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:

LAG3 ectodomain structure reveals functional interfaces for ligand and antibody recognition

Abstract

The immune checkpoint receptor lymphocyte activation gene 3 protein (LAG3) inhibits T cell function upon binding to major histocompatibility complex class II (MHC class II) or fibrinogen-like protein 1 (FGL1). Despite the emergence of LAG3 as a target for next-generation immunotherapies, we have little information describing the molecular structure of the LAG3 protein or how it engages cellular ligands. Here we determined the structures of human and murine LAG3 ectodomains, revealing a dimeric assembly mediated by Ig domain 2. Epitope mapping indicates that a potent LAG3 antagonist antibody blocks interactions with MHC class II and FGL1 by binding to a flexible ‘loop 2’ region in LAG3 domain 1. We also defined the LAG3–FGL1 interface by mapping mutations onto structures of LAG3 and FGL1 and established that FGL1 cross-linking induces the formation of higher-order LAG3 oligomers. These insights can guide LAG3-based drug development and implicate ligand-mediated LAG3 clustering as a mechanism for disrupting T cell activation.

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: Structures of human and murine LAG3 ECDs.
Fig. 2: Structural and biochemical characterization of human and murine LAG3 dimers.
Fig. 3: Epitope mapping and functional characterization of LAG3 antagonist scFvs.
Fig. 4: Identification of an FGL1-binding surface on the LAG3.
Fig. 5: Structure and LAG3-interacting residues of FGL1FD.
Fig. 6: FGL1-induced LAG3 clustering correlates with FGL1 suppression mechanism.

Similar content being viewed by others

Data availability

Crystallography data have been deposited in the Protein Data Bank under the accession no. 7TZG, 7TZH, 7TZE and 7TZ2 for the structures of hLAG3*-F7, hLAG3D34-F7, mLAG3D12 and FGL1FD, respectively.

References

  1. Robert, C. A decade of immune-checkpoint inhibitors in cancer therapy. Nat. Commun. 11, 3801 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  2. Topalian, S. L., Taube, J. M., Anders, R. A. & Pardoll, D. M. Mechanism-driven biomarkers to guide immune checkpoint blockade in cancer therapy. Nat. Rev. Cancer 16, 275–287 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  3. Okazaki, T., Chikuma, S., Iwai, Y., Fagarasan, S. & Honjo, T. A rheostat for immune responses: the unique properties of PD-1 and their advantages for clinical application. Nat. Immunol. 14, 1212–1218 (2013).

    Article  CAS  PubMed  Google Scholar 

  4. Nguyen, L. T. & Ohashi, P. S. Clinical blockade of PD1 and LAG3—potential mechanisms of action. Nat. Rev. Immunol. 15, 45–56 (2015).

    Article  CAS  PubMed  Google Scholar 

  5. Lui, Y. & Davis, S. J. LAG-3: a very singular immune checkpoint. Nat. Immunol. 19, 1278–1279 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  6. Lecocq, Q., Keyaerts, M., Devoogdt, N. & Breckpot, K. The next-generation immune checkpoint LAG-3 and its therapeutic potential in oncology: third time’s a charm. Int. J. Mol. Sci. 22, 75 (2021).

    Article  CAS  Google Scholar 

  7. Okazaki, T. et al. PD-1 and LAG-3 inhibitory co-receptors act synergistically to prevent autoimmunity in mice. J. Exp. Med. 208, 395–407 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  8. Zhang, Q. et al. LAG-3 limits regulatory T cell proliferation and function in autoimmune diabetes. Sci. Immunol. 2, eaah4569 (2017).

    Article  PubMed  PubMed Central  Google Scholar 

  9. Andrews, L. P. et al. Resistance to PD1 blockade in the absence of metalloprotease-mediated LAG3 shedding. Sci. Immunol. 5, eabc2728 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  10. Takamatsu, K. et al. Profiling the inhibitory receptors LAG-3, TIM-3, and TIGIT in renal cell carcinoma reveals malignancy. Nat. Commun. 12, 5547 (2021).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  11. Incyte Corporation. Phase 1-2 study of combination therapy with INCMGA00012 (Anti-PD-1), INCAGN02385 (Anti-LAG-3), and INCAGN02390 (Anti-TIM-3) in participants with select advanced malignancies. ClinicalTrials.gov https://clinicaltrials.gov/ct2/show/NCT04370704 (2021).

  12. Yap, T. A. et al. A first-in-human phase I study of FS118, an anti-LAG-3/PD-L1 bispecific antibody in patients with solid tumors that have progressed on prior PD-1/PD-L1 therapy. J. Clin. Oncol. 37, https://doi.org/10.1200/JCO.2019.37.15_suppl.TPS2652 (2019).

  13. Hoffmann-La Roche. An open label, multicenter, dose escalation, phase 1 study to evaluate safety/tolerability, pharmacokinetics, pharmacodynamics and preliminary anti tumor activity of RO7247669, a PD1-LAG3 bispecific antibody, in patients with advanced and/or metastatic solid tumors. ClinicalTrials.gov https://clinicaltrials.gov/ct2/show/NCT04140500 (2021).

  14. Multiple Myeloma Research Consortium. A phase I/II assessment of combination immuno-oncology drugs elotuzumab, anti-LAG-3 (BMS-986016) and anti-TIGIT (BMS-986207). ClinicalTrials.gov https://clinicaltrials.gov/ct2/show/NCT04150965 (2021).

  15. Tawbi, H. A. et al. Relatlimab and nivolumab versus nivolumab in untreated advanced melanoma. N. Engl. J. Med. 386, 24–34 (2022).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  16. Grebinoski, S. & Vignali, D. A. Inhibitory receptor agonists: the future of autoimmune disease therapeutics? Curr. Opin. Immunol. 67, 1–9 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  17. Kadowaki, A. et al. Gut environment-induced intraepithelial autoreactive CD4+ T cells suppress central nervous system autoimmunity via LAG-3. Nat. Commun. 7, 11639 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  18. Angin, M., Brignone, C. & Triebel, F. A LAG-3-specific agonist antibody for the treatment of T cell-induced autoimmune diseases. J. Immunol. 204, 810–818 (2020).

    Article  CAS  PubMed  Google Scholar 

  19. Huard, B., Prigent, P., Tournier, M., Bruniquel, D. & Triebel, F. CD4/major histocompatibility complex class II interaction analyzed with CD4 and lymphocyte activation gene-3 (LAG-3)-Ig fusion proteins. Eur. J. Immunol. 25, 2718–2721 (1995).

    Article  CAS  PubMed  Google Scholar 

  20. Miceli, M. C. & Parnes, J. R. The roles of CD4 and CD8 in T cell activation. Semin. Immunol. 3, 133–141 (1991).

    CAS  PubMed  Google Scholar 

  21. Maruhashi, T., Sugiura, D., Okazaki, I.-M. & Okazaki, T. LAG-3: from molecular functions to clinical applications. J. Immunother. Cancer 8, e001014 (2020).

    Article  PubMed  PubMed Central  Google Scholar 

  22. Hannier, S., Tournier, M., Bismuth, G. & Triebel, F. CD3/TCR complex-associated lymphocyte activation gene-3 molecules inhibit CD3/TCR signaling. J. Immunol. 161, 4058–4065 (1998).

    Article  CAS  PubMed  Google Scholar 

  23. Bhagwat, B. et al. Establishment of engineered cell-based assays mediating LAG3 and PD1 immune suppression enables potency measurement of blocking antibodies and assessment of signal transduction. J. Immunol. Methods 456, 7–14 (2018).

    Article  CAS  PubMed  Google Scholar 

  24. Workman, C. J., Dugger, K. J. & Vignali, D. A. A. Cutting edge: molecular analysis of the negative regulatory function of lymphocyte activation gene-3. J. Immunol. 169, 5392–5395 (2002).

    Article  CAS  PubMed  Google Scholar 

  25. Maeda, T. K., Sugiura, D., Okazaki, I.-M., Maruhashi, T. & Okazaki, T. Atypical motifs in the cytoplasmic region of the inhibitory immune co-receptor LAG-3 inhibit T cell activation. J. Biol. Chem. 294, 6017–6026 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  26. Huard, B., Tournier, M., Hercend, T., Triebel, F. & Faure, F. Lymphocyte-activation gene 3/major histocompatibility complex class II interaction modulates the antigenic response of CD4+ T lymphocytes. Eur. J. Immunol. 24, 3216–3221 (1994).

    Article  CAS  PubMed  Google Scholar 

  27. Weber, S. & Karjalainen, K. Mouse CD4 binds MHC class II with extremely low affinity. Int. Immunol. 5, 695–698 (1993).

    Article  CAS  PubMed  Google Scholar 

  28. Maruhashi, T. et al. LAG-3 inhibits the activation of CD4+ T cells that recognize stable pMHCII through its conformation-dependent recognition of pMHCII. Nat. Immunol. 19, 1415–1426 (2018).

    Article  CAS  PubMed  Google Scholar 

  29. Huard, B. et al. Characterization of the major histocompatibility complex class II binding site on LAG-3 protein. Proc. Natl Acad. Sci. USA 94, 5744–5749 (1997).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  30. Graydon, C. G., Mohideen, S. & Fowke, K. R. LAG3’s enigmatic mechanism of action. Front. Immunol. 11, 615317 (2021).

    Article  PubMed  PubMed Central  Google Scholar 

  31. Wang, J. et al. Fibrinogen-like protein 1 is a major immune inhibitory ligand of LAG-3. Cell 176, 334–347.e12 (2019).

    Article  CAS  PubMed  Google Scholar 

  32. Xu, F. et al. LSECtin expressed on melanoma cells promotes tumor progression by inhibiting antitumor T-cell responses. Cancer Res. 74, 3418–3428 (2014).

    Article  CAS  PubMed  Google Scholar 

  33. Kouo, T. et al. Galectin-3 shapes antitumor immune responses by suppressing CD8+ T cells via LAG-3 and inhibiting expansion of plasmacytoid dendritic cells. Cancer Immunol. Res. 3, 412–423 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  34. Mao, X. et al. Pathological α-synuclein transmission initiated by binding lymphocyte-activation gene 3. Science 353, aah3374 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  35. Zhang, S. et al. Mechanistic basis for receptor-mediated pathological α-synuclein fibril cell-to-cell transmission in Parkinson’s disease. Proc. Natl Acad. Sci. USA 118, e2011196118 (2021).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  36. Mao, X. et al. Aplp1 and the Aplp1-Lag3 complex facilitates transmission of pathologic α-synuclein. Preprint at bioRxiv https://doi.org/10.1101/2021.05.01.442157 (2021).

  37. Ascione, A. et al. Development of a novel human phage display-derived anti-LAG3 scFv antibody targeting CD8+ T lymphocyte exhaustion. BMC Biotechnol. 19, 67 (2019).

    Article  PubMed  PubMed Central  Google Scholar 

  38. Guy, H. R. Amino acid side-chain partition energies and distribution of residues in soluble proteins. Biophys. J. 47, 61–70 (1985).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  39. Ohmura, T., Ueda, T., Hashimoto, Y. & Imoto, T. Tolerance of point substitution of methionine for isoleucine in hen egg white lysozyme. Protein Eng. 14, 421–425 (2001).

    Article  CAS  PubMed  Google Scholar 

  40. Wang, J. H. et al. Crystal structure of the human CD4 N-terminal two-domain fragment complexed to a class II MHC molecule. Proc. Natl Acad. Sci. USA 98, 10799–10804 (2001).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  41. Li, N., Workman, C. J., Martin, S. M. & Vignali, D. A. A. Biochemical analysis of the regulatory T cell protein lymphocyte activation gene-3 (LAG-3; CD223). J. Immunol. 173, 6806–6812 (2004).

    Article  CAS  PubMed  Google Scholar 

  42. Su, L. F., Del Alcazar, D., Stelekati, E., Wherry, E. J. & Davis, M. M. Antigen exposure shapes the ratio between antigen-specific Tregs and conventional T cells in human peripheral blood. Proc. Natl Acad. Sci. USA 113, E6192–E6198 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  43. Mendel Kerlero de Rosbo, N. & Ben-Nun, A. Delineation of the minimal encephalitogenic epitope within the immunodominant region of myelin oligodendrocyte glycoprotein: diverse Vβ gene usage by T cells recognizing the core epitope encephalitogenic for T cell receptor Vβb and T cell receptor Vβa H-2b mice. Eur. J. Immunol. 26, 2470–2479 (1996).

    Article  CAS  PubMed  Google Scholar 

  44. Petersen, T. R. et al. Characterization of MHC- and TCR-binding residues of the myelin oligodendrocyte glycoprotein 38–51 peptide. Eur. J. Immunol. 34, 165–173 (2004).

    Article  CAS  PubMed  Google Scholar 

  45. MacLachlan, B. J. et al. Molecular characterization of HLA class II binding to the LAG-3 T cell co-inhibitory receptor. Eur. J. Immunol. 51, 331–341 (2021).

    Article  CAS  PubMed  Google Scholar 

  46. Holling, T. M., Schooten, E., Langerak, A. W. & van den Elsen, P. J. Regulation of MHC class II expression in human T-cell malignancies. Blood 103, 1438–1444 (2004).

    Article  CAS  PubMed  Google Scholar 

  47. Grandal, M. M. et al. Anti-lag-3 antibodies and compositions. WIPO (PCT) patent no. US20220056126A3 (2019).

  48. Krissinel, E. & Henrick, K. Secondary-structure matching (SSM), a new tool for fast protein structure alignment in three dimensions. Acta Crystallogr. D Biol. Crystallogr. 60, 2256–2268 (2004).

    Article  CAS  PubMed  Google Scholar 

  49. Shrive, A. K. et al. Crystal structure of the tetrameric fibrinogen-like recognition domain of fibrinogen C domain containing 1 (FIBCD1) protein. J. Biol. Chem. 289, 2880–2887 (2014).

    Article  CAS  PubMed  Google Scholar 

  50. Leppänen, V.-M., Saharinen, P. & Alitalo, K. Structural basis of Tie2 activation and Tie2/Tie1 heterodimerization. Proc. Natl Acad. Sci. USA 114, 4376–4381 (2017).

    Article  PubMed  PubMed Central  Google Scholar 

  51. Nagdas, S. K., Winfrey, V. P. & Olson, G. E. Two fibrinogen-like proteins, FGL1 and FGL2 are disulfide-linked subunits of oligomers that specifically bind nonviable spermatozoa. Int. J. Biochem. Cell Biol. 80, 163–172 (2016).

    Article  CAS  PubMed  Google Scholar 

  52. Mohan, K. et al. Topological control of cytokine receptor signaling induces differential effects in hematopoiesis. Science 364, eaav7532 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  53. Freed, D. M., Alvarado, D. & Lemmon, M. A. Ligand regulation of a constitutively dimeric EGF receptor. Nat. Commun. 6, 7380 (2015).

    Article  PubMed  Google Scholar 

  54. Moore, J. O., Lemmon, M. A. & Ferguson, K. M. Dimerization of Tie2 mediated by its membrane-proximal FNIII domains. Proc. Natl Acad. Sci. USA 114, 4382–4387 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  55. Iouzalen, N., Andreae, S., Hannier, S. & Triebel, F. LAP, a lymphocyte activation gene-3 (LAG-3)-associated protein that binds to a repeated EP motif in the intracellular region of LAG-3, may participate in the down-regulation of the CD3/TCR activation pathway. Eur. J. Immunol. 31, 2885–2891 (2001).

    Article  CAS  PubMed  Google Scholar 

  56. Liu, Z. & Ukomadu, C. Fibrinogen-like protein 1, a hepatocyte derived protein is an acute phase reactant. Biochem. Biophys. Res. Commun. 365, 729–734 (2008).

    Article  CAS  PubMed  Google Scholar 

  57. Kabsch, W. Integration, scaling, space-group assignment and post-refinement. Acta Crystallogr. D Biol. Crystallogr. 66, 133–144 (2010).

  58. Otwinowski, Z. & Minor, W. Processing of X-ray diffraction data collected in oscillation mode. Methods Enzymol. 276, 307–326 (1997).

    Article  CAS  PubMed  Google Scholar 

  59. Minor, W., Cymborowski, M., Otwinowski, Z. & Chruszcz, M. HKL -3000: the integration of data reduction and structure solution—from diffraction images to an initial model in minutes. Acta Crystallogr. D Biol. Crystallogr. 62, 859–866 (2006).

    Article  PubMed  Google Scholar 

  60. McCoy, A. J. et al. Phaser crystallographic software. J. Appl. Crystallogr. 40, 658–674 (2007).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  61. Kelley, L. A., Mezulis, S., Yates, C. M., Wass, M. N. & Sternberg, M. J. E. The Phyre2 web portal for protein modeling, prediction and analysis. Nat. Protoc. 10, 845–858 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  62. Adams, P. D. et al. PHENIX: a comprehensive Python-based system for macromolecular structure solution. Acta Crystallogr. D Biol. Crystallogr. 66, 213–221 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  63. Afonine, P. V. et al. Towards automated crystallographic structure refinement with phenix.refine. Acta Crystallogr. D Biol. Crystallogr. 68, 352–367 (2012).

  64. Emsley, P. & Cowtan, K. Coot: model-building tools for molecular graphics. Acta Crystallogr. D Biol. Crystallogr. 60, 2126–2132 (2004).

    Article  PubMed  Google Scholar 

  65. Krissinel, E. & Henrick, K. Inference of macromolecular assemblies from crystalline state. J. Mol. Biol. 372, 774–797 (2007).

    Article  CAS  PubMed  Google Scholar 

  66. The PyMOL Molecular Graphics System, Version 1.7.2.1, Schrödinger, LLC (2014).

  67. Lee, R. A., Razaz, M. & Hayward, S. The DynDom Database of Protein Domain Motions. Bioinformatics 19, 1290–1291 (2003).

    Article  CAS  PubMed  Google Scholar 

  68. van Zundert, G. C. P. et al. The HADDOCK2.2 Web Server: User-Friendly Integrative Modeling of Biomolecular Complexes. Journal of Molecular Biology 428, 720–725 (2016).

    Article  CAS  PubMed  Google Scholar 

  69. Honorato, R. V. et al. Structural Biology in the Clouds: The WeNMR-EOSC Ecosystem. Frontiers in Molecular Biosciences 8, 708 (2021).

    Article  Google Scholar 

Download references

Acknowledgements

We thank the staff at the 23-ID-D, 22-ID and 19-ID beamlines of the Advanced Photon Source for assistance with remote X-ray data collection. We thank D. Gonzalez-Perez and E. Medina for proofreading and thoughtful comments. The following reagents were obtained through the NIH Tetramer Core Facility: biotinylated MHCII proteins including I-A(b)MOG, I-A(b)CLIP, I-A(d)CLIP, I-A(g7)CLIP and I-E(k)CLIP, HLA-DR4CLIP, HLA-DP2CLIP and HLA-DQ2CLIP. V.C.L., Q.M., S.S. and C.M. are supported by a V Scholar grant from the V Foundation, a Rita Allen Scholars grant and NIH grant no. R35GM133482. G.K.A., C.W. and A.C. are supported by NIH grant no. P01AI120943. B.R. is supported by an NIH/National Cancer Institute grant no. R01CA230610. Support for T.H.T. and shared resources were provided by a Moffitt Cancer Center Support NIH grant no. P30CA076292.

Author information

Authors and Affiliations

Authors

Contributions

V.C.L., Q.M., D.P.C. and B.R. designed the experiments. Q.M. and C.M. purified the recombinant proteins. Q.M. performed the structural studies of the LAG3 and FGL1 proteins, including crystallization, data collection, data processing, structure solution and refinement. T.H.T. assisted in data processing and refinement. Q.M. performed the SPR experiments and NFAT reporter signaling assays. Q.M., C.M. and S.S. performed the yeast display experiments. D.P.C. and B.R. performed microscopy imaging and imaging analyses. C.W., A.R.C. and G.K.A. performed the MALS experiments and assisted with data analysis. S.D. assisted with the recombinant MHC class II construct design and provided the HLA-DR4 plasmid. V.C.L. and Q.M. wrote the manuscript.

Corresponding author

Correspondence to Vincent C. Luca.

Ethics declarations

Competing interests

V.C.L. is a consultant on an unrelated project for Cellestia Biotech. The other authors declare no competing interests.

Peer review

Peer review information

Nature Immunology thanks Philippe Pierre and the other, anonymous, reviewer(s) for their contribution to the peer review of this work. N. Bernard was the primary editor on this article and managed its editorial process and peer review in collaboration with the rest of the editorial 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 Yeast display selections and protein engineering of hLAG3.

a, Yeast display selection strategy used to isolate high-affinity FGL1FD binders from the hLAG3D12 mutant library. Flow cytometry dot plots depict the first round (upper panel) and final round of selection (lower panel) stained with 100 nM FGL1FD tetramers and 20 nM FGL1FD monomers, respectively. The gating strategies for yeast and single cells were shown in the bottom. b, Mutation-frequency map generated from sequencing of the five clones isolated from the final round of yeast selection. The recurring mutation, M171I, is highlighted in blue. c, Table of clones containing the M171I mutation. d, Dose-response titrations of yeast expressing hLAG3D12 or LAG3D12 variants with biotinylated FGL1FD. e, SDS-PAGE analysis of LAG3 protein expression and secretion. Two replicated transfections were performed for hLAG3, hLAG3*, hLAG3D12, hLAG3*D12 (hLAG3D12 with M171I mutation). High-Five cells were inoculated with consistent titer of virus. After 48 hours, proteins were retrieved using Ni-NTA. Quantifications of the relative intensity of the bands (right panel) were based on triplicates inoculation.

Extended Data Fig. 2 Analysis of LAG3 structures.

a, Electron density maps of hLAG3*:F7 complex structure at 3.71 Å. Upper panel: the composite omit map; lower panel: the 2Fo-Fc map from the final round of refinement. The maps were contoured at 0.8 σ. b, Alignment of D3 and D4 domains from the two protomers in the hLAG3 dimer reveals a rotation angle of 161° about the D2-D3 hinge. The two protomers are colored in cyan and grey. c, the M171I mutation and its surrounding residues in hLAG3* structure. M167 in mLAG3 is the equivalent site to M171I. d, Structural comparison of D1 domains from LAG3 (mLAG3) and CD4 (PDB ID: 3T0E). MHCII-binding residues of CD4 D1 are shown as sticks and are clustered in an analogous region to LAG3 Loop2. CD4: pink; mLAG3: green; Loop2: blue.

Extended Data Fig. 3 LAG3 dimer interface and LAG3:MHCII binding analyses with SPR.

a, The 2Fo-Fc electron density map, contoured at 1.0 σ, surrounding the mLAG3 dimer interface residue Trp180. b, SPR analyses of hLAG3* binding to HLA-DR4HIV. Biotinylated HLA-DR4HIV was immobilized on a SA chip. hLAG3* protein dilutions starting from 20,000 nM were injected in turn. c,d, SPR sensorgram of recombinant mLAG3 flowed over an SA-chip immobilized with I-A(b)MOG (c) or I-A(b)CLIP (d). e, SPR analyses of LAG3 binding affinity binding to human MHCII allomorphs. MHCII proteins bound to CLIP peptides were biotinylated and immobilized on an SA chip and hLAG3* proteins were flowed over the chip. The curves were then fitted to determine the KD. Mean KD ± s.d. was from two independent replicates. f, Sequence alignment of the D2-dimer interface region from multiple LAG3 orthologs. Residues forming dimer contacts are outlined in red, conserved residues are highlighted in green, and biochemically similar residues are colored in green.

Extended Data Fig. 4 LAG3 and MHCII protein colocalization on cell surface.

a,b, Untreated HuT 78 cells or HuT 78 cells treated with anti-CD3 (1 ug/ml), anti-CD3 /anti-CD28 (1 ug/ml) or PMA (50 ng/ml) for 48 h and LAG3 expression was analyzed by flow cytometry. c, Flow cytometry analyses of HLA-DR expression on unstimulated or stimulated Hut-78 cells. The cells were stained with anti-HLA-DR (clone L243, BioLegend, 1:100). d, Subcellular localization of LAG-3 and HLA-DR on HuT-78 cells. Confocal fluorescence images displaying nuclei (blue), LAG-3 (green) and HLA-DR (red) in HuT-78 cells, either untreated or treated with anti-CD3/anti-CD28. Representative confocal images for the individual channels and merged images are shown in the zoomed panel. One of three representative experiments is shown. Horizontal bars in (b) represent the mean and statistics was determined by one-way ANOVA with P values noted in the figure.

Extended Data Fig. 5 Epitope mapping and functional characterization of LAG3 antagonist antibodies.

ac, SPR sensograms measuring hLAG3 binding to MHCII in the presence of scFv-Fc fusion proteins. d, SPR was used to determine scFv:LAG3 binding affinities. e, Flow cytometry analyses of binding of recombinant HLA-DR4HIV(4 µM) to yeast-displayed CD4 or hLAG3 and its mutants. f,g, A luciferase reporter assay was used to assess the ability of 15011-Fc to inhibit LAG3 in the presence of FGL1. To activate TCR signaling, LAG3+ Jurkat cells were either co-cultured with MHCII-expressing Raji cells (f) or stimulated with anti-CD3 (α-CD3) (g). In f and g, the cells were supplemented with 10 nM FGL1 and an isotype control antibody, 15011-Fc, or F7-Fc (300 nM). The luminescence was measured after incubating the cells overnight. In f and g, the graph represents mean ± SD of three replicates from representative of two independent experiments. All statistics was determined by one-way ANOVA, with P values noted in the figure.

Extended Data Fig. 6 Mapping of FGL1 binding sites on hLAG3.

a, Flow chart depicting the selection strategy used to isolate mutations that decreased LAG3 binding to FGL1. In the first two rounds, the hLAG3D12 library was negatively selected against Alexa Fluor 647-labeled streptavidin to remove non-specific binders. In rounds 3 through 5, the indicated concentrations of biotinylated FGL1FD were incubated with the library and negative selections were performed to remove FGL1 binders. b, Yeast expressing LAG3 mutants were stained with MHCII tetramers (200 nM) to determine whether the FGL1 loss-of-binding mutations affected LAG3 binding to MHCII. c, SPR-sensograms recorded following the injection of samples containing fixed concentrations of hLAG3 (800 nM) and varying concentrations 15011.scFv over a chip coated with FGL1FD.

Extended Data Fig. 7 Mapping the binding interface between FGL1FD and hLAG3.

a, Structural alignment of the three copies of FGL1FD (FD-a, FD-b and FD-c) in the crystal asymmetric unit. Different P subdomain loop conformations differences are depicted in the zoom panel. b, Structural alignment of fibrinogen-like domains from FGL1, Ang1 (RMSD = 1.12 Å) and FIBCD1 (RMSD = 1.21 Å). The FD-b protomer from the FGL1FD structure was used in this alignment. FGL1, blue; Ang1, yellow; FICD1, gray. c, Grouped residues mutated in the alanine scanning assay. Residues that were mutated in a single construct are colored accordingly. d, Flow cytometry histogram plots depicting increased binding of FGL1FD variant populations following iterative rounds of selection against hLAG3 (left) and a heat map showing the mutation frequency of amino acid substitutions after sequencing six clones (right). e, Yeast expressing hLAG3 were stained with biotinylated FGL1FD and FGL1FD containing the affinity-enhancing G290E mutation (FDG290E).

Extended Data Fig. 8 FGL1-induced LAG3 clustering correlates with FGL1 suppression mechanism.

a, LAG3-NFAT Jurkat T cells were treated with 1, 10 or 100 nM of FGL1 protein and incubated for 30 min. Binding was detected using an anti-LAG3 antibody. Data reflect the mean ± SD of n = 3 technical replicates, with representative of two experiments is shown. b, In the presence of the blocking antibody, 15011Fc, the FGL1-LAG3 clustering effect was diminished. Stacked confocal microscopy images displaying LAG3 (green) and DAPI (blue) in Jurkat T cells expressing LAG3, either untreated or treated with the indicated reagents: 10 nM FGL1, 15011Fc, isotype control, 10 nM FGL1 and 300 nM 15011Fc, or 10 nM FGL1 and 300 nM isotype control. c, The number of LAG3 clusters per cell was quantified and plotted as the average per field of view (FOV) with analyses of n = 7 FOV from one of three representative experiments. d, NFAT luciferase reporter assay performed to assess the effect of FGL1 on LAG3-mediated suppression of T cell activation mediated by anti-CD3. The relative activation was obtained by normalizing baseline-subtracted luciferase signals of tested concentrations to the buffer control group. The graph represents mean ± SD of three independent replicates from one of two representative experiments. All significance was determined by one-way ANOVA, with P values noted in the figure.

Supplementary information

Supplementary Information

Supplementary Tables 1 and 2.

Reporting Summary

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Ming, Q., Celias, D.P., Wu, C. et al. LAG3 ectodomain structure reveals functional interfaces for ligand and antibody recognition. Nat Immunol 23, 1031–1041 (2022). https://doi.org/10.1038/s41590-022-01238-7

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41590-022-01238-7

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