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:

Synaptic adhesion molecule IgSF11 regulates synaptic transmission and plasticity

Abstract

Synaptic adhesion molecules regulate synapse development and plasticity through mechanisms that include trans-synaptic adhesion and recruitment of diverse synaptic proteins. We found that the immunoglobulin superfamily member 11 (IgSF11), a homophilic adhesion molecule that preferentially expressed in the brain, is a dual-binding partner of the postsynaptic scaffolding protein PSD-95 and AMPA glutamate receptors (AMPARs). IgSF11 required PSD-95 binding for its excitatory synaptic localization. In addition, IgSF11 stabilized synaptic AMPARs, as determined by IgSF11 knockdown–induced suppression of AMPAR-mediated synaptic transmission and increased surface mobility of AMPARs, measured by high-throughput, single-molecule tracking. IgSF11 deletion in mice led to the suppression of AMPAR-mediated synaptic transmission in the dentate gyrus and long-term potentiation in the CA1 region of the hippocampus. IgSF11 did not regulate the functional characteristics of AMPARs, including desensitization, deactivation or recovery. These results suggest that IgSF11 regulates excitatory synaptic transmission and plasticity through its tripartite interactions with PSD-95 and AMPARs.

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

Figure 1: IgSF11 interacts with PSD-95 and is targeted to excitatory synapses in a PDZ interaction–dependent manner.
Figure 2: IgSF11 interacts with AMPARs and recruits both PSD-95 and AMPARs to sites of homophilic IgSF11 adhesion.
Figure 3: IgSF11 knockdown reduces surface clustering of AMPARs without affecting PSD-95 clustering.
Figure 4: IgSF11 knockdown suppresses mEPSCs and evoked EPSCs.
Figure 5: IgSF11 knockdown increases surface mobility and endocytosis of AMPARs.
Figure 6: Dominant-negative inhibition of IgSF11 suppresses mEPSC amplitude in DG granule cells.
Figure 7: IgSF11 has no effect on the functional properties of AMPARs.
Figure 8: Reduced mEPSC amplitude in DG cells and suppressed LTP at SC-CA1 synapses in the Igsf11−/− hippocampus.

Similar content being viewed by others

References

  1. Yuzaki, M. Cbln1 and its family proteins in synapse formation and maintenance. Curr. Opin. Neurobiol. 21, 215–220 (2011).

    Article  CAS  PubMed  Google Scholar 

  2. Takahashi, H. & Craig, A.M. Protein tyrosine phosphatases PTPdelta, PTPsigma, and LAR: presynaptic hubs for synapse organization. Trends Neurosci. 36, 522–534 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  3. Krueger, D.D., Tuffy, L.P., Papadopoulos, T. & Brose, N. The role of neurexins and neuroligins in the formation, maturation, and function of vertebrate synapses. Curr. Opin. Neurobiol. 22, 412–422 (2012).

    Article  CAS  PubMed  Google Scholar 

  4. de Wit, J., Hong, W., Luo, L. & Ghosh, A. Role of leucine-rich repeat proteins in the development and function of neural circuits. Annu. Rev. Cell Dev. Biol. 27, 697–729 (2011).

    Article  CAS  PubMed  Google Scholar 

  5. Shen, K. & Scheiffele, P. Genetics and cell biology of building specific synapse connectivity. Annu. Rev. Neurosci. 33, 473–507 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  6. Dalva, M.B., McClelland, A.C. & Kayser, M.S. Cell adhesion molecules: signalling functions at the synapse. Nat. Rev. Neurosci. 8, 206–220 (2007).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  7. Biederer, T. & Stagi, M. Signaling by synaptogenic molecules. Curr. Opin. Neurobiol. 18, 261–269 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  8. Um, J.W. & Ko, J. LAR-RPTPs: synaptic adhesion molecules that shape synapse development. Trends Cell Biol. 23, 465–475 (2013).

    Article  CAS  PubMed  Google Scholar 

  9. Südhof, T.C. Neuroligins and neurexins link synaptic function to cognitive disease. Nature 455, 903–911 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  10. Nuriya, M. & Huganir, R.L. Regulation of AMPA receptor trafficking by N-cadherin. J. Neurochem. 97, 652–661 (2006).

    Article  CAS  PubMed  Google Scholar 

  11. Saglietti, L. et al. Extracellular interactions between GluR2 and N-cadherin in spine regulation. Neuron 54, 461–477 (2007).

    Article  CAS  PubMed  Google Scholar 

  12. Budreck, E.C. et al. Neuroligin-1 controls synaptic abundance of NMDA-type glutamate receptors through extracellular coupling. Proc. Natl. Acad. Sci. USA 110, 725–730 (2013).

    Article  CAS  PubMed  Google Scholar 

  13. Herrera-Molina, R. et al. Structure of excitatory synapses and GABAA receptor localization at inhibitory synapses are regulated by neuroplastin-65. J. Biol. Chem. 289, 8973–8988 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  14. Suzu, S. et al. Molecular cloning of a novel immunoglobulin superfamily gene preferentially expressed by brain and testis. Biochem. Biophys. Res. Commun. 296, 1215–1221 (2002).

    Article  CAS  PubMed  Google Scholar 

  15. Harada, H., Suzu, S., Hayashi, Y. & Okada, S. BT-IgSF, a novel immunoglobulin superfamily protein, functions as a cell adhesion molecule. J. Cell. Physiol. 204, 919–926 (2005).

    Article  CAS  PubMed  Google Scholar 

  16. Eom, D.S. et al. Melanophore migration and survival during zebrafish adult pigment stripe development require the immunoglobulin superfamily adhesion molecule Igsf11. PLoS Genet. 8, e1002899 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  17. Brennand, K.J. et al. Modelling schizophrenia using human induced pluripotent stem cells. Nature 473, 221–225 (2011).

    CAS  PubMed  PubMed Central  Google Scholar 

  18. Raschperger, E., Engstrom, U., Pettersson, R.F. & Fuxe, J. CLMP, a novel member of the CTX family and a new component of epithelial tight junctions. J. Biol. Chem. 279, 796–804 (2004).

    Article  CAS  PubMed  Google Scholar 

  19. Opazo, P., Sainlos, M. & Choquet, D. Regulation of AMPA receptor surface diffusion by PSD-95 slots. Curr. Opin. Neurobiol. 22, 453–460 (2012).

    Article  CAS  PubMed  Google Scholar 

  20. Giannone, G. et al. Dynamic superresolution imaging of endogenous proteins on living cells at ultra-high density. Biophys. J. 99, 1303–1310 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  21. Schnell, E. et al. Direct interactions between PSD-95 and stargazin control synaptic AMPA receptor number. Proc. Natl. Acad. Sci. USA 99, 13902–13907 (2002).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  22. Schreiber, J., Langhorst, H., Jüttner, R. & Rathjen, F.G. The IgCAMs CAR, BT-IgSF and CLMP: structure, function and diseases. in Cell Adhesion Molecules: Implications in Neurological Diseases (eds. Berezin, V. & Walmod, P.S.) (Springer, 2014).

  23. Siddiqui, T.J. et al. An LRRTM4-HSPG complex mediates excitatory synapse development on dentate gyrus granule cells. Neuron 79, 680–695 (2013).

    Article  CAS  PubMed  Google Scholar 

  24. Jackson, A.C. & Nicoll, R.A. The expanding social network of ionotropic glutamate receptors: TARPs and other transmembrane auxiliary subunits. Neuron 70, 178–199 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  25. Chih, B., Engelman, H. & Scheiffele, P. Control of excitatory and inhibitory synapse formation by neuroligins. Science 307, 1324–1328 (2005).

    Article  CAS  PubMed  Google Scholar 

  26. Soler-Llavina, G.J., Fuccillo, M.V., Ko, J., Sudhof, T.C. & Malenka, R.C. The neurexin ligands, neuroligins and leucine-rich repeat transmembrane proteins, perform convergent and divergent synaptic functions in vivo. Proc. Natl. Acad. Sci. USA 108, 16502–16509 (2011).

    CAS  PubMed  PubMed Central  Google Scholar 

  27. Chubykin, A.A. et al. Activity-dependent validation of excitatory versus inhibitory synapses by neuroligin-1 versus neuroligin-2. Neuron 54, 919–931 (2007).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  28. Jung, S.Y. et al. Input-specific synaptic plasticity in the amygdala is regulated by neuroligin-1 via postsynaptic NMDA receptors. Proc. Natl. Acad. Sci. USA 107, 4710–4715 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  29. Blundell, J. et al. Increased anxiety-like behavior in mice lacking the inhibitory synapse cell adhesion molecule neuroligin 2. Genes Brain Behav. 8, 114–126 (2009).

    Article  CAS  PubMed  Google Scholar 

  30. Shipman, S.L. & Nicoll, R.A. A subtype-specific function for the extracellular domain of neuroligin 1 in hippocampal LTP. Neuron 76, 309–316 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  31. Kim, J. et al. Neuroligin-1 is required for normal expression of LTP and associative fear memory in the amygdala of adult animals. Proc. Natl. Acad. Sci. USA 105, 9087–9092 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  32. Mondin, M. et al. Neurexin-neuroligin adhesions capture surface-diffusing AMPA receptors through PSD-95 scaffolds. J. Neurosci. 31, 13500–13515 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  33. Shipman, S.L. et al. Functional dependence of neuroligin on a new non-PDZ intracellular domain. Nat. Neurosci. 14, 718–726 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  34. Cingolani, L.A. et al. Activity-dependent regulation of synaptic AMPA receptor composition and abundance by beta3 integrins. Neuron 58, 749–762 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  35. Pozo, K. et al. beta3 integrin interacts directly with GluA2 AMPA receptor subunit and regulates AMPA receptor expression in hippocampal neurons. Proc. Natl. Acad. Sci. USA 109, 1323–1328 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  36. Tai, C.Y., Kim, S.A. & Schuman, E.M. Cadherins and synaptic plasticity. Curr. Opin. Cell Biol. 20, 567–575 (2008).

    Article  CAS  PubMed  Google Scholar 

  37. Song, M. et al. Slitrk5 mediates BDNF-dependent TrkB receptor trafficking and signaling. Dev. Cell 33, 690–702 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  38. Taniguchi, H. et al. Silencing of neuroligin function by postsynaptic neurexins. J. Neurosci. 27, 2815–2824 (2007).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  39. Schwenk, J. et al. Functional proteomics identify cornichon proteins as auxiliary subunits of AMPA receptors. Science 323, 1313–1319 (2009).

    Article  CAS  PubMed  Google Scholar 

  40. Shanks, N.F. et al. Differences in AMPA and kainate receptor interactomes facilitate identification of AMPA receptor auxiliary subunit GSG1L. Cell Reports 1, 590–598 (2012).

    Article  CAS  PubMed  Google Scholar 

  41. von Engelhardt, J. et al. CKAMP44: a brain-specific protein attenuating short-term synaptic plasticity in the dentate gyrus. Science 327, 1518–1522 (2010).

    Article  CAS  PubMed  Google Scholar 

  42. Kalashnikova, E. et al. SynDIG1: an activity-regulated, AMPA receptor–interacting transmembrane protein that regulates excitatory synapse development. Neuron 65, 80–93 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  43. Chen, L. et al. Stargazin regulates synaptic targeting of AMPA receptors by two distinct mechanisms. Nature 408, 936–943 (2000).

    Article  CAS  PubMed  Google Scholar 

  44. Schwenk, J. et al. High-resolution proteomics unravel architecture and molecular diversity of native AMPA receptor complexes. Neuron 74, 621–633 (2012).

    Article  CAS  PubMed  Google Scholar 

  45. Lee, K. et al. MDGAs interact selectively with neuroligin-2, but not other neuroligins to regulate inhibitory synapse development. Proc. Natl. Acad. Sci. USA 110, 336–341 (2013).

    Article  CAS  PubMed  Google Scholar 

  46. Pettem, K.L., Yokomaku, D., Takahashi, H., Ge, Y. & Craig, A.M. Interaction between autism-linked MDGAs and neuroligins suppresses inhibitory synapse development. J. Cell Biol. 200, 321–336 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  47. Fogel, A.I., Stagi, M., Perez de Arce, K. & Biederer, T. Lateral assembly of the immunoglobulin protein SynCAM 1 controls its adhesive function and instructs synapse formation. EMBO J. 30, 4728–4738 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  48. Song, Y.S., Lee, H.J., Prosselkov, P., Itohara, S. & Kim, E. Trans-induced cis interaction in the tripartite NGL-1, netrin-G1 and LAR adhesion complex promotes development of excitatory synapses. J. Cell Sci. 126, 4926–4938 (2013).

    Article  CAS  PubMed  Google Scholar 

  49. Patzke, C. et al. The coxsackievirus-adenovirus receptor reveals complex homophilic and heterophilic interactions on neural cells. J. Neurosci. 30, 2897–2910 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  50. Lucic´, V., Yang, T., Schweikert, G., Forster, F. & Baumeister, W. Morphological characterization of molecular complexes present in the synaptic cleft. Structure 13, 423–434 (2005).

    Article  PubMed  CAS  Google Scholar 

  51. Kim, E., Niethammer, M., Rothschild, A., Jan, Y.N. & Sheng, M. Clustering of Shaker-type K+ channels by interaction with a family of membrane-associated guanylate kinases. Nature 378, 85–88 (1995).

    Article  CAS  PubMed  Google Scholar 

  52. Choi, J. et al. Regulation of dendritic spine morphogenesis by insulin receptor substrate 53, a downstream effector of Rac1 and Cdc42 small GTPases. J. Neurosci. 25, 869–879 (2005).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  53. Neve, R.L., Neve, K.A., Nestler, E.J. & Carlezon, W.A. Jr. Use of herpes virus amplicon vectors to study brain disorders. Biotechniques 39, 381–391 (2005).

    Article  CAS  PubMed  Google Scholar 

  54. Kim, J., Kwon, J.T., Kim, H.S., Josselyn, S.A. & Han, J.H. Memory recall and modifications by activating neurons with elevated CREB. Nat. Neurosci. 17, 65–72 (2014).

    Article  CAS  PubMed  Google Scholar 

  55. Yang, J. et al. DGKiota regulates presynaptic release during mGluR-dependent LTD. EMBO J. 30, 165–180 (2011).

    Article  CAS  PubMed  Google Scholar 

  56. Kim, M.H. et al. Enhanced NMDA receptor-mediated synaptic transmission, enhanced long-term potentiation, and impaired learning and memory in mice lacking IRSp53. J. Neurosci. 29, 1586–1595 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  57. Sheng, M., Cummings, J., Roldan, L.A., Jan, Y.N. & Jan, L.Y. Changing subunit composition of heteromeric NMDA receptors during development of rat cortex. Nature 368, 144–147 (1994).

    Article  CAS  PubMed  Google Scholar 

  58. Wyszynski, M., Kim, E., Yang, F.C. & Sheng, M. Biochemical and immunocytochemical characterization of GRIP, a putative AMPA receptor anchoring protein, in rat brain. Neuropharmacology 37, 1335–1344 (1998).

    Article  CAS  PubMed  Google Scholar 

  59. Han, S. et al. Altered expression of synaptotagmin 13 mRNA in adult mouse brain after contextual fear conditioning. Biochem. Biophys. Res. Commun. 425, 880–885 (2012).

    Article  CAS  PubMed  Google Scholar 

  60. Huttner, W.B., Schiebler, W., Greengard, P. & De Camilli, P. Synapsin I (protein I), a nerve terminal-specific phosphoprotein. III. Its association with synaptic vesicles studied in a highly purified synaptic vesicle preparation. J. Cell Biol. 96, 1374–1388 (1983).

    Article  CAS  PubMed  Google Scholar 

  61. Cho, K.O., Hunt, C.A. & Kennedy, M.B. The rat brain postsynaptic density fraction contains a homolog of the Drosophila discs-large tumor suppressor protein. Neuron 9, 929–942 (1992).

    Article  CAS  PubMed  Google Scholar 

  62. Han, K. et al. Regulated RalBP1 binding to RalA and PSD-95 controls AMPA receptor endocytosis and LTD. PLoS Biol. 7, e1000187 (2009).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  63. Jo, J. et al. Abeta(1–42) inhibition of LTP is mediated by a signaling pathway involving caspase-3, Akt1 and GSK-3beta. Nat. Neurosci. 14, 545–547 (2011).

    Article  CAS  PubMed  Google Scholar 

  64. Giannone, G. et al. Dynamic super-resolution imaging of endogenous proteins on living cells at ultra-high density. Biophys. J. 99, 1303–1310 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  65. Groc, L. et al. Differential activity-dependent regulation of the lateral mobilities of AMPA and NMDA receptors. Nat. Neurosci. 7, 695–696 (2004).

    Article  CAS  PubMed  Google Scholar 

  66. Skarnes, W.C. et al. A conditional knockout resource for the genome-wide study of mouse gene function. Nature 474, 337–342 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  67. O'Gorman, S., Dagenais, N.A., Qian, M. & Marchuk, Y. Protamine-Cre recombinase transgenes efficiently recombine target sequences in the male germ line of mice, but not in embryonic stem cells. Proc. Natl. Acad. Sci. USA 94, 14602–14607 (1997).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  68. Carlin, R.K., Grab, D.J., Cohen, R.S. & Siekevitz, P. Isolation and characterization of postsynaptic densities from various brain regions: enrichment of different types of postsynaptic densities. J. Cell Biol. 86, 831–845 (1980).

    Article  CAS  PubMed  Google Scholar 

  69. Won, H. et al. GIT1 is associated with ADHD in humans and ADHD-like behaviors in mice. Nat. Med. 17, 566–572 (2011).

    Article  CAS  PubMed  Google Scholar 

  70. Schoch, S. et al. RIM1alpha forms a protein scaffold for regulating neurotransmitter release at the active zone. Nature 415, 321–326 (2002).

    Article  CAS  PubMed  Google Scholar 

  71. Patzke, C. et al. The coxsackievirus-adenovirus receptor reveals complex homophilic and heterophilic interactions on neural cells. J. Neurosci. 30, 2897–2910 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  72. Arnold, K., Bordoli, L., Kopp, J. & Schwede, T. The SWISS-MODEL workspace: a web-based environment for protein structure homology modelling. Bioinformatics 22, 195–201 (2006).

    Article  CAS  PubMed  Google Scholar 

  73. Lyskov, S. & Gray, J.J. The RosettaDock server for local protein-protein docking. Nucleic Acids Res. 36, W233–W238 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  74. Schrodinger, LLC. The PyMOL Molecular Graphics System, Version 1.3r1, (http://pymol.org) (2010).

Download references

Acknowledgements

We would like to thank R. Neve for the kind gift of the HSV system, J.-R. Lee and K. Han for the PSD fraction sample, J.W. Um and J. Ko for the subcellular fraction sample, M.-H. Kim for the mini analysis program, and Mihyeon Kim for helping with IgSF11 antibody generation. This work was supported by the Institute for Basic Science (to E.K.), the Brain Research Program through the National Research Foundation of Korea(NRF) funded by the Ministry of Science, ICT & Future Planning (2013-056732 to H.K.), National Honor Scientist Program of the NRF (to B.K.K.), and Wellcome Trust and UK MRC Neurodegenerative Disease Initiative Programme (to K.C.).

Author information

Authors and Affiliations

Authors

Contributions

D.L. and H.K. performed the in situ hybridization analysis. S.J., D.O., Y.L., H.S., D.K. and S.G.K. performed immunoblot and co-immunoprecipitation experiments. D.O., S.K.K., J.W. and H.S. performed neuron culture, co-culture, co-clustering experiments, homophilic binding assay and antibody feeding experiments. D.O., Y.L. and H.S. performed neuronal shRNA knockdown experiments. S.J., C.V.R., D.W., J.M.W., J.J., S.G.K., S.M.U., S.K.K., M.H.K., J.D.R., H.J., W.M. and K.C. performed electrophysiological experiments and analyzed the data. E.H. performed uPAINT experiments. Y.L. and J.D.R. carried out HSV experiments. S.J. generated and characterized Igsf11–/– (LacZ) and Igsf11–/– mice. D.K. modeled IgSF11 structures. H.K., B.K.K., K.C., J.S.R., D.C. and E.K. supervised the project and wrote the manuscript.

Corresponding authors

Correspondence to Jeong-Seop Rhee, Daniel Choquet or Eunjoon Kim.

Ethics declarations

Competing interests

The authors declare no competing financial interests.

Integrated supplementary information

Supplementary Figure 1 Amino acid sequence alignment of IgSF11 proteins from various vertebrate species.

Amino acids in black and gray backgrounds indicate fully and partially conserved residues. The following amino acid sequences were used for comparison; human (NP_001015887.1), chimpanzee (XP_001161500.1), macaque monkey (NP_001244835.1), mouse (NP_733548.2), rat (NP_001013138.1), chicken (XP_416568.3), finch (ENSTGUP00000013797), lizard (ENSACAP00000002845), Xenopus (ENSXETP00000049789), and zebrafish (XP_001343768.4).

Supplementary Figure 2 Amino acid sequence alignment of CAR subgroup members and their in situ hybridization patterns in the brain.

(a) Amino acids in black and gray backgrounds indicate fully and partially conserved residues. The following mouse amino acid sequences were used for comparison; IgSF11 (NP_733548.2), CAR (NP_001020363.1), ESAM (NP_081378.1), and CLMP (NP_598494.2). (b) In situ mRNA distribution patterns of IgSF11, CAR, ESAM, and CLMP in adult (8 weeks) male mouse brains obtained from the Allen Brain Atlas (IgSF11, http://mouse.brain-map.org/experiment/show?id=68861816, CAR, http://mouse.brain-map.org/experiment/show?id=77454980, ESAM, http://mouse.brain-map.org/experiment/show?id=69262331, and CLMP, http://mouse.brain-map.org/experiment/show?id=70445538). Scale bar, 1 mm.

Supplementary Figure 3 IgSF11 interacts with PSD-95 in yeast two-hybrid assays.

The C-terminal PDZ-binding motif of IgSF11 interacts with PDZ domains of PSD-95 family proteins (PSD-95, PSD-93/chapsyn-110 and SAP97) in yeast two-hybrid assays. pBHA, bait vector; pGAD10, prey vector; WT, wild type (last seven residues); VA, a mutant IgSF11 in which the last residue (valine) was changed to alanine. PDZ domains from GRIP2, Shank1 and NHERF were used as controls. β-Galactosidase (β-gal) activity: +++, < 45 min; ++, 45–90 min; +, 90–240 min; –, no significant β-gal activity. Three independent experiments were performed.

Supplementary Figure 4 IgSF11 does not induce clustering of pre- or postsynaptic proteins in contacting axons and dendrites, and does not interact with other synaptic adhesion molecules.

(a-c) Cultured rat hippocampal neurons were cocultured with HEK293T cells expressing rat IgSF11 (untagged), or EGFP (control; DIV 10–13), followed by staining for synapsin I (a presynaptic protein), PSD-95 (an excitatory postsynaptic protein), and gephyrin (an inhibitory postsynaptic protein). Scale bar, 30 μm. (d) IgSF11 does not interact with other synaptic adhesion molecules, as shown by the lack of binding of the ectodomain of IgSF11 fused to human Fc (IgSF11-ecto-Fc) to known synaptic adhesion molecules expressed in HEK293T cells. IgSF11-expressing HEK293T cells were used as a control. Scale bar, 10 μm. Three independent experiments were performed.

Supplementary Figure 5 The transmembrane domain of IgSF11 is important for GluA1 binding.

(a and b) HEK293T cells were cotransfected with GluA1 and the indicated deletion variants of HA-IgSF11, followed by immunoprecipitation with HA antibodies and immunoblotting with HA and GluA1 antibodies. Three independent experiments were performed. Full-length blots are presented in Supplementary Figure 17c.

Supplementary Figure 6 Characterization of IgSF11-knockdown and IgSF11 rescue constructs.

(a) Acute knockdown of IgSF11 expression in heterologous cells. HEK293T cells were doubly transfected with rIgSF11 (rat IgSF11) and sh-vec (empty vector), sh-IgSF11#1, or sh-IgSF11#1* (a mutant incapable of knocking down IgSF11) for two days, and the cell lysates were characterized by immunoblotting for IgSF11 and α-tubulin (loading control). n = 3, * p < 0.05, ** p < 0.01, ns, not significant, one-way ANOVA. Full-length blots are presented in Supplementary Figure 17d.

(b) Acute knockdown of IgSF11 expression by an independent IgSF11 knockdown construct (sh-IgSF11#2). HEK293T cells were transfected with doubly transfected with IgSF11 and sh-vec (empty vector) or sh-IgSF11#2 for two days, and the cell lysates were characterized by immunoblotting for IgSF11 and α-tubulin (loading control). n = 3, *** p < 0.001, Student’s t-test. Full-length blots are presented in Supplementary Figure 17e.

(c and d) Characterization of an sh-IgSF11#1-resistant IgSF11 expression construct. HEK293T cells were doubly transfected with IgSF11 (rat) + sh-vec, IgSF11 + sh-IgSF11#1, or sh-IgSF11#1-resistant IgSF11 (WT or ΔC3) + sh-IgSF11#1, followed by immunoblotting with IgSF11 antibodies. Three independent experiments were performed. Full-length blots are presented in Supplementary Figure 17f,g.

Supplementary Figure 7 IgSF11 knockdown in cultured neurons suppresses surface clustering of GluA1 and GluA2 but has no effect on excitatory synapses.

(a-c) Acute knockdown of IgSF11 decreases surface clustering of GluA1 in cultured hippocampal neurons. Cultured neurons were transfected with sh-vec (empty vector), sh-IgSF11#1, or sh-IgSF11#1* (DIV 14–17), and stained for surface GluA1. au, arbitrary unit; n = 18 for sh-vec, 17 for sh-IgSF11#1, and 17 for sh-IgSF11#1*, *p < 0.05, **p < 0.01, ***p < 0.001, one-way ANOVA. Scale bar, 10 μm.

(d-g) IgSF11 knockdown has no effect on excitatory synapses, defined by synapsin I-positive PSD-95 clusters. Cultured neurons were transfected with sh-vec, sh-IgSF11#1, or sh-IgSF11#1* (DIV14–17), and immunostained for PSD-95 and synapsin I. n = 22 for sh-vec, 23 for sh-IgSF11#1, and 21 for sh-IgSF11#1*, ns, not significant, one-way ANOVA. Scale bar, 10 μm.

(h-k) Independent IgSF11 knockdown suppresses surface GluA2 clustering but has no effect on PSD-95 clusters. Cultured hippocampal neurons were transfected with sh-vec (empty vector) or sh-IgSF11#2 (DIV 14–17), and stained for surface GluA2 or PSD-95. au, arbitrary unit; n = 18 for sh-vec, 17 for sh-IgSF11#2, ***p < 0.001, ns, not significant, Student’s t-test. Scale bar, 10 μm.

Supplementary Figure 8 IgSF11 knockdown reduces steady-state levels of surface GluA2.

Cultured hippocampal neurons doubly transfected with N-terminally HA-tagged GluA2 (HA-GluA2) and sh-vec, sh-IgSF11#1, or IgSF11#1* (DIV 14–17) were measured of surface and internal levels of GluA2. n = 19 for sh-vec, 19 for sh-IgSF11#1, and 18 for sh-IgSF11#1*, *p < 0.05, **p < 0.01, one-way ANOVA. Scale bar, 30 μm.

Supplementary Figure 9 Generation and characterization of Igsf11−/− (LacZ) and Igsf11−/− mice.

(a) Schematic diagram showing the strategy for the generation of the gene-trapped Igsf11−/− (LacZ) mice and exon 3-deleted Igsf11−/− mice. Note that exon 2 is captured by the inserted cassette to yield a truncated protein fused to β-geo (β-galactosidase + neomycin), and that crossing of Igsf11 floxed mice with protamine-Cre mice leads to the deletion of exon 3 in the IgSF11 gene. Primer locations for genotyping are indicated. (b) Genotyping of wild-type (Igsf11+/+; WT), Igsf11+/− (LacZ), and Igsf11−/− (LacZ) mice by PCR. Full-length gels are presented in Supplementary Figure 17h. (c) Igsf11−/− (LacZ) mice show no detectable levels of IgSF11 proteins in the brain. α-Tubulin blot was used as a control. Full-length blots are presented in Supplementary Figure 17i.

Supplementary Figure 10 Spatiotemporal IgSF11 expression revealed by X-Gal staining.

(a and b) The spatial expression pattern of IgSF11 was characterized by X‑Gal staining of sagittal brain slices from Igsf11−/− (LacZ) mice at P5 and postnatal weeks 2, 3, and 12. Scale bars, 1 mm for whole brain, 200 μm for hippocampus. Three independent experiments were performed.(c) No differences are observed in left and right Igsf11−/− (LacZ) hippocampal regions at 8 weeks. (d) IgSF11 protein levels in mouse brains are rapidly increased during the first and second weeks and are maintained at later stages. Three independent experiments were performed. Full-length blots are presented in Supplementary Figure 17j.

Supplementary Figure 11 Normal expression levels of synaptic proteins in the Igsf11−/− brain.

(a and b) Hippocampal samples from WT and Igsf11−/− (LacZ) brains (3 weeks) were immunoblotted for the synaptic proteins indicated, followed by quantification. Protein amounts in Igsf11−/− brains normalized to α-tubulin were normalized to those of WT brains. n = 3 mice for WT and KO, ns, not significant, Student’s t-test. Full-length blots are presented in Supplementary Figure 17k.

Supplementary Figure 12 Igsf11−/− hippocampal CA1 pyramidal neurons display unaltered levels extrasynaptic AMPAR currents.

(a and b) Igsf11−/− (LacZ) DG granule cells in acute hippocampal slices (3 weeks) were treated with AMPA (1 μM), a condition known to induce extrasynaptic AMPAR-mediated currents, and measured of evoked EPSCs. n = 8 slices (2 mice) for WT and KO, ns, not significant, Student’s t-test.

Supplementary Figure 13 Normal levels of basal transmission and paired pulse ratio at Igsf11−/− MPP-DG and SC-CA1 synapses.

(a and b) Basal transmission (input-out relationship) and paired pulse ratio at WT and Igsf11−/− (LacZ) MPP-DG synapses (3–6 months). n = 12 slices (6 mice) for WT, and 10, 6 for KO. (c and d) Basal transmission and paired pulse ratio at WT and Igsf11−/− (LacZ) SC-CA1 synapses (3–6 months). n = 10, 6 for WT and KO.

Supplementary Figure 14 The Igsf11−/− hippocampus shows normal levels of immature GluA2.

Hippocampal lysates from WT and Igsf11−/− (LacZ) mice were digested with endoglycosidase H (Endo H) or peptide N-glycosidase F (PNGase F), and measured of the ratio of Endo H-sensitive GluA2 over total (Endo H-resistant + Endo H-sensitive) GluA2. PNGase F, which completely removes all types of N-linked oligosaccharides on GluA2, was used as a control. n = 3 mice for WT and KO (3 weeks), ns, not significant, Student’s t-test. Full-length blots are presented in Supplementary Figure 17l.

Supplementary Figure 15 Molecular modeling of IgSF11-dependent homophilic adhesion.

Three possible molecular models of homophilic and trans-cellular adhesion mediated by the two Ig domains of IgSF11. These models were based on the X-ray crystal structure of CAR (a relative of IgSF11) 49.

Supplementary Figure 16 Full-length images of immunoblots and gels.

(a) Full-length immunoblots for Fig. 1d. (b) Fig. 1e. (c) Fig. 1f. (d) Fig. 1g. (e) Fig. 1h. (f) Fig. 1i. (g) Fig. 2a. (h) Fig. 2b.

Supplementary Figure 17 Full-length images of immunoblots and gels.

(a) A full-length gel for Fig. 8a. (b) Full-length immunoblots for Fig. 8b. (c) Supplementary Fig. 5b. (d) Supplementary Fig. 6a. (e) Supplementary Fig. 6b. (f) Supplementary Fig. 6c. (g) Supplementary Fig. 6d. (h) A full-length gel for Supplementary Fig. 9b. (i) Full-length immunoblots for Supplementary Fig. 9c. (j) Supplementary Fig. 10d. (k) Supplementary Fig. 11a. The blackened immunoblot panels indicate overexposed films, which could be properly exposed in other immunoblot experiments. (l) A full-length immunoblot for Supplementary Fig. 14a.

Supplementary information

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Jang, S., Oh, D., Lee, Y. et al. Synaptic adhesion molecule IgSF11 regulates synaptic transmission and plasticity. Nat Neurosci 19, 84–93 (2016). https://doi.org/10.1038/nn.4176

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/nn.4176

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