PSEN1ΔE9, APPswe, and APOE4 Confer Disparate Phenotypes in Human iPSC-Derived Microglia

Henna Konttinen, Mauricio E Castro Cabral-da-Silva, Sohvi Ohtonen, Sara Wojciechowski, Anastasia Shakirzyanova, Simone Caligola, Rosalba Giugno, Yevheniia Ishchenko, Damián Hernández, Mohammad Feroze Fazaludeen, Shaila Eamen, Mireia Gómez Budia, Ilkka Fagerlund, Flavia Scoyni, Paula Korhonen, Nadine Huber, Annakaisa Haapasalo, Alex W Hewitt, James Vickers, Grady C Smith, Minna Oksanen, Caroline Graff, Katja M Kanninen, Sarka Lehtonen, Nicholas Propson, Michael P Schwartz, Alice Pébay, Jari Koistinaho, Lezanne Ooi, Tarja Malm, Henna Konttinen, Mauricio E Castro Cabral-da-Silva, Sohvi Ohtonen, Sara Wojciechowski, Anastasia Shakirzyanova, Simone Caligola, Rosalba Giugno, Yevheniia Ishchenko, Damián Hernández, Mohammad Feroze Fazaludeen, Shaila Eamen, Mireia Gómez Budia, Ilkka Fagerlund, Flavia Scoyni, Paula Korhonen, Nadine Huber, Annakaisa Haapasalo, Alex W Hewitt, James Vickers, Grady C Smith, Minna Oksanen, Caroline Graff, Katja M Kanninen, Sarka Lehtonen, Nicholas Propson, Michael P Schwartz, Alice Pébay, Jari Koistinaho, Lezanne Ooi, Tarja Malm

Abstract

Here we elucidate the effect of Alzheimer disease (AD)-predisposing genetic backgrounds, APOE4, PSEN1ΔE9, and APPswe, on functionality of human microglia-like cells (iMGLs). We present a physiologically relevant high-yield protocol for producing iMGLs from induced pluripotent stem cells. Differentiation is directed with small molecules through primitive erythromyeloid progenitors to re-create microglial ontogeny from yolk sac. The iMGLs express microglial signature genes and respond to ADP with intracellular Ca2+ release distinguishing them from macrophages. Using 16 iPSC lines from healthy donors, AD patients and isogenic controls, we reveal that the APOE4 genotype has a profound impact on several aspects of microglial functionality, whereas PSEN1ΔE9 and APPswe mutations trigger minor alterations. The APOE4 genotype impairs phagocytosis, migration, and metabolic activity of iMGLs but exacerbates their cytokine secretion. This indicates that APOE4 iMGLs are fundamentally unable to mount normal microglial functionality in AD.

Keywords: APOE; APPswe; Alzheimer disease; E9; PSEN1Δ; iPSC; metabolism; microglia; mitochondria; phagocytosis.

Copyright © 2019 The Author(s). Published by Elsevier Inc. All rights reserved.

Figures

Graphical abstract
Graphical abstract
Figure 1
Figure 1
iPSCs Differentiate into iMGLs through Primitive Hematopoiesis (A–E) Schematic protocol (A). Percentages of positive cells analyzed by flow cytometry for markers of (B) pluripotency, (C) EMPs and mesodermal brachyury (BRAC), (D) primitive EMPs, and (E) and mature microglia. n = 4 cell lines, repeated in 3 batches. (F) The expression of microglial signature genes in RNA sequencing (RNA-seq) data of D24 iMGLs as log2 CPM values. n = 3 batches, 4 cell lines. (G) Hierarchical clustering of RNA-seq data shows that our iMGLs cluster with published iMGLs and human microglia (MG), but are distinct from dendritic cells (DCs), monocytes (CD14M and CD16M), iPSCs, and hematopoietic progenitor cells (HPCs) (Abud et al., 2017). (H–J) Immunostainings of D24 iMGLs (H). Repeated with two batches for all cell lines. Images of iMGLs labeled with IBA1 (red) in (I) 3D-Matrigel co-culture with neurons and in (J) cerebral brain organoids. Repeated with two batches for 2–4 cell lines. Scale bars as μm. Data presented mean ± SEM. See also Figures S1 and S2; Tables S1 and S2.
Figure 2
Figure 2
iMGLs Express APP and PSEN1 Proteins, and PSEN1ΔE9 Mutation Leads to Expected Alterations in PSEN1 Endoproteolysis (A–L) Western blots for full-length (FL) and C-terminal fragment (CTF) of PSEN1 and APP proteins from 3 batches of control (CTRL) and APPswe (pAPP, spAPP) iMGLs (A). GAPDH and b-ACTIN as loading controls. Quantification of blots normalized to GAPDH for (B) PSEN1-FL, (C) PSEN1-CTF, and (D) APP protein. n = 3 batches. Respective western blots (E) and quantification (F–H) for PSEN1ΔE9 iMGLs (pPSEN, spPSEN) and their isogenic controls (pISO and spISO). n = 2–5 batches. Western blots (I) for APOE3 and APOE4 iMGLs and quantification (J–L) for the proteins. n = 3 batches. (M) Aβ 1-42 levels in cell culture medium after 48 h analyzed by ELISA. n = 2–5 batches for APP and PSEN; n = 3 wells for APOE repeated in three batches. Data presented mean ± SEM unpaired two-tailed t test,∗p < 0.05, ∗∗p < 0.01, ∗∗∗p < 0.001. p, presymptomatic; sp, symptomatic. See also Figure S3.
Figure 3
Figure 3
ATP and ADP Evoke Intracellular Calcium [Ca2+]i Transients in iMGLs (A) Example traces of [Ca2+]i transients following 100 μM ATP (left panel) and ADP (right panel) applications for 5 s (indicated by bars) in iMGLs loaded with the Ca2+ indicator Fluo-4 AM. (B) The ratio of maximum amplitudes normalized to amplitudes evoked by ionomycin that was applied in the end of experiment and used as inclusion criteria. n = 4 batches, each with 9–10 coverslips, altogether 3,994 CTRL, 3,015 pAPP, and 3,906 spAPP cells. (C–E) Percentages of ATP- and ADP-responsive cells in APPswe lines compared with control iMGLs (C). Ratio of maximum amplitudes (D) and percentages of responsive cells (E) obtained from isogenic and PSENΔE9 iMGLs. n = 4 batches, each with 9–12 coverslips, altogether 1,969 pISO, 2,355 spISO, 1,856 pPSEN, and 2,823 spPSEN cells. (F and G) Similar data for APOE3 and APOE4 iMGLs. n = 4 coverslips, altogether 482 APOE3 and 991 APOE4 cells, repeated in three batches. Data presented mean ± SEM unpaired two-tailed t test or one-way ANOVA followed by Bonferroni's post hoc test, ∗p < 0.05, ∗∗p < 0.01. CTRL, control; p, presymptomatic; sp, symptomatic; PSEN, PSEN1ΔE9; APP, APPswe; and ISO, isogenic control iMGLs.
Figure 4
Figure 4
Chemokinesis Is Accelerated in APPswe and PSEN1ΔE9 iMGLs but Decelerated in APOE4 iMGLs (A) Representative images of iMGLs in scratch wound migration assay at 0, 12, and 24 h time points. Scale bar 300 µm. (B) Wound densities measured for 25 h with vehicle (VEH), 100 μM ATP, 100 μM ADP, or 1-μM soluble sAβ treatments. (C–E) Wound densities at 24 h normalized to vehicle (C). Time curves for (D) control (CTRL) and APPswe (APP), and (E) APOE3 and APOE4 iMGLs. (F) Wound densities at 24 h normalized to control or isogenic (ISO) iMGLs. (G) A heatmap for increase (darker color) or decrease (lighter color) in wound density compared with vehicle. White asterisks indicate significance compared with vehicle and black asterisks to control genotype. (H) Time curves for wound density with 100 μM fractalkine (CX3CL1) treatment in APO3 iMGLs. (I) Corresponding wound density at 24 h normalized to vehicle for APOE iMGLs. Curve graphs show a representative experiment of three replicates, n = 3–5 wells. Boxplots and heatmap show normalized results from n = 3–5 replicate batches. Data presented mean ± SEM, unpaired two-tailed t test, ∗p < 0.05, ∗p < 0.01, ∗∗∗p < 0.001. p, presymptomatic; sp, symptomatic. See also Figure S4.
Figure 5
Figure 5
Phagocytosis Is Dampened in APOE4 iMGLs, but not in APPswe or PSEN1ΔE9 iMGLs (A and B) Representative images of phagocytosed green pHrodo Zymosan A bioparticles in iMGLs at 5 h. (C) Time curves for pHrodo fluorescence intensity in control (CTRL) and APPswe (APP) iMGLs normalized to cell amount. (D) Respective boxplots at 5 h normalized to control or isogenic (ISO) iMGLs. (E and F) Representative images of phagocytosed FITC Zymosan A bioparticles in iMGLs. (G) pHrodo time curves for APOE3 and APOE4 iMGLs. (H) Percentages of APOE iMGLs that internalized certain number of FITC particles per cell. n = 290–750 cells. (I–L) pHrodo intensity at 5 h, after 24 h pretreatment with 100 ng/mL LPS, 20 ng/mL IFN-γ, or LPS-IFN-γ, or with simultaneous treatment with 0.5 μM soluble sAβ or fibrillar fAβ, compared with vehicle (Veh) in APPswe (I), pPSEN (J), spPSEN (K), and APOE (L) iMGLs. (M) Representative image of phagocytosed green fluor-Aβ1-42 in iMGLs at 5 h. (N) Time curves for fluorescence intensity of fluor-Aβ in control and APPswe iMGLs. (O) Respective bar graphs at 5 h normalized to control iMGLs. (P) Representative image of iMGLs treated with fluor-Aβ and fAβ depicting enlarged vacuoles. Scale bars, 50 μm. Curve graphs show a representative experiment of 3 replicates, n = 4 wells. Boxplots and bar graphs show normalized results from n = 2–6 replicate batches. Data presented mean ± SEM unpaired two-tailed t test or two-way ANOVA with Bonferroni's post hoc test,∗p < 0.05, ∗∗p < 0.01, ∗∗∗p < 0.001. p, presymptomatic; sp, symptomatic; PSEN, PSEN1ΔE9 iMGLs. See also Figure S4.
Figure 6
Figure 6
Cytokine Release under Proinflammatory Conditions Is Aggravated in APOE4 iMGLs but Decreased in PSEN1ΔE9 and APPswe iMGLs (A) iMGLs secrete cytokines when stimulated for 24 h with LPS 100 ng/mL, IFN-γ 20 ng/mL, or their combination LPS-IFN-γ as measured from media by cytometric bead array assay. Representative graphs. n = 4 wells. (B) spAPP iMGLs released less TNF-α, and pAPP less TNF-α and MCP1, compared with control iMGLs in response to LPS-IFN-γ treatment. (C) PSEN1ΔE9 iMGLs released less IL-6, TNF-α, and RANTES compared with isogenic iMGLs. (D) In contrast, APOE4 iMGLs released aggregated amounts of TNF-α and IL-8 compared with APOE3. For (B–D) n = 3–6 batches, each with 4 wells. Data presented mean ± SEM unpaired two-tailed t test, ∗p < 0.05, ∗p < 0.01, ∗∗∗p < 0.001. See also Figure S4. CTRL, control; p, presymptomatic; sp, symptomatic; PSEN, PSEN1ΔE9; APP, APPswe; and ISO, isogenic control iMGLs.
Figure 7
Figure 7
Metabolism of iMGLs Is Altered under Pro- and Anti-inflammatory Stimuli and by APOE4 Genetic Background (A) Representative oxygen consumption rate (OCR) curves for iMGLs following 24 h vehicle (VEH), LPS, IL-4, IFN-γ, and LPS-IFN-γ treatments, all 20 ng/mL. n = 3–5 wells. (B–D) Corresponding extracellular acidification rate (ECAR) curves (B). Mitochondrial parameters calculated from (C) OCRs in (A and D) from ECARs in (B). (E–H) Heatmap indicating decrease (blue) or increase (red) in fold change of mitochondrial parameters of LPS-treated iMGLs compared with vehicle (E). White equals 1. n = 5 CTRL, n = 4 pAPP, n = 2 spAPP, n = 3 APOE3, and n = 2 APOE4 batches with 10 wells; n = 1 isogenic, and n = 3 PSEN1 batches with 4–5 wells. Representative OCR and ECAR curves for (F) control and APPswe, (G) isogenic and PSEN1ΔE9, and for (H) APOE4 and APOE3 iMGLs. n = 5–10 wells, repeated with three batches. (I) Mitochondrial parameters calculated from OCRs and ECARs in (H) ∗p < 0.05, ∗∗p < 0.01, ∗∗∗p < 0.001 compared with vehicle, #compared with LPS, †compared with IFN-γ, two-tailed unpaired t test. Olig, oligomycin; FCCP, carbonyl cyanide-4-(trifluoromethoxy)phenylhydrazone; R/A, rotenone and antimycin A, each 1 μM. CTRL, control; p, presymptomatic; sp, symptomatic; PSEN, PSEN1ΔE9; APP, APPswe; and ISO, isogenic control iMGLs.

References

    1. Abud E.M., Ramirez R.N., Martinez E.S., Healy L.M., Nguyen C.H.H., Newman S.A., Yeromin A.V., Scarfone V.M., Marsh S.E., Fimbres C. iPSC-derived human microglia-like cells to study neurological diseases. Neuron. 2017;94:278–293.e9.
    1. Bagyinszky E., Youn Y.C., An S.S., Kim S. The genetics of Alzheimer's disease. Clin. Interv. Aging. 2014;9:535–551.
    1. Balez R., Steiner N., Engel M., Munoz S.S., Lum J.S., Wu Y., Wang D., Vallotton P., Sachdev P., O'Connor M. Neuroprotective effects of apigenin against inflammation, neuronal excitability and apoptosis in an induced pluripotent stem cell model of Alzheimer's disease. Sci. Rep. 2016;6:31450.
    1. Banati R.B., Gehrmann J., Czech C., Monning U., Jones L.L., Konig G., Beyreuther K., Kreutzberg G.W. Early and rapid de novo synthesis of Alzheimer beta A4-amyloid precursor protein (APP) in activated microglia. Glia. 1993;9:199–210.
    1. Barroeta-Espar I., Weinstock L.D., Perez-Nievas B.G., Meltzer A.C., Siao Tick Chong M., Amaral A.C., Murray M.E., Moulder K.L., Morris J.C., Cairns N.J. Distinct cytokine profiles in human brains resilient to Alzheimer's pathology. Neurobiol. Dis. 2019;121:327–337.
    1. Bennett M.L., Bennett F.C., Liddelow S.A., Ajami B., Zamanian J.L., Fernhoff N.B., Mulinyawe S.B., Bohlen C.J., Adil A., Tucker A. New tools for studying microglia in the mouse and human CNS. Proc. Natl. Acad. Sci. U S A. 2016;113:E1738–E1746.
    1. Butovsky O., Jedrychowski M.P., Moore C.S., Cialic R., Lanser A.J., Gabriely G., Koeglsperger T., Dake B., Wu P.M., Doykan C.E. Identification of a unique TGF-beta-dependent molecular and functional signature in microglia. Nat. Neurosci. 2014;17:131–143.
    1. Caldeira C., Cunha C., Vaz A.R., Falcao A.S., Barateiro A., Seixas E., Fernandes A., Brites D. Key aging-associated alterations in primary microglia response to beta-amyloid stimulation. Front. Aging Neurosci. 2017;9:277.
    1. Colonna M., Butovsky O. Microglia function in the central nervous system during health and neurodegeneration. Annu. Rev. Immunol. 2017;35:441–468.
    1. Crombie D.E., Daniszewski M., Liang H.H., Kulkarni T., Li F., Lidgerwood G.E., Conquest A., Hernandez D., Hung S.S., Gill K.P. Development of a modular automated system for maintenance and differentiation of adherent human pluripotent stem cells. SLAS Discov. 2017;22:1016–1025.
    1. Crook R., Verkkoniemi A., Perez-Tur J., Mehta N., Baker M., Houlden H., Farrer M., Hutton M., Lincoln S., Hardy J. A variant of Alzheimer's disease with spastic paraparesis and unusual plaques due to deletion of exon 9 of presenilin 1. Nat. Med. 1998;4:452–455.
    1. De Simone R., Niturad C.E., De Nuccio C., Ajmone-Cat M.A., Visentin S., Minghetti L. TGF-beta and LPS modulate ADP-induced migration of microglial cells through P2Y1 and P2Y12 receptor expression. J. Neurochem. 2010;115:450–459.
    1. Douvaras P., Sun B., Wang M., Kruglikov I., Lallos G., Zimmer M., Terrenoire C., Zhang B., Gandy S., Schadt E. Directed differentiation of human pluripotent stem cells to Microglia. Stem Cell Reports. 2017;8:1516–1524.
    1. Engel M., Balez R., Munoz S.S., Cabral-da-Silva M.C., Stevens C.H., Bax M., Do-Ha D., Sidhu K., Sachdev P., Ooi L. Viral-free generation and characterization of a human induced pluripotent stem cell line from dermal fibroblasts. Stem Cell Res. 2018;32:135–138.
    1. Ghosh S., Castillo E., Frias E.S., Swanson R.A. Bioenergetic regulation of microglia. Glia. 2018;66:1200–1212.
    1. Ginhoux F., Greter M., Leboeuf M., Nandi S., See P., Gokhan S., Mehler M.F., Conway S.J., Ng L.G., Stanley E.R. Fate mapping analysis reveals that adult microglia derive from primitive macrophages. Science. 2010;330:841–845.
    1. Ginhoux F., Lim S., Hoeffel G., Low D., Huber T. Origin and differentiation of microglia. Front. Cell Neurosci. 2013;7:45.
    1. Haenseler W., Sansom S.N., Buchrieser J., Newey S.E., Moore C.S., Nicholls F.J., Chintawar S., Schnell C., Antel J.P., Allen N.D. A highly efficient human pluripotent stem cell Microglia model displays a neuronal-co-culture-specific expression profile and inflammatory response. Stem Cell Reports. 2017;8:1727–1742.
    1. Hoffmann A., Kann O., Ohlemeyer C., Hanisch U.K., Kettenmann H. Elevation of basal intracellular calcium as a central element in the activation of brain macrophages (microglia): suppression of receptor-evoked calcium signaling and control of release function. J. Neurosci. 2003;23:4410–4419.
    1. Holmqvist S., Lehtonen S., Chumarina M., Puttonen K.A., Azevedo C., Lebedeva O., Ruponen M., Oksanen M., Djelloul M., Collin A. Creation of a library of induced pluripotent stem cells from Parkinsonian patients. NPJ Parkinson's Dis. 2016;2:16009.
    1. Jayadev S., Case A., Alajajian B., Eastman A.J., Moller T., Garden G.A. Presenilin 2 influences miR146 level and activity in microglia. J. Neurochem. 2013;127:592–599.
    1. Kennedy M., D'Souza S.L., Lynch-Kattman M., Schwantz S., Keller G. Development of the hemangioblast defines the onset of hematopoiesis in human ES cell differentiation cultures. Blood. 2007;109:2679–2687.
    1. Kierdorf K., Erny D., Goldmann T., Sander V., Schulz C., Perdiguero E.G., Wieghofer P., Heinrich A., Riemke P., Holscher C. Microglia emerge from erythromyeloid precursors via Pu.1- and Irf8-dependent pathways. Nat. Neurosci. 2013;16:273–280.
    1. Koenigsknecht-Talboo J., Landreth G.E. Microglial phagocytosis induced by fibrillar beta-amyloid and IgGs are differentially regulated by proinflammatory cytokines. J. Neurosci. 2005;25:8240–8249.
    1. Krasemann S., Madore C., Cialic R., Baufeld C., Calcagno N., El Fatimy R., Beckers L., O'Loughlin E., Xu Y., Fanek Z. The TREM2-APOE pathway drives the transcriptional phenotype of dysfunctional microglia in neurodegenerative diseases. Immunity. 2017;47:566–581.e9.
    1. Lambert C., Ase A.R., Seguela P., Antel J.P. Distinct migratory and cytokine responses of human microglia and macrophages to ATP. Brain Behav. Immun. 2010;24:1241–1248.
    1. Lanzrein A.S., Johnston C.M., Perry V.H., Jobst K.A., King E.M., Smith A.D. Longitudinal study of inflammatory factors in serum, cerebrospinal fluid, and brain tissue in Alzheimer disease: interleukin-1beta, interleukin-6, interleukin-1 receptor antagonist, tumor necrosis factor-alpha, the soluble tumor necrosis factor receptors I and II, and alpha1-antichymotrypsin. Alzheimer Dis. Assoc. Disord. 1998;12:215–227.
    1. Lavin Y., Winter D., Blecher-Gonen R., David E., Keren-Shaul H., Merad M., Jung S., Amit I. Tissue-resident macrophage enhancer landscapes are shaped by the local microenvironment. Cell. 2014;159:1312–1326.
    1. Lee C.Y., Landreth G.E. The role of microglia in amyloid clearance from the AD brain. J. Neural Transm. (Vienna) 2010;117:949–960.
    1. Lin Y.T., Seo J., Gao F., Feldman H.M., Wen H.L., Penney J., Cam H.P., Gjoneska E., Raja W.K., Cheng J. APOE4 causes widespread molecular and cellular alterations associated with Alzheimer's disease phenotypes in human iPSC-derived brain cell types. Neuron. 2018;98:1141–1154.e7.
    1. Liu C.C., Liu C.C., Kanekiyo T., Xu H., Bu G. Apolipoprotein E and Alzheimer disease: risk, mechanisms and therapy. Nat. Rev. Neurol. 2013;9:106–118.
    1. Manocha G.D., Floden A.M., Rausch K., Kulas J.A., McGregor B.A., Rojanathammanee L., Puig K.R., Puig K.L., Karki S., Nichols M.R. APP regulates microglial phenotype in a mouse model of Alzheimer's disease. J. Neurosci. 2016;36:8471–8486.
    1. McCaughey T., Liang H.H., Chen C., Fenwick E., Rees G., Wong R.C., Vickers J.C., Summers M.J., MacGregor C., Craig J.E. An interactive multimedia approach to improving informed consent for induced pluripotent stem cell research. Cell Stem Cell. 2016;18:307–308.
    1. McQuade A., Coburn M., Tu C.H., Hasselmann J., Davtyan H., Blurton-Jones M. Development and validation of a simplified method to generate human microglia from pluripotent stem cells. Mol. Neurodegener. 2018;13:67.
    1. Muffat J., Li Y., Yuan B., Mitalipova M., Omer A., Corcoran S., Bakiasi G., Tsai L.H., Aubourg P., Ransohoff R.M. Efficient derivation of microglia-like cells from human pluripotent stem cells. Nat. Med. 2016;22:1358–1367.
    1. Mullan M., Crawford F., Axelman K., Houlden H., Lilius L., Winblad B., Lannfelt L. A pathogenic mutation for probable Alzheimer's disease in the APP gene at the N-terminus of beta-amyloid. Nat. Genet. 1992;1:345–347.
    1. Munoz S.S., Balez R., Castro Cabral-da-Silva M.E., Berg T., Engel M., Bax M., Do-Ha D., Stevens C.H., Greenough M., Bush A. Generation and characterization of human induced pluripotent stem cell lines from a familial Alzheimer's disease PSEN1 A246E patient and a non-demented family member bearing wild-type PSEN1. Stem Cell Res. 2018;31:227–230.
    1. Nadler Y., Alexandrovich A., Grigoriadis N., Hartmann T., Rao K.S., Shohami E., Stein R. Increased expression of the gamma-secretase components presenilin-1 and nicastrin in activated astrocytes and microglia following traumatic brain injury. Glia. 2008;56:552–567.
    1. Oberstein T.J., Spitzer P., Klafki H.W., Linning P., Neff F., Knolker H.J., Lewczuk P., Wiltfang J., Kornhuber J., Maler J.M. Astrocytes and microglia but not neurons preferentially generate N-terminally truncated Abeta peptides. Neurobiol. Dis. 2015;73:24–35.
    1. Okita K., Matsumura Y., Sato Y., Okada A., Morizane A., Okamoto S., Hong H., Nakagawa M., Tanabe K., Tezuka K. A more efficient method to generate integration-free human iPS cells. Nat. Methods. 2011;8:409–412.
    1. Oksanen M., Hyotylainen I., Voutilainen J., Puttonen K.A., Hamalainen R.H., Graff C., Lehtonen S., Koistinaho J. Generation of a human induced pluripotent stem cell line (LL008 1.4) from a familial Alzheimer's disease patient carrying a double KM670/671NL (Swedish) mutation in APP gene. Stem Cell Res. 2018;31:181–185.
    1. Oksanen M., Petersen A.J., Naumenko N., Puttonen K., Lehtonen S., Gubert Olive M., Shakirzyanova A., Leskela S., Sarajarvi T., Viitanen M. PSEN1 mutant iPSC-derived model reveals severe astrocyte pathology in Alzheimer's disease. Stem Cell Reports. 2017;9:1885–1897.
    1. Olah M., Patrick E., Villani A.C., Xu J., White C.C., Ryan K.J., Piehowski P., Kapasi A., Nejad P., Cimpean M. A transcriptomic atlas of aged human microglia. Nat. Commun. 2018;9:539.
    1. Ooi L., Sidhu K., Poljak A., Sutherland G., O'Connor M.D., Sachdev P., Munch G. Induced pluripotent stem cells as tools for disease modelling and drug discovery in Alzheimer's disease. J. Neural Transm. (Vienna) 2013;120:103–111.
    1. Orihuela R., McPherson C.A., Harry G.J. Microglial M1/M2 polarization and metabolic states. Br. J. Pharmacol. 2016;173:649–665.
    1. Pandya H., Shen M.J., Ichikawa D.M., Sedlock A.B., Choi Y., Johnson K.R., Kim G., Brown M.A., Elkahloun A.G., Maric D. Differentiation of human and murine induced pluripotent stem cells to microglia-like cells. Nat. Neurosci. 2017;20:753–759.
    1. Pulido-Salgado M., Vidal-Taboada J.M., Barriga G.G., Sola C., Saura J. RNA-Seq transcriptomic profiling of primary murine microglia treated with LPS or LPS + IFNgamma. Sci. Rep. 2018;8:16096.
    1. Qu C., Puttonen K.A., Lindeberg H., Ruponen M., Hovatta O., Koistinaho J., Lammi M.J. Chondrogenic differentiation of human pluripotent stem cells in chondrocyte co-culture. Int. J. Biochem. Cell Biol. 2013;45:1802–1812.
    1. Rustenhoven J., Park T.I., Schweder P., Scotter J., Correia J., Smith A.M., Gibbons H.M., Oldfield R.L., Bergin P.S., Mee E.W. Isolation of highly enriched primary human microglia for functional studies. Sci. Rep. 2016;6:19371.
    1. Saijo K., Glass C.K. Microglial cell origin and phenotypes in health and disease. Nat. Rev. Immunol. 2011;11:775–787.
    1. Scheuner D., Eckman C., Jensen M., Song X., Citron M., Suzuki N., Bird T.D., Hardy J., Hutton M., Kukull W. Secreted amyloid beta-protein similar to that in the senile plaques of Alzheimer's disease is increased in vivo by the presenilin 1 and 2 and APP mutations linked to familial Alzheimer's disease. Nat. Med. 1996;2:864–870.
    1. Schwartz M.P., Hou Z., Propson N.E., Zhang J., Engstrom C.J., Santos Costa V., Jiang P., Nguyen B.K., Bolin J.M., Daly W. Human pluripotent stem cell-derived neural constructs for predicting neural toxicity. Proc. Natl. Acad. Sci. U S A. 2015;112:12516–12521.
    1. Selkoe D.J. The cell biology of beta-amyloid precursor protein and presenilin in Alzheimer's disease. Trends Cell Biol. 1998;8:447–453.
    1. Shi Y., Holtzman D.M. Interplay between innate immunity and Alzheimer disease: APOE and TREM2 in the spotlight. Nat. Rev. Immunol. 2018;18:759–772.
    1. Smith J.A., Das A., Ray S.K., Banik N.L. Role of pro-inflammatory cytokines released from microglia in neurodegenerative diseases. Brain Res. Bull. 2012;87:10–20.
    1. Sturgeon C.M., Ditadi A., Awong G., Kennedy M., Keller G. Wnt signaling controls the specification of definitive and primitive hematopoiesis from human pluripotent stem cells. Nat. Biotechnol. 2014;32:554–561.
    1. Ta T.T., Dikmen H.O., Schilling S., Chausse B., Lewen A., Hollnagel J.O., Kann O. Priming of microglia with IFN-gamma slows neuronal gamma oscillations in situ. Proc. Natl. Acad. Sci. U S A. 2019
    1. Townsend K.P., Town T., Mori T., Lue L.F., Shytle D., Sanberg P.R., Morgan D., Fernandez F., Flavell R.A., Tan J. CD40 signaling regulates innate and adaptive activation of microglia in response to amyloid beta-peptide. Eur. J. Immunol. 2005;35:901–910.
    1. Uenishi G., Theisen D., Lee J.H., Kumar A., Raymond M., Vodyanik M., Swanson S., Stewart R., Thomson J., Slukvin I. Tenascin C promotes hematoendothelial development and T lymphoid commitment from human pluripotent stem cells in chemically defined conditions. Stem Cell Reports. 2014;3:1073–1084.
    1. Ulland T.K., Song W.M., Huang S.C., Ulrich J.D., Sergushichev A., Beatty W.L., Loboda A.A., Zhou Y., Cairns N.J., Kambal A. TREM2 maintains microglial metabolic fitness in Alzheimer's disease. Cell. 2017;170:649–663.e13.
    1. Wang W.Y., Tan M.S., Yu J.T., Tan L. Role of pro-inflammatory cytokines released from microglia in Alzheimer's disease. Ann. Transl. Med. 2015;3:136.
    1. Weuve J., Hebert L.E., Scherr P.A., Evans D.A. Deaths in the United States among persons with Alzheimer's disease (2010-2050) Alzheimers Dement. 2014;10:e40–e46.
    1. Xu M., Zhang L., Liu G., Jiang N., Zhou W., Zhang Y. Pathological changes in alzheimer's disease analyzed using induced pluripotent stem cell-derived human Microglia-like cells. J. Alzheimers Dis. 2019;67:357–368.
    1. Zhang Y., Chen K., Sloan S.A., Bennett M.L., Scholze A.R., O'Keeffe S., Phatnani H.P., Guarnieri P., Caneda C., Ruderisch N. An RNA-sequencing transcriptome and splicing database of glia, neurons, and vascular cells of the cerebral cortex. J. Neurosci. 2014;34:11929–11947.
    1. Zhao Y., Li X., Huang T., Jiang L.L., Tan Z., Zhang M., Cheng I.H., Wang X., Bu G., Zhang Y.W. Intracellular trafficking of TREM2 is regulated by presenilin 1. Exp. Mol. Med. 2017;49:e405.

Source: PubMed

Подписаться