PD-1, GITR, LAG3 and TIGIT in exTreg cells (left). Corresponding FMO controls were used to set the gates. Right, contour plots showing the expression of these markers in all CD4+T cells and in Treg cells from the same donor.

Extended Data Fig. 7 Cytotoxic and T cell activation marker expression on stimulated human exTreg vs NK cells.

Contour plots show intracellular expression of CD40L (X-axis) and Perforin (Y-axis) in exTregs and NK cells from unstimulated and PMA+ionomycin stimulated PBMCs. Data from three independent donors. 33.33% male, 66.67% female donors, age: 25-43 yrs.

Extended Data Fig. 8 Gating strategy and representative plots.

(a) Gating strategy used to analyze granzyme B, perforin and TNF in Treg cells and exTregs. (b) Contour plots show surface expression of chemokine receptors CCR5, CXCR2, CXCR3, CXCR4 and CX3CR1 on exTreg cells (left) and their corresponding expression in all CD4+T cells (right). Individual FMO controls were used to set the gate for expression of each receptor.

Supplementary information

Reporting Summary

Peer Review File

Supplementary Tables 1–14

Supplementary Table 1: Mouse and human ortholog genes. Supplementary Table 2: SVM classifier: Top 60 genes classifying Treg cells and exTreg cells. Supplementary Table 3: DEGs of mouse exTreg cells versus Treg cells. Supplementary Table 4: Number and frequency of CD4+ T cells expressing each candidate gene combination. Supplementary Table 5: Genes driving PC1 in human bulk RNA-seq data. Supplementary Table 6: DEGs of human exTreg cells versus Treg cells. Supplementary Table 7: Results obtained from the TCRβ-seq of human exTreg cells, Treg cells and Tn (naïve) cells. Supplementary Table 8: 345 significantly enriched and expanded GLIPH patterns with human exTreg TCRs. Supplementary Table 9: Differential frequency analysis of exTreg cell and Treg TCRs in the 178 shared GLIPH groups. Supplementary Table 10: Human Treg DEGs (significantly up in Treg cells compared to helper T cells) that are still expressed in exTreg cells. Supplementary Table 11: Human exTreg DEGs present in both bulk RNA-seq and scRNA-seq datasets. Supplementary Table 12: Human (h) and mouse (m) antibodies used in the study. Supplementary Table 13: Details of donors and PBMCs used for the Bulk RNA-seq and TCRβ-seq experiments. Supplementary Table 14: Details of CDR3β sequences identified in immunoSEQ assay (Adaptive).

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.

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

Freuchet, A., Roy, P., Armstrong, S.S. et al. Identification of human exTreg cells as CD16+CD56+ cytotoxic CD4+ T cells. Nat Immunol 24, 1748–1761 (2023). https://doi.org/10.1038/s41590-023-01589-9

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41590-023-01589-9

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