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:

The innate immunity protein IFITM3 modulates γ-secretase in Alzheimer’s disease

Abstract

Innate immunity is associated with Alzheimer’s disease1, but the influence of immune activation on the production of amyloid-β is unknown2,3. Here we identify interferon-induced transmembrane protein 3 (IFITM3) as a γ-secretase modulatory protein, and establish a mechanism by which inflammation affects the generation of amyloid-β. Inflammatory cytokines induce the expression of IFITM3 in neurons and astrocytes, which binds to γ-secretase and upregulates its activity, thereby increasing the production of amyloid-β. The expression of IFITM3 is increased with ageing and in mouse models that express familial Alzheimer’s disease genes. Furthermore, knockout of IFITM3 reduces γ-secretase activity and the formation of amyloid plaques in a transgenic mouse model (5xFAD) of early amyloid deposition. IFITM3 protein is upregulated in tissue samples from a subset of patients with late-onset Alzheimer’s disease that exhibit higher γ-secretase activity. The amount of IFITM3 in the γ-secretase complex has a strong and positive correlation with γ-secretase activity in samples from patients with late-onset Alzheimer’s disease. These findings reveal a mechanism in which γ-secretase is modulated by neuroinflammation via IFITM3 and the risk of Alzheimer’s disease is thereby increased.

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: IFITM3 directly binds to γ-secretase.
Fig. 2: Effect of IFITM3 on γ-secretase activity for APP cleavage.
Fig. 3: Effect of ageing and the expression of APP or PS1 on IFITM3 and γ-secretase.
Fig. 4: The association of IFITM3 with the γ-secretase complex in human brains.
Fig. 5: IFITM3 is located near the active site of γ-secretase.

Similar content being viewed by others

Data availability

Source data are provided with this paper. All other data are available from the corresponding authors upon reasonable request. Source data are provided with this paper.

References

  1. Shi, Y. & Holtzman, D. M. Interplay between innate immunity and Alzheimer disease: APOE and TREM2 in the spotlight. Nat. Rev. Immunol. 18, 759–772 (2018).

    CAS  PubMed  PubMed Central  Google Scholar 

  2. Griciuc, A. et al. Alzheimer’s disease risk gene CD33 inhibits microglial uptake of amyloid beta. Neuron 78, 631–643 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  3. Wang, Y. et al. TREM2 lipid sensing sustains the microglial response in an Alzheimer’s disease model. Cell 160, 1061–1071 (2015).

    CAS  PubMed  PubMed Central  Google Scholar 

  4. De Strooper, B. Aph-1, Pen-2, and Nicastrin with Presenilin generate an active γ-secretase complex. Neuron 38, 9–12 (2003).

    PubMed  Google Scholar 

  5. Gertsik, N., Chiu, D. & Li, Y. M. Complex regulation of γ-secretase: from obligatory to modulatory subunits. Front. Aging Neurosci. 6, 342 (2015).

    PubMed  PubMed Central  Google Scholar 

  6. Villa, J. C. et al. Nontranscriptional role of Hif-1α in activation of γ-secretase and notch signaling in breast cancer. Cell Rep. 8, 1077–1092 (2014).

    CAS  PubMed  PubMed Central  Google Scholar 

  7. Crump, C. J., Johnson, D. S. & Li, Y.-M. Development and mechanism of γ-secretase modulators for Alzheimer’s disease. Biochemistry 52, 3197–3216 (2013).

    CAS  PubMed  Google Scholar 

  8. Wagner, S. L. et al. Soluble γ-secretase modulators selectively inhibit the production of the 42-amino acid amyloid β peptide variant and augment the production of multiple carboxy-truncated amyloid β species. Biochemistry 53, 702–713 (2014).

    CAS  PubMed  PubMed Central  Google Scholar 

  9. Wagner, S. L. et al. Pharmacological and toxicological properties of the potent oral γ-secretase modulator BPN-15606. J. Pharmacol. Exp. Ther. 362, 31–44 (2017).

    CAS  PubMed  PubMed Central  Google Scholar 

  10. Kounnas, M. Z. et al. Modulation of γ-secretase reduces β-amyloid deposition in a transgenic mouse model of Alzheimer’s disease. Neuron 67, 769–780 (2010).

    CAS  PubMed  PubMed Central  Google Scholar 

  11. Pozdnyakov, N. et al. γ-Secretase modulator (GSM) photoaffinity probes reveal distinct allosteric binding sites on presenilin. J. Biol. Chem. 288, 9710–9720 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  12. Bailey, C. C., Zhong, G., Huang, I. C. & Farzan, M. IFITM-family proteins: the cell’s first line of antiviral defense. Annu. Rev. Virol. 1, 261–283 (2014).

    PubMed  PubMed Central  Google Scholar 

  13. Kumar, D. K. et al. Amyloid-β peptide protects against microbial infection in mouse and worm models of Alzheimer’s disease. Sci. Transl. Med. 8, 340ra72 (2016).

    PubMed  Google Scholar 

  14. Eimer, W. A. et al. Alzheimer’s disease-associated β-amyloid is rapidly seeded by Herpesviridae to protect against brain infection. Neuron 100, 1527–1532 (2018).

    CAS  PubMed  Google Scholar 

  15. Oakley, H. et al. Intraneuronal β-amyloid aggregates, neurodegeneration, and neuron loss in transgenic mice with five familial Alzheimer’s disease mutations: potential factors in amyloid plaque formation. J. Neurosci. 26, 10129–10140 (2006).

    CAS  PubMed  PubMed Central  Google Scholar 

  16. Allen, M. et al. Human whole genome genotype and transcriptome data for Alzheimer’s and other neurodegenerative diseases. Sci. Data 3, 160089 (2016).

    CAS  PubMed  PubMed Central  Google Scholar 

  17. GTEx Consortium. Human genomics. The Genotype-Tissue Expression (GTEx) pilot analysis: multitissue gene regulation in humans. Science 348, 648–660 (2015).

    Google Scholar 

  18. Wang, M. et al. The Mount Sinai cohort of large-scale genomic, transcriptomic and proteomic data in Alzheimer’s disease. Sci. Data 5, 180185 (2018).

    CAS  PubMed  PubMed Central  Google Scholar 

  19. Allen, E. K. et al. SNP-mediated disruption of CTCF binding at the IFITM3 promoter is associated with risk of severe influenza in humans. Nat. Med. 23, 975–983 (2017).

    CAS  PubMed  PubMed Central  Google Scholar 

  20. Frost, G. R. & Li, Y. M. The role of astrocytes in amyloid production and Alzheimer’s disease. Open Biol. 7, 170228 (2017).

    PubMed  PubMed Central  Google Scholar 

  21. Zhang, B. et al. Integrated systems approach identifies genetic nodes and networks in late-onset Alzheimer’s disease. Cell 153, 707–720 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  22. Imamura, Y. et al. Inhibition of gamma-secretase activity by helical beta-peptide foldamers. J. Am. Chem. Soc. 131, 7353–7359 (2009).

    CAS  PubMed  Google Scholar 

  23. Wakabayashi, T. et al. Analysis of the γ-secretase interactome and validation of its association with tetraspanin-enriched microdomains. Nat. Cell Biol. 11, 1340–1346 (2009).

    CAS  PubMed  Google Scholar 

  24. Jonsson, T. et al. Variant of TREM2 associated with the risk of Alzheimer’s disease. N. Engl. J. Med. 368, 107–116 (2013).

    CAS  PubMed  Google Scholar 

  25. Guerreiro, R. et al. TREM2 variants in Alzheimer’s disease. N. Engl. J. Med. 368, 117–127 (2013).

    CAS  PubMed  Google Scholar 

  26. Bertram, L. et al. Genome-wide association analysis reveals putative Alzheimer’s disease susceptibility loci in addition to APOE. Am. J. Hum. Genet. 83, 623–632 (2008).

    CAS  PubMed  PubMed Central  Google Scholar 

  27. Baruch, K. et al. Aging-induced type I interferon response at the choroid plexus negatively affects brain function. Science 346, 89–93 (2014).

    ADS  CAS  PubMed  PubMed Central  Google Scholar 

  28. Li, Y. M. et al. Photoactivated gamma-secretase inhibitors directed to the active site covalently label presenilin 1. Nature 405, 689–694 (2000).

    ADS  CAS  PubMed  Google Scholar 

  29. Chun, J., Yin, Y. I., Yang, G., Tarassishin, L. & Li, Y. M. Stereoselective synthesis of photoreactive peptidomimetic gamma-secretase inhibitors. J. Org. Chem. 69, 7344–7347 (2004).

    CAS  PubMed  Google Scholar 

  30. Gertsik, N., Chau, D. M. & Li, Y. M. γ-secretase inhibitors and modulators induce distinct conformational changes in the active sites of γ-secretase and signal peptide peptidase. ACS Chem. Biol. 10, 1925–1931 (2015).

    CAS  PubMed  PubMed Central  Google Scholar 

  31. Crump, C. J. et al. Development of sulfonamide photoaffinity inhibitors for probing cellular γ-secretase. ACS Chem. Neurosci. 7, 1166–1173 (2016).

    CAS  PubMed  PubMed Central  Google Scholar 

  32. Crump, C. J. et al. Piperidine acetic acid based γ-secretase modulators directly bind to Presenilin-1. ACS Chem. Neurosci. 2, 705–710 (2011).

    CAS  PubMed  PubMed Central  Google Scholar 

  33. Crump, C. J. et al. BMS-708,163 targets presenilin and lacks notch-sparing activity. Biochemistry 51, 7209–7211 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  34. Yang, G. et al. Stereo-controlled synthesis of novel photoreactive gamma-secretase inhibitors. Bioorg. Med. Chem. Lett. 19, 922–925 (2009).

    CAS  PubMed  Google Scholar 

  35. Xu, M. et al. γ -Secretase: characterization and implication for Alzheimer disease therapy. Neurobiol. Aging 23, 1023–1030 (2002).

    CAS  PubMed  Google Scholar 

  36. Pettersson, M. et al. Discovery of indole-derived pyridopyrazine-1,6-dione γ-secretase modulators that target presenilin. Bioorg. Med. Chem. Lett. 25, 908–913 (2015).

    CAS  PubMed  Google Scholar 

  37. Sung, J. Y. et al. WAVE1 controls neuronal activity-induced mitochondrial distribution in dendritic spines. Proc. Natl Acad. Sci. USA 105, 3112–3116 (2008).

    ADS  CAS  PubMed  Google Scholar 

  38. Ho, S. M. et al. Rapid Ngn2-induction of excitatory neurons from hiPSC-derived neural progenitor cells. Methods 101, 113–124 (2016).

    CAS  PubMed  Google Scholar 

  39. Stanley, S. et al. Profiling of glucose-sensing neurons reveals that GHRH neurons are activated by hypoglycemia. Cell Metab. 18, 596–607 (2013).

    CAS  PubMed  Google Scholar 

  40. Heiman, M., Kulicke, R., Fenster, R. J., Greengard, P. & Heintz, N. Cell type-specific mRNA purification by translating ribosome affinity purification (TRAP). Nat. Protoc. 9, 1282–1291 (2014).

    CAS  PubMed  Google Scholar 

  41. R Core Team. A Language and Environment for Statistical Computing, R Foundation for Statistical Computing; https://www.R-project.org/ (Vienna, Austria 2017).

  42. Chau, D. M., Crump, C. J., Villa, J. C., Scheinberg, D. A. & Li, Y. M. Familial Alzheimer disease presenilin-1 mutations alter the active site conformation of γ-secretase. J. Biol. Chem. 287, 17288–17296 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  43. Shelton, C. C. et al. Modulation of γ-secretase specificity using small molecule allosteric inhibitors. Proc. Natl Acad. Sci. USA 106, 20228–20233 (2009).

    ADS  CAS  PubMed  Google Scholar 

  44. Wong, E. et al. GSAP modulates γ-secretase specificity by inducing conformational change in PS1. Proc. Natl Acad. Sci. USA 116, 6385–6390 (2019).

    CAS  PubMed  Google Scholar 

  45. Hur, J. Y. et al. Active γ-secretase is localized to detergent-resistant membranes in human brain. FEBS J. 275, 1174–1187 (2008).

    CAS  PubMed  Google Scholar 

  46. Placanica, L., Zhu, L. & Li, Y. M. Gender- and age-dependent gamma-secretase activity in mouse brain and its implication in sporadic Alzheimer disease. PLoS ONE 4, e5088 (2009).

    ADS  PubMed  PubMed Central  Google Scholar 

  47. Placanica, L. et al. Pen2 and presenilin-1 modulate the dynamic equilibrium of presenilin-1 and presenilin-2 γ-secretase complexes. J. Biol. Chem. 284, 2967–2977 (2009).

    CAS  PubMed  PubMed Central  Google Scholar 

  48. Li, Y. M. et al. Presenilin 1 is linked with γ-secretase activity in the detergent solubilized state. Proc. Natl Acad. Sci. USA 97, 6138–6143 (2000).

    ADS  CAS  PubMed  Google Scholar 

  49. Frykman, S. et al. Synaptic and endosomal localization of active gamma-secretase in rat brain. PLoS ONE 5, e8948 (2010).

    ADS  PubMed  PubMed Central  Google Scholar 

  50. Hur, J. Y. et al. Identification of novel γ-secretase-associated proteins in detergent-resistant membranes from brain. J. Biol. Chem. 287, 11991–12005 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  51. Ahn, K. et al. Activation and intrinsic γ-secretase activity of presenilin 1. Proc. Natl Acad. Sci. USA 107, 21435–21440 (2010).

    ADS  CAS  PubMed  Google Scholar 

  52. Shelton, C. C., Tian, Y., Frattini, M. G. & Li, Y. M. An exo-cell assay for examining real-time gamma-secretase activity and inhibition. Mol. Neurodegener. 4, 22 (2009).

    PubMed  PubMed Central  Google Scholar 

  53. Placanica, L. et al. Pen2 and presenilin-1 modulate the dynamic equilibrium of presenilin-1 and presenilin-2 gamma-secretase complexes. J. Biol. Chem. 284, 2967–2977 (2009).

    CAS  PubMed  PubMed Central  Google Scholar 

  54. Tian, Y., Bassit, B., Chau, D. & Li, Y. M. An APP inhibitory domain containing the Flemish mutation residue modulates gamma-secretase activity for Aβ production. Nat. Struct. Mol. Biol. 17, 151–158 (2010).

    CAS  PubMed  Google Scholar 

  55. Tian, Y., Crump, C. J. & Li, Y. M. Dual role of α-secretase cleavage in the regulation of γ-secretase activity for amyloid production. J. Biol. Chem. 285, 32549–32556 (2010).

    CAS  PubMed  PubMed Central  Google Scholar 

  56. Muller, P. Y., Janovjak, H., Miserez, A. R. & Dobbie, Z. Processing of gene expression data generated by quantitative real-time RT-PCR. Biotechniques 32, 1372–1374, 1376, 1378–1379 (2002).

    CAS  PubMed  Google Scholar 

  57. Wang, J. C. et al. Risk for nicotine dependence and lung cancer is conferred by mRNA expression levels and amino acid change in CHRNA5. Hum. Mol. Genet. 18, 3125–3135 (2009).

    CAS  PubMed  PubMed Central  Google Scholar 

  58. Wang, J. C. et al. Genetic variation in the CHRNA5 gene affects mRNA levels and is associated with risk for alcohol dependence. Mol. Psychiatry 14, 501–510 (2009).

    CAS  PubMed  Google Scholar 

  59. Smyth, G. K. Linear models and empirical bayes methods for assessing differential expression in microarray experiments. Stat. Appl. Genet Mol. Biol. 3, (2004).

  60. Readhead, B. et al. Multiscale analysis of independent alzheimer’s cohorts finds disruption of molecular, genetic, and clinical networks by human herpesvirus. Neuron 99, 64–82 (2018).

    CAS  PubMed  PubMed Central  Google Scholar 

  61. Inoue, M. et al. Human brain proteins showing neuron-specific interactions with γ-secretase. FEBS J. 282, 2587–2599 (2015).

    CAS  PubMed  Google Scholar 

  62. Teranishi, Y. et al. Proton myo-inositol cotransporter is a novel γ-secretase associated protein that regulates Aβ production without affecting Notch cleavage. FEBS J. 282, 3438–3451 (2015).

    CAS  PubMed  Google Scholar 

Download references

Acknowledgements

We thank A. Kentsis and R. Hendrickson for LC–MS/MS and proteomic analysis. We thank H. Brogdon and N. L. Nadon for mouse brains. We thank E. Sikora for Apoe genotyping, D. Yarilin for immunostaining, S. Fujisawa and Y. Romin for help with microscopy, and S. Shuldberg for human brain samples. We thank S. Wagner for providing NGP-97555. This work is supported by the JPB Foundation (Y.-M.L., P.G. and A.G.), the Fisher Center for Alzheimer’s Research Foundation (P.G.), Cure Alzheimer’s Fund (Y.-M.L), the National Institutes of Health R01NS096275 (Y.-M.L.), RF1AG057593 (Y.-M.L.), R01AG061350 (Y.-M.L.), R01AG046170 (B.Z.), RF1AG057440 (B.Z.) and R01AG057907 (B.Z.). We also acknowledge the MSK Cancer Center Support Grant/Core Grant (grant P30 CA008748) the ADRC parent grant (P30 AG062429), Mr. William H. Goodwin and Mrs. Alice Goodwin and the Commonwealth Foundation for Cancer Research, the Experimental Therapeutics Center of MSKCC, and the William Randolph Hearst Fund in Experimental Therapeutics.

Author information

Authors and Affiliations

Authors

Contributions

J.-Y.H. and Y.-M.L. conceived the study. J.-Y.H., G.R.F. and Y.-M.L. planned, and J.-Y.H., G.R.F. and X.W. performed most of the experiments, and J.-Y.H., G.R.F. and Y.-M.L. analysed the data. X.W. and C.C. performed proteomics analysis. S.J.P. and E.W. cloned the constructs and S.J.P. expressed a recombinant substrate protein. E.W. performed a photolabelling study in primary neurons and Notch AlphaLISA assay in knockout cells. M.B. performed immnuofluorescence. T.L. perfused mouse brains and T.L., Y.Z. and Y.K. helped with primary neuronal culture. P.N. synthesized compounds. J.C.W. assisted with gene expression studies and analysed the large human brain dataset. J.T. cultured human iPS cell-derived neurons and astrocytes and performed immunofluorescence. L.G., A.M., C.M., X.Z., M.W. and B.Z. analysed gene expression in a large human brain dataset. Y.S. performed qPCR in mouse hippocampal cells and analysed the data. A.E.R. and B.T.E. performed SNP genotyping and analysed the data. K.R.S. and R.V. coordinated and provided 5xFAD transgenic tissue samples. I.T., R.A.R. and E.M. coordinated and provided the collection of human brain samples. G.R.F. summarized findings in illustrations. J.-Y.H., J.C.W., R.V., B.Z., D.S.J., E.M., P.G., A.G. and Y.-M.L. analysed the data. J.-Y.H. wrote the initial draft of the manuscript and, J.-Y.H., G.R.F. and Y.-M.L. revised the manuscript. All authors discussed and commented on the manuscript.

Corresponding author

Correspondence to Yue-Ming Li.

Ethics declarations

Competing interests

Y.-M.L. is co-inventor of intellectual property (assay for gamma secretase activity and screening method for gamma secretase inhibitors) owned by MSKCC and licensed to Jiangsu Continental Medical Development.

Additional information

Peer review information Nature thanks Paul Kellam, Rudolph Tanzi and the other, anonymous, reviewer(s) for their contribution to the peer review of this work.

Extended data figures and tables

Extended Data Fig. 1 Identification of IFITM3 as γ-secretase binding protein.

a, LC–MS/MS analysis of the 15-kDa band identified four peptides that match with human IFITM3. b, Western blotting analysis of PS1-NTF protein labelled with E2012-BPyne (500 nM). c, Structures of imidazole GSMs, acid GSM and GSIs. d, Western blotting analysis of E2012-BPyne-labelled proteins in the absence or presence of imidazole GSMs, acid GSM and GSIs. Labelled proteins were captured and analysed by western blotting for IFITM3. e, Structures of GY6 and 163-BP-l-biotin. f, IFITM3 co-immunoprecipitates with γ-secretase subunits. CHAPSO-solubilized cell membranes were immunoprecipitated with anti-IFITM3 antibody and probed with antibodies against PS2-CTF and PEN-2. Rabbit IgG was used as a negative control. g, IFITM3 does not co-immunoprecipitate with SPP. CHAPSO-solubilized cell membranes were immunoprecipitated with a monoclonal anti-IFITM3 antibody (9D11) and probed with antibodies against SPP and IFITM3. Mouse IgG was used as a negative control. h, Analysis of the total protein level in wild-type MEF or PS1/PS2 double knockout MEF cells. The same amount of membrane proteins was loaded and analysed by western blotting. i, Ifitm3 mRNA expression levels were measured by RT–PCR in wild-type MEFs or PS1/PS2 double knockout MEF cells (n = 6). All western blotting images and graphs are representative of three independent experiments (except in a and d, which represent two replicates). Data are mean ± s.d. ns, not significant. P values were determined by two-sided Student’s t-test.

Source data

Extended Data Fig. 2 Effect of IFITM3 knockdown and knockout on γ-secretase.

a, Quantification of western blotting (Fig. 2a) showed that IFITM3 knockdown did not change protein expression levels of APP, NCT and PS1-NTF in HEK-APP wild-type cells (n = 3). b, Schematic representation of cell-free γ-secretase assay. γ-secretase is incubated with a recombinant APP substrate in the presence of 0.25% CHAPSO. Cleaved Aβ40 and Aβ42 species are measured with cleavage specific antibodies and AlphaLISA technology. c, Schematic model showing different GSM and GSI binding sites in γ-secretase: E2012 (imidazole GSM), GSM-1 (acid GSM), and L458 (transition state analogue inhibitor, GSI). df, Comparison of IC50 values of GSM-25 (EV: n = 9, KO: n = 8) for Aβ40 (****P < 0.0001) and Aβ42 (ns) (d), GSM-1 (n = 6) for Aβ40 (**P = 0.0039) and Aβ42 (ns) (e), and L458 (n = 3) for Aβ40 (ns) and Aβ42 (ns) (f) cleavages in the U138 EV or KO cell lines (n ≥ 3). g, IFITM3 knockdown (KD) does not affect expression of γ-secretase subunits. IFITM3 was knocked down by siRNA (6 pmol, n = 3) in HEK-Notch∆E cells and scramble siRNA (SC, n = 3) was used as a negative control. Cell lysates were probed by antibodies against NCT, PS1-NTF and IFITM3. β-Actin was used as a loading control. h, Effect of IFITM3 knockdown on γ-secretase activity. IFITM3 knockdown increased γ-secretase cleaved product NICD, analysed by western blotting. Cell lysates were probed by antibodies against c-Myc (NotchΔE) and NICD and a representative quantification of NICD (n = 8, ***P = 0.001) is shown (bottom). i, Cell-based NICD AlphaLISA assay (left) revealed an increase in NICD production with IFITM3 knockdown. Quantification of NICD (n = 8, ***P = 0.001) is shown in the right panel. j, Effect of IFITM3 knockout on γ-secretase activity. Knockout cells lines have increased γ-secretase activity as compared to the EV cell line. The NICD cleavage in vitro was measured by AlphaLISA assay (n = 3, **P = 0.0096). All western blotting images and graphs are representative of three independent experiments. Data are mean ± s.d. P values were determined by two-sided Student’s t-test.

Source data

Extended Data Fig. 3 Effect of ageing and the expression of APP/PS1 on the level of γ-secretase and IFITM3.

a, Western blotting for NCT and PS1-NTF (Fig. 3a) were quantified by Odyssey imaging (n = 5 mice pooled per group, except n = 4 for 28F, ns). b, Effect of ageing on subcellular localizations of IFITM3. A hemibrain from male wild-type C57BL/6 mouse at 4 and 28 months (n = 1 per group) were homogenized and layered on iodixanol gradient (2.5–30%). Fractions were collected from the top and resolved by western blotting for γ-secretase, IFITM3 and different subcellular markers. c, Western blotting for APP, NCT and PS1-NTF (Fig. 3e) were quantified by Odyssey imaging (n = 5 mice per group except n = 4 for WT at 12 months). APP: 3moWT-3mo5x: ****P < 0.0001, 3mo5x-12mo5x: ****P < 0.0001, 12moWT-12mo5x: ****P < 0.0001). NCT: 3moWT-12moWT: ****P < 0.0001, 3mo5x-12mo5x: ****P < 0.0001, 12moWT-12mo5x: ****P < 0.0001. PS1-NTF: 3moWT-12mo5x: ***P = 0.0004, 12moWT-12mo5x: ****P < 0.0001. d, Immunostaining of IFITM3 in mouse brains. Fluorescence microscopy of IFITM3 expression in 12-month-old PFA-perfused mice (WT, top; 5xFAD, bottom). Representative images of cortex, hippocampus and subiculum (left to right) show IFITM3 (green) and DAPI (blue). Scale bars, 1 mm, 200 μm, 100 μm (left to right). Total IFITM3 fluorescence area within the hippocampus and cortex of WT and 5xFAD was quantified using FIJI. Total IFITM3 was divided by tissue area and IFITM3 expression was normalized to average of WT (WT: n = 7, 5xFAD: n = 9) (cortex: **P = 0.0035, hippocampus: ****P < 0.0001). e, IFITM3 expression in astrocytes and microglia is upregulated in 5xFAD mice compared to wild-type mice. Fluorescence microscopy of IFITM3, GFAP (top) and IBA1 (bottom) expression in 12-month-old PFA-perfused mice (WT, top; 5xFAD, bottom). Representative images of the hippocampus and cortex show IFITM3 (red), GFAP (green, top), IBA1 (green, bottom), and DAPI (blue). Scale bar, 500 μm. Inset panels (left to right) show GFAP or IBA1 (green), IFITM3 (red) and merged staining. Scale bar, 50 μm. All western blotting images and graphs are representative of three independent experiments (except in b, which denotes two replicates). Data mean ± s.d. P values were determined by two-sided Student’s t-test (except in c, which was by one-way ANOVA followed by Tukey).

Source data

Extended Data Fig. 4 Expression profile of IFITM3 and other markers in LOAD and age-matched controls.

a, Spearman’s correlation of mRNA expression of human IFITM3 gene with age was analysed in the cortex (n = 158) and hippocampus (n = 123) of normal human brains using the GTEx cohort. b, mRNA expression in samples from non-demented control participants (n = 10) and patients with LOAD (n = 18) of MAP2 (ns), GFAP (**P = 0.0046), and AIF1 (ns) were measured, which were used in Fig. 4c, d. c, Expression profiles of MAP2 (ns), GFAP (****P < 0.0001), and AIF1 (ns) in the temporal cortex of samples from human control participants (n = 76) and patients with LOAD (n = 80) using the Mayo Clinic cohort data. Correlation analyses were carried out and P values were calculated. d, The protein levels of NCT (****P < 0.0001) and PS1-NTF (**P = 0.0042) in human brain membranes (control and LOAD). The samples were analysed by western blotting and quantified (n = 10 and 18, respectively). Signal was normalized to HeLa cell membrane. e, f, IFITM3 SNP genotypes. e, Allelic discrimination plot depicting rs34481144 genotype calls for control (n = 9), LOAD-L (n = 10), and LOAD-H (n = 8) brain samples. The axes show delta Rn values obtained from TaqMan SNP genotyping analysis. Samples without genomic DNA were used as non-template controls (shown as black squares in the left lower quadrant, n = 2). f, Allele frequency of rs34481144 genotype in control (n = 9) and LOAD (n = 18). g, mRNA level of Ifitm3 in four types of eGFP/L10a-expressing mouse hippocampal neurons (GAD2 (glutamate decarboxylase 2), CCK (cholecystokinin), PV (parvalbumin), and CORT (cortistatin) expressing GABAergic neurons) (n = 4 per group, GAD2-PV: ***P = 0.0003, CCK-PV: ****P < 0.0001, CCK-Cort: *P = 0.0363, PV-Cort: **P = 0.0059). h, mRNA levels of IFITM3 in human iPS-cell-derived neurons (n = 4) and human primary astrocytes (n = 3) were measured by qPCR (****P < 0.0001). i, j, Human iPS-cell-derived neurons (i) and human primary astrocytes (j) were stained for IFITM3 with MAP2 (neuronal marker) or S100β (astrocyte marker). DAPI was used for nucleus staining. Scale bars, 200 μm and 500 μm. k, Induction of IFITM3 by IFNα in primary neurons. E16 mouse primary neurons were treated with 100 ng ml−1 of IFNα at DIV12 for 24 h. The protein levels of γ-secretase and IFITM3 were analysed by western blotting (n = 4 per group). β-tubulin III was used as a loading control. l, Effect of IFITM3 induction on γ-secretase activity for Aβ40 (*P = 0.0116) and Aβ42 (*P = 0.0319) activity. Membranes from primary neurons were incubated with the recombinant APP substrate C100-∆ID-FLAG and γ-secretase activity (Aβ cleavage rate) was assayed by human Aβ three-plex MSD kits (n = 12, 10). m, JC8 whole-cell photolabelling. Neuronal membranes were photolabelled with JC8 in the absence or presence of L458 and analysed by anti-PS1-NTF antibody. Photolabelled PS1-NTF protein level was quantified by Odyssey imaging (n = 3, *P = 0.0210). n, Spearman’s correlation between the expression level of IFITM3 and viruses. In the BA-22 region in the MSBB cohort, the expression level of IFITM3 is positively correlated with the expression level of HHV-6B (rho = 0.248, P = 0.044, n = 66). In the BA-36 region, the expression level of IFITM3 is positively correlated with the expression of hepatitis C virus genotype 4 (rho = 0.255, P = 0.033, n = 70). All western blotting images and graphs are representative of two independent experiments (except in b and e, which denote one replicate; in h, data are pooled from two experiments; in k, which denotes three replicates; and in l, data are pooled from four experiments). Data are mean ± s.d. Violin plots represent median (middle line) and interquartile range (outer lines). Two-sided Student’s t-test (except in g, which was by one-way ANOVA followed by Tukey).

Source data

Extended Data Fig. 5 Correlation between L505 labelled protein and γ-secretase activity.

a, Structures of JC8, L505, L646, GY4 and L631. b, Western blotting analysis of 11Bt-labelled proteins in the absence or presence of its parent compound, a substrate binding site inhibitor pep11 and imidazole GSM, E2012. Labelled proteins were captured and analysed by western blotting for PS1-NTF (n = 1). c, Pearson’s correlation between γ-secretase activity (Aβ40, Aβ42 cleavage rates in Fig. 4d) and L505-labelled PS1-NTF (in Fig. 5c) in LOAD samples (n = 10). Linear regression analysis was used to calculate R and P values.

Source data

Supplementary information

Supplementary Figure

Supplementary Figure 1: Gel Source Data.

Reporting Summary

Supplementary Table

Supplementary Table 1: List of human brain tissues used in Figure 4.

Supplementary Table

Supplementary Table 2: Spearman’s correlation between the expression level of IFITM3 and Aβ load.

Supplementary Table

Supplementary Table 3: Human IFITM3 SNP Genotypes.

Supplementary Table

Supplementary Table 4: Correlation between IFITM3 and cytokines in a large AD human dataset.

Supplementary Table

Supplementary Table 5: Correlation between IFITM3 and of type I IFN responsive genes in a large AD human dataset.

Source data

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Hur, JY., Frost, G.R., Wu, X. et al. The innate immunity protein IFITM3 modulates γ-secretase in Alzheimer’s disease. Nature 586, 735–740 (2020). https://doi.org/10.1038/s41586-020-2681-2

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41586-020-2681-2

This article is cited by

Comments

By submitting a comment you agree to abide by our Terms and Community Guidelines. If you find something abusive or that does not comply with our terms or guidelines please flag it as inappropriate.

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