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:

Intrinsically disordered domain of transcription factor TCF-1 is required for T cell developmental fidelity

Abstract

In development, pioneer transcription factors access silent chromatin to reveal lineage-specific gene programs. The structured DNA-binding domains of pioneer factors have been well characterized, but whether and how intrinsically disordered regions affect chromatin and control cell fate is unclear. Here, we report that deletion of an intrinsically disordered region of the pioneer factor TCF-1 (termed L1) leads to an early developmental block in T cells. The few T cells that develop from progenitors expressing TCF-1 lacking L1 exhibit lineage infidelity distinct from the lineage diversion of TCF-1-deficient cells. Mechanistically, L1 is required for activation of T cell genes and repression of GATA2-driven genes, normally reserved to the mast cell and dendritic cell lineages. Underlying this lineage diversion, L1 mediates binding of TCF-1 to its earliest target genes, which are subject to repression as T cells develop. These data suggest that the intrinsically disordered N terminus of TCF-1 maintains T cell lineage fidelity.

This is a preview of subscription content, access via your institution

Access options

Buy this article

Prices may be subject to local taxes which are calculated during checkout

Fig. 1: The N terminus of TCF-1 is intrinsically disordered.
Fig. 2: Loss of TCF-1’s L1 domain limits DN1-to-DN2 transition.
Fig. 3: GATA2-driven mast cell gene signature is identified in developing T cells lacking L1.
Fig. 4: L1 modulates binding and transcriptional outcomes in early T cell development.
Fig. 5: L1 can be functionally substituted with another unstructured domain.
Fig. 6: Loss of L1 has limited effect on chromatin accessibility in committed T cells.

Similar content being viewed by others

Data availability

ATAC-seq, RNA-seq and CUT&RUN date have been deposited in the NCBI Gene Expression Omnibus (GEO) under accession code GSE213238. The MS proteomics data have been deposited to the ProteomeXchange Consortium via the PRIDE partner repository under dataset identifier PXD043586. Proteome raw files were searched with the SwissProt mouse database (updated January 2023; https://www.uniprot.org/help/downloads). Other publicly available datasets for microarray experiments were accessed in the GEO under accession code GSE82044. Source data are provided with this paper.

Code availability

Code is available upon reasonable request.

References

  1. Cirillo, L. A. et al. Opening of compacted chromatin by early developmental transcription factors HNF3 (FoxA) and GATA-4. Mol. Cell 9, 279–289 (2002).

    CAS  PubMed  Google Scholar 

  2. Zaret, K. S. & Carroll, J. S. Pioneer transcription factors: establishing competence for gene expression. Genes Dev. 25, 2227–2241 (2011).

    CAS  PubMed  PubMed Central  Google Scholar 

  3. Lambert, S. A. et al. The human transcription factors. Cell 172, 650–665 (2018).

    CAS  PubMed  Google Scholar 

  4. Verbeek, S. et al. An HMG-box-containing T cell factor required for thymocyte differentiation. Nature 374, 70–74 (1995).

    CAS  PubMed  Google Scholar 

  5. Johnson, J. L. et al. Lineage-determining transcription factor TCF-1 initiates the epigenetic identity of T cells. Immunity 48, 243–257 (2018).

    CAS  PubMed  PubMed Central  Google Scholar 

  6. Emmanuel, A. O. et al. TCF-1 and HEB cooperate to establish the epigenetic and transcription profiles of CD4+CD8+ thymocytes. Nat. Immunol. 19, 1366–1378 (2018).

    CAS  PubMed  PubMed Central  Google Scholar 

  7. Wang, W. et al. TCF-1 promotes chromatin interactions across topologically associating domains in T cell progenitors. Nat. Immunol. 23, 1052–1062 (2022).

    CAS  PubMed  PubMed Central  Google Scholar 

  8. Xu, Z. et al. Cutting edge: beta-catenin-interacting Tcf1 isoforms are essential for thymocyte survival but dispensable for thymic maturation transitions. J. Immunol. 198, 3404–3409 (2017).

    CAS  PubMed  Google Scholar 

  9. Zhao, X., Shan, Q. & Xue, H. H. TCF1 in T cell immunity: a broadened frontier. Nat. Rev. Immunol. 22, 147–157 (2022).

    CAS  PubMed  Google Scholar 

  10. Love, J. J. et al. Structural basis for DNA bending by the architectural transcription factor LEF-1. Nature 376, 791–795 (1995).

    CAS  PubMed  Google Scholar 

  11. Xing, S. et al. Tcf1 and Lef1 transcription factors establish CD8+ T cell identity through intrinsic HDAC activity. Nat. Immunol. 17, 695–703 (2016).

    CAS  PubMed  PubMed Central  Google Scholar 

  12. Wang, C., Uversky, V. N. & Kurgan, L. Disordered nucleiome: abundance of intrinsic disorder in the DNA- and RNA-binding proteins in 1121 species from Eukaryota, Bacteria and Archaea. Proteomics 16, 1486–1498 (2016).

    CAS  PubMed  Google Scholar 

  13. Shin, Y. & Brangwynne, C. P. Liquid phase condensation in cell physiology and disease. Science https://doi.org/10.1126/science.aaf4382 (2017).

  14. Jumper, J. et al. Highly accurate protein structure prediction with AlphaFold. Nature 596, 583–589 (2021).

    CAS  PubMed  PubMed Central  Google Scholar 

  15. Boija, A. et al. Transcription factors activate genes through the phase-separation capacity of their activation domains. Cell 175, 1842–1855 (2018).

    CAS  PubMed  Google Scholar 

  16. Ben Chorin, A. et al. ConSurf-DB: an accessible repository for the evolutionary conservation patterns of the majority of PDB proteins. Protein Sci. 29, 258–267 (2020).

    PubMed  Google Scholar 

  17. Peng, K., Radivojac, P., Vucetic, S., Dunker, A. K. & Obradovic, Z. Length-dependent prediction of protein intrinsic disorder. BMC Bioinformatics 7, 208 (2006).

    PubMed  PubMed Central  Google Scholar 

  18. Skinner, J. J., Lim, W. K., Bedard, S., Black, B. E. & Englander, S. W. Protein hydrogen exchange: testing current models. Protein Sci. 21, 987–995 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  19. Bai, Y., Milne, J. S., Mayne, L. & Englander, S. W. Primary structure effects on peptide group hydrogen exchange. Proteins 17, 75–86 (1993).

    CAS  PubMed  PubMed Central  Google Scholar 

  20. Connelly, G. P., Bai, Y., Jeng, M. F. & Englander, S. W. Isotope effects in peptide group hydrogen exchange. Proteins 17, 87–92 (1993).

    CAS  PubMed  Google Scholar 

  21. Hosokawa, H. et al. Stage-specific action of Runx1 and GATA3 controls silencing of PU.1 expression in mouse pro-T cells. J. Exp. Med. https://doi.org/10.1084/jem.20202648 (2021).

  22. Hu, G. et al. Transformation of accessible chromatin and 3D nucleome underlies lineage commitment of early T cells. Immunity 48, 227–242 (2018).

    CAS  PubMed  PubMed Central  Google Scholar 

  23. Rothenberg, E. V., Hosokawa, H. & Ungerback, J. Mechanisms of action of hematopoietic transcription factor PU.1 in initiation of T cell development. Front Immunol. 10, 228 (2019).

    CAS  PubMed  PubMed Central  Google Scholar 

  24. Shin, B. et al. Runx1 and Runx3 drive progenitor to T-lineage transcriptome conversion in mouse T cell commitment via dynamic genomic site switching. Proc. Natl Acad. Sci. USA https://doi.org/10.1073/pnas.2019655118 (2021).

  25. Ungerback, J. et al. Pioneering, chromatin remodeling, and epigenetic constraint in early T cell gene regulation by SPI1 (PU.1). Genome Res. 28, 1508–1519 (2018).

    PubMed  PubMed Central  Google Scholar 

  26. Zhou, W., Gao, F., Romero-Wolf, M., Jo, S. & Rothenberg, E. V. Single-cell deletion analyses show control of pro-T cell developmental speed and pathways by Tcf7, Spi1, Gata3, Bcl11a, Erg, and Bcl11b. Sci. Immunol. 7, eabm1920 (2022).

    CAS  PubMed  PubMed Central  Google Scholar 

  27. Zhou, W. et al. Single-cell analysis reveals regulatory gene expression dynamics leading to lineage commitment in early T cell development. Cell Syst. 9, 321–337 (2019).

    CAS  PubMed  PubMed Central  Google Scholar 

  28. Schmitt, T. M. & Zuniga-Pflucker, J. C. T cell development, doing it in a dish. Immunol. Rev. 209, 95–102 (2006).

    PubMed  Google Scholar 

  29. Schmitt, T. M. & Zuniga-Pflucker, J. C. Induction of T cell development from hematopoietic progenitor cells by delta-like-1 in vitro. Immunity 17, 749–756 (2002).

    CAS  PubMed  Google Scholar 

  30. de Boer, J. et al. Transgenic mice with hematopoietic and lymphoid specific expression of Cre. Eur. J. Immunol. 33, 314–325 (2003).

    PubMed  Google Scholar 

  31. Yang, Q. et al. TCF-1 upregulation identifies early innate lymphoid progenitors in the bone marrow. Nat. Immunol. 16, 1044–1050 (2015).

    CAS  PubMed  PubMed Central  Google Scholar 

  32. Weber, B. N. et al. A critical role for TCF-1 in T-lineage specification and differentiation. Nature 476, 63–68 (2011).

    CAS  PubMed  PubMed Central  Google Scholar 

  33. Yoshida, H. et al. The cis-regulatory atlas of the mouse immune system. Cell 176, 897–912 (2019).

    CAS  PubMed  PubMed Central  Google Scholar 

  34. Ling, K. W. et al. GATA-2 plays two functionally distinct roles during the ontogeny of hematopoietic stem cells. J. Exp. Med. 200, 871–882 (2004).

    CAS  PubMed  PubMed Central  Google Scholar 

  35. Li, Y. et al. GATA2 regulates mast cell identity and responsiveness to antigenic stimulation by promoting chromatin remodeling at super-enhancers. Nat. Commun. 12, 494 (2021).

    CAS  PubMed  PubMed Central  Google Scholar 

  36. Onodera, K. et al. GATA2 regulates dendritic cell differentiation. Blood 128, 508–518 (2016).

    CAS  PubMed  PubMed Central  Google Scholar 

  37. Tsai, F. Y. & Orkin, S. H. Transcription factor GATA-2 is required for proliferation/survival of early hematopoietic cells and mast cell formation, but not for erythroid and myeloid terminal differentiation. Blood 89, 3636–3643 (1997).

    CAS  PubMed  Google Scholar 

  38. Jin, W. et al. Critical POU domain residues confer Oct4 uniqueness in somatic cell reprogramming. Sci. Rep. 6, 20818 (2016).

    CAS  PubMed  PubMed Central  Google Scholar 

  39. Boller, S. et al. Pioneering activity of the C-terminal domain of EBF1 shapes the chromatin landscape for B cell programming. Immunity 44, 527–541 (2016).

    CAS  PubMed  Google Scholar 

  40. Wang, Y. et al. A prion-like domain in transcription factor EBF1 promotes phase separation and enables B cell programming of progenitor chromatin. Immunity 53, 1151–1167 (2020).

    CAS  PubMed  Google Scholar 

  41. de Bruijn, M. F. & Speck, N. A. Core-binding factors in hematopoiesis and immune function. Oncogene 23, 4238–4248 (2004).

    PubMed  Google Scholar 

  42. Xing, S. et al. Tle corepressors are differentially partitioned to instruct CD8+ T cell lineage choice and identity. J. Exp. Med. 215, 2211–2226 (2018).

    CAS  PubMed  PubMed Central  Google Scholar 

  43. Hosokawa, H. et al. Transcription factor PU.1 represses and activates gene expression in early T cells by redirecting partner transcription factor binding. Immunity 48, 1119–1134 (2018).

    CAS  PubMed  PubMed Central  Google Scholar 

  44. Frederick, M. A. et al. A pioneer factor locally opens compacted chromatin to enable targeted ATP-dependent nucleosome remodeling. Nat. Struct. Mol. Biol. 30, 31–37 (2023).

    CAS  PubMed  Google Scholar 

  45. Minderjahn, J. et al. Mechanisms governing the pioneering and redistribution capabilities of the non-classical pioneer PU.1. Nat. Commun. 11, 402 (2020).

    CAS  PubMed  PubMed Central  Google Scholar 

  46. Xue, B., Dunbrack, R. L., Williams, R. W., Dunker, A. K. & Uversky, V. N. PONDR-FIT: a meta-predictor of intrinsically disordered amino acids. Biochim. Biophys. Acta 1804, 996–1010 (2010).

    CAS  PubMed  PubMed Central  Google Scholar 

  47. Nguyen, D., Mayne, L., Phillips, M. C. & Walter Englander, S. Reference parameters for protein hydrogen exchange rates. J. Am. Soc. Mass. Spectrom. 29, 1936–1939 (2018).

    CAS  PubMed  PubMed Central  Google Scholar 

  48. Carleton, M. et al. Signals transduced by CD3ε, but not by surface pre-TCR complexes, are able to induce maturation of an early thymic lymphoma in vitro. J. Immunol. 163, 2576–2585 (1999).

    CAS  PubMed  Google Scholar 

  49. Ogilvy, S. et al. Promoter elements of vav drive transgene expression in vivo throughout the hematopoietic compartment. Blood 94, 1855–1863 (1999).

    CAS  PubMed  Google Scholar 

  50. Shimshek, D. R. et al. Codon-improved Cre recombinase (iCre) expression in the mouse. Genesis 32, 19–26 (2002).

    CAS  PubMed  Google Scholar 

  51. Dou, Z. et al. Autophagy mediates degradation of nuclear lamina. Nature 527, 105–109 (2015).

    CAS  PubMed  PubMed Central  Google Scholar 

  52. Schindelin, J. et al. Fiji: an open-source platform for biological-image analysis. Nat. Methods 9, 676–682 (2012).

    CAS  PubMed  Google Scholar 

  53. Aguilan, J. T., Kulej, K. & Sidoli, S. Guide for protein fold change and P value calculation for non-experts in proteomics. Mol. Omics 16, 573–582 (2020).

    CAS  PubMed  Google Scholar 

  54. Buenrostro, J. D., Giresi, P. G., Zaba, L. C., Chang, H. Y. & Greenleaf, W. J. Transposition of native chromatin for fast and sensitive epigenomic profiling of open chromatin, DNA-binding proteins and nucleosome position. Nat. Methods 10, 1213–1218 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  55. Fasolino, M. et al. Genetic variation in type 1 diabetes reconfigures the 3D chromatin organization of T cells and alters gene expression. Immunity 52, 257–274 (2020).

    CAS  PubMed  PubMed Central  Google Scholar 

  56. Stirling, D. R. et al. CellProfiler 4: improvements in speed, utility and usability. BMC Bioinformatics 22, 433 (2021).

    PubMed  PubMed Central  Google Scholar 

  57. Mayne, L. Hydrogen exchange mass spectrometry. Methods Enzymol. 566, 335–356 (2016).

    CAS  PubMed  Google Scholar 

  58. Mayne, L. et al. Many overlapping peptides for protein hydrogen exchange experiments by the fragment separation-mass spectrometry method. J. Am. Soc. Mass. Spectrom. 22, 1898–1905 (2011).

    CAS  PubMed  PubMed Central  Google Scholar 

  59. Kan, Z. Y., Ye, X., Skinner, J. J., Mayne, L. & Englander, S. W. ExMS2: an integrated solution for hydrogen–deuterium exchange mass spectrometry data analysis. Anal. Chem. 91, 7474–7481 (2019).

    CAS  PubMed  Google Scholar 

  60. Glasoe, P. K. & Long, F. A. Use of glass electrodes to measure acidities in deuterium oxide. J. Phys. Chem. 64, 188–190 (1960).

    CAS  Google Scholar 

Download references

Acknowledgements

We thank helpful discussions with members of the laboratory of G.V. in addition to M. Fasolino, K. Zaret and N. Speck. We thank K. Gupta and the Johnson Foundation Structural Biology and Biophysics Core at the Perelman School of Medicine (RRID: SCR_022414) for HX–MS resources. Flow cytometry data were generated in the Penn Cytomics and Cell Sorting Shared Resource Laboratory at the University of Pennsylvania (RRID: SCR_022376), which is partially supported by P30CA016520 (Abramson Cancer Center). We thank laboratories at the University of Pennsylvania including the Marmorstein laboratory for protein purification equipment, the Berger laboratory for 6xHis-TEV bacterial expression vector, and the Lynch laboratory for Typhoon imager access. The work in this manuscript was supported by F30AI161873 (to A.V.), R01AI091627 (to I.M.), F30AI174776 (to M.A.S.), R01-CA230800 and R01-CA248041 (to R.B.F.), the Burroughs Welcome Fund, the Chan Zuckerberg Initiative Award, W. W. Smith Charitable Trust, Sloan Foundation and the NIH grants R01AI168136, R01AI168240, UC4DK112217, U01DK112217, R01HL145754, U01DK127768 and U01DA052715 (to G.V.).

Author information

Authors and Affiliations

Authors

Contributions

Concepts and experimental design were provided by N.G. and G.V. with help from A.V., I.M. and M.A.S. Experiments were conducted by N.G., I.J., A.C., M.A.S., J.A. and S.S. Analysis of genomics data was done by N.G. with help from A.R.P. Imaging and analysis were done by A.J., N.G. and E.K.F. HX experiments and analysis were done by M.A.S. and N.G. with support from L.M. Experimental support and discussion were provided by Y.Z., R.B.F., E.J.W. and H.-H.X. The paper was written by N.G. and G.V.

Corresponding author

Correspondence to Golnaz Vahedi.

Ethics declarations

Competing interests

E.J.W. is a member of the Parker Institute for Cancer Immunotherapy that supports research in the laboratory of E.J.W. E.J.W. is an advisor for Danger Bio, Marengo, Janssen, NewLimit, Pluto Immunotherapeutics Related Sciences, Santa Ana Bio, Synthekine and Surface Oncology. E.J.W. is a founder of and holds stock in Surface Oncology, Danger Bio and Arsenal Biosciences. The other authors declare no competing interests.

Peer review

Peer review information

Nature Immunology thanks David Weiss and the other, anonymous, reviewer(s) for their contribution to the peer review of this work. Primary Handling Editor: N. Bernard, 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 The N terminus of TCF-1 is intrinsically disordered.

a) Size exclusion chromatography purification of affinity- and ion exchange-purified recombinant TCF-1 protein expressed in E. coli. Chromatogram (left) displays measured A280 vs. elution volume for a representative injection of pooled ion exchange fractions. Vertical dashed lines indicate approximate position of fractions pooled for analysis by HX-MS. Gel image displays SDS-PAGE analysis (stained with SYPRO Orange) of final protein input to HX-MS. Chromatogram (right) displays A280 vs. elution time for repeated analysis of indicated fractions from a prior Superose 6 Increase 10/300GL run during the initial purification. b) Plots of normalized deuterium uptake (relative to measured deuterium content after 23hrs of H-to-D exchange) at each indicated measured sample pH (pHmeas) for each TCF-1 peptide observation at the indicated exchange times. For observations with technical replicates (n = 3 independent samples for pHmeas 6.0 4sec, 10sec), center line represents mean value with error bars corresponding to standard deviation. Shaded columns indicated pHmeas and time conditions where approximately equivalent exchange is expected given the pH dependency of HX rates. c) Representative mass spectra of indicated peptide observations (generated using ExMS259). Relative to the all-1H sample (treated as HX time = 0sec), the change in m/z of each centroid distribution reflects the indicated change in mass due to deuterium incorporation. d) Time-scaled, back exchange-corrected deuterium content vs. time experimental data as in Fig. 1c was fit to a stretched-exponential function for each peptide across the indicated regions of TCF-1. Boxplots of estimated kex values are approximate estimates of observed peptide-level HX rate constants. Center line of box plots is median, limits are 1st and 3rd quartiles, and whiskers are maximum and minimum values. e) Comparison of approximate predicted (random coil) kpred vs. observed kex peptide-level HX rate constants across peptides from the L1-L7 regions of TCF-1. Values for kex are as in (d). Values for kpred were estimated from the stretched-exponential fitting approach using predicted deuterium content vs. time data calculated from the predicted residue-specific HX rate constants across each respective peptide sequence under the assumption of no protection from exchange. Pearson correlation coefficient and corresponding correlation P value are displayed.

Extended Data Fig. 2 Loss of TCF-1′s L1 domain limits DN1 to DN2 transition.

a) Thy1+ CD25+ cells (b) DN1 (CD44+ CD25), DN2 (CD44+CD25+), and DN3 (CD44- CD25+) in OP9-DLL1 co-cultures of Tcf7 cKOs transduced with EV, WT, or mutant TCF-1 (ΔL1-7) on day 13. c) Quantification of Thy1+ CD25+ cells (top) and DN2s, and DN3s by CD44 and CD25 surface expression (bottom) in OP9-DLL1 co-cultures of Tcf7 cKOs transduced with EV, WT, or mutant TCF-1 on day 13. d) DN1 (CD44 CD25), DN2 (CD44+ CD25+), and DN3 (CD44 CD25+) cells in co-cultures of wild type (WT) ckit+ bone marrow (BM) progenitors transduced with WT TCF-1, EV, or mutant TCF-1 (ΔL1-L7) (GFP+) on OP9-DLL1 cells at day 5. e) Frequency of GFP+ (transduced) and GFP (un-transduced) DN2s and DN3s in (D) (top). Analysis of ratio of GFP+ to GFP- Thy1+ CD25+ cells in (d) (middle). Frequency of GFP+ and GFP Thy1+ CD25+ cells in (d) (bottom). f) DN1 (CD44 CD25), DN2 (CD44+ CD25+) and DN3 (CD44 CD25+) cells in WT ckit+ BM progenitors transduced with WT TCF-1, ΔL1, or EV (GFP+) on OP9-DLL4 after 5 days. g) DN1 (CD44+ CD25), DN2 (CD44+ CD25+), and DN3 (CD44 CD25+) cells in WT ckit+ BM progenitors transduced with WT TCF-1 (GFP+) on OP9-DLL1 cells (left) and OP9 cells (right) for 5 days. h) Histogram depicting TCF-1 intracellular flow cytometry in Tcf7 cKO progenitors un-transduced (Vex) or transduced with WT TCF-1, ΔL1, or EV (Vex+) as well as WT TCF-1 sufficient progenitors. i) Frequency of B220+ Vex+ (transduced) cells in Tcf7 cKO OP9-DLL1 at day 5 (left) and 13 (right). j) Number (left) and frequency (right) of CD11b+ CD25 cells at day 7 in Tcf7 cKO OP9-DLL1 co-cultures. All cells (a-j) are pre-gated on SSC-A/FSC-A, Singlets, Live cell (Viability-), CD45+ and for a,b,c,i,j Vex+ and all data are representative of at least 3 independent experiments. In c, e, i,j, bars show mean from n = 2 independent animals, dots represent individual data points. In c, i, P values are determined by one-way ANOVA followed by Dunnett’s multiple comparison test with WT TCF-1 (P45) as a control. *P < 0.05, ** P < 0.01, *** P < 0.001, and **** P < 0.001.

Extended Data Fig. 3 GATA2 driven mast cell gene signature is unmasked in developing T cells lacking L1.

a) Heatmaps depicting gene ontology enrichment in significantly differential gene sets (adjusted P<0.05, |Log2FoldChange|>1). P values are calculated using a hypergeometric test. b) Heatmap demonstrating differentially expressed genes (adjusted P <0.05) between wild type (WT) TCF-1 transduced DN1 and DN2s from Tcf7 cKO cells on OP9-DLL1 co-cultures at day 7. P-values are calculated by the Wald test and adjusted using the Benjamini and Hochberg method. c) Principle component plot of RNA-sequencing on 293T human cell line transduced with empty vector (EV), wild type (WT) human TCF-1, and an internal deletion mutant lacking the analogous L1 region of human TCF-1; human ΔL1 (upper panel). GSEA depicts the enrichment of genes in GSE22601_IMMATURE_CD4_SINGLE_POSITIVE VS_DOUBLE_POSITIVE_THYMOCYTE_UP gene set within genes upregulated in 293T cells with human TCF-1 vs. EV. d) Heatmap depicting transcription factors differentially upregulated in ΔL1 and WT TCF-1 transduced DN2s from Tcf7 cKO cells on OP9-DLL1 co-cultures at day 7 (adjusted P<0.05 and |Log2FoldChange|>1). P-values are calculated by the Wald test and adjusted using the Benjamini and Hochberg method. e) Bar plots depicting select gene expression (in RPKM) values in DN1 and DN2s from Tcf7 cKO cells on OP9-DLL1 co-cultures at day 7. Bars represent mean RPKM values, error bars represent Standard deviation (SD), and individual data points are represented with dots.

Extended Data Fig. 4 GATA2 driven mast cell transcriptional signature is unmasked in developing T cells lacking the L1 region of TCF-1.

a–e. Representative genome browser views of counts per million normalized strand specific RNA-seq tracks at Gata2 (a.), Gata3 (b.), Thy1 (c.), Mcpt1/2/4 (d.) and Bcl11b loci (e).

Extended Data Fig. 5 The L1 domain of TCF-1 modulates binding and transcriptional outcomes in early T cell development independent of chromatin accessibility.

a) SeqLogo depicting top enriched motifs from de novo HOMER motif analysis of differentially accessible ATAC-seq peaks in WT vs. EV, ΔL1 vs. EV, and WT vs ΔL1 transduced DN1 and DN2s with non-differential peaks as background. P values are calculated using a hypergeometric test. b) Venn-diagram representing TCF-1 CUT&RUN experiments and associated unique and overlapping WT TCF-1 and ΔL1 binding events in DN1 and DN2s. c) Principal component analysis of TCF-1 CUT&RUN and chromatin accessibility measurements in DN1 and DN2s. Counts of ATAC-seq and TCF-1 binding in CUT&RUN measurements were generated across the union of all peaks across all ATAC-seq and CUT&RUN conditions. d) SeqLogo depicting top enriched motifs from de novo HOMER motif analysis of L1 dependent and independent binding events in DN1 and DN2s compared to randomly generated background. P values are calculated using a hypergeometric test. e) Heatmap depicting TCF-1 binding events measured by TCF-1 CUT&RUN and chromatin accessibility in DN1 and DN2s at differentially accessible peaks open in WT DN2 vs. DN1s.

Extended Data Fig. 6 Loss of the L1 domain of TCF-1 has limited effect on chromatin accessibility in committed T cells.

a) Principal component plot of RNA-seq on Tcf7−/− DN3 like Scid.adh cells transduced with empty vector (EV), wild type (WT) TCF-1, and internal deletion mutants: ΔL1, ΔL2, ΔL6, and ΔL7. b) Volcano plot demonstrating significantly differential genes comparing WT TCF-1 and EV (left), ΔL1 and WT TCF-1 (middle), and ΔL7 and WT TCF-1 (right) transduced Tcf7/ DN3 cells. (adjusted P<0.05 and |Log2FoldChange|>1) P-values are calculated by the Wald test and adjusted using the Benjamini and Hochberg method. c) Heatmap depicting significantly up and down-regulated genes comparing WT TCF-1 and EV transduced Tcf7−/− DN3 cells. (adjusted P <0.05 and |Log2FC|>1). P-values are calculated by the Wald test and adjusted using the Benjamini and Hochberg method. d) Pathway enrichment analysis of differential gene sets depicted in B. P values are calculated using a hypergeometric test. e) Quantification of number of WT TCF-1 and ΔL1 binding events profiled by TCF-1 and FLAG CUT&RUN in Tcf7−/− KO DN3 cells. Bars represent mean number of binding sites from n = 2 biologically independent samples. f) Principal component plot of WT TCF-1 and ΔL1 binding events as measured in E.

Extended Data Fig. 7 Proteomics measurements suggest the interaction between RUNX1 and TCF-1 is dependent on the L1 domain.

a) Representative immunofluorescence images depicting GFP tagged wild type (WT) TCF-1, ΔL1 mutant TCF-1 and empty vector (EV). DAPI staining of nuclei and overlay images are included (right). Boxplot of granularity of GFP signal in DN3 cells transduced with either EV, WT TCF-1 or ΔL1 fused with GFP. Granularity indicates the percentage of highest intensity elements of 8 pixels subtracted relative to the background (see Methods). Cells with a more granular pattern or punctate localization are indicated by a lower percentage. Center line of box plots represent median granularity, limits represent 1st and 3rd quartiles, whiskers represent maximum and minimum values, data points represent outliers. Cells analyzed per condition EV: n = 189, WT TCF-1: n = 237, ΔL1: n = 190. P values were determined by a two-tailed Mann-Whitney test: *P < 0.05, ** P < 0.01, *** P < 0.001, and **** P < 0.001. Scale bar: 4µm. b) Heatmap indicating the Z score of the log2 normalized abundance of top 100 proteins detected with a higher enrichment between DN3 cells expressing WT TCF-1 and both EV and ΔL1 in mass spectrometry of a TCF-1 immunoprecipitation in DN3 cells. c) Depiction of L1 dependent TCF-1 protein-protein interaction network identified by mass spectrometry of a TCF-1 immunoprecipitation in DN3 cells. Node size and color indicate fold change in log normalized abundance between DN3 cells expressing WT TCF-1 and EV. d) Network terms corresponding to Uniprot keywords are highlighted in the network depicted in c.

Supplementary information

Source data

Source Data Fig. 1

Unprocessed immunoblots.

Source Data Fig. 2

Flow cytometry gating strategy.

Source Data Fig. 5

Unprocessed immunoblots.

Source Data Fig. 6

Unprocessed immunoblots.

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

Goldman, N., Chandra, A., Johnson, I. et al. Intrinsically disordered domain of transcription factor TCF-1 is required for T cell developmental fidelity. Nat Immunol 24, 1698–1710 (2023). https://doi.org/10.1038/s41590-023-01599-7

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41590-023-01599-7

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