Hostname: page-component-848d4c4894-x5gtn Total loading time: 0 Render date: 2024-06-02T10:19:46.071Z Has data issue: false hasContentIssue false

Modelling Alzheimer's disease using human brain organoids: current progress and challenges

Published online by Cambridge University Press:  15 December 2022

Mario Yanakiev
Affiliation:
Institute of Medical Sciences, School of Medicine, Medical Sciences and Nutrition, University of Aberdeen, Aberdeen AB25 2ZD, UK
Olivia Soper
Affiliation:
Institute of Medical Sciences, School of Medicine, Medical Sciences and Nutrition, University of Aberdeen, Aberdeen AB25 2ZD, UK
Daniel A. Berg
Affiliation:
Institute of Medical Sciences, School of Medicine, Medical Sciences and Nutrition, University of Aberdeen, Aberdeen AB25 2ZD, UK
Eunchai Kang*
Affiliation:
Institute of Medical Sciences, School of Medicine, Medical Sciences and Nutrition, University of Aberdeen, Aberdeen AB25 2ZD, UK
*
Author for correspondence: Eunchai Kang, E-mail: eunchai.kang@abdn.ac.uk

Abstract

Alzheimer's disease (AD) is a progressive neurodegenerative disorder characterised by gradual memory loss and declining cognitive and executive functions. AD is the most common cause of dementia, affecting more than 50 million people worldwide, and is a major health concern in society. Despite decades of research, the cause of AD is not well understood and there is no effective curative treatment so far. Therefore, there is an urgent need to increase understanding of AD pathophysiology in the hope of developing a much-needed cure. Dissecting the cellular and molecular mechanisms of AD pathogenesis has been challenging as the most commonly used model systems such as transgenic animals and two-dimensional neuronal culture do not fully recapitulate the pathological hallmarks of AD. The recent advent of three-dimensional human brain organoids confers unique opportunities to study AD in a humanised model system by encapsulating many aspects of AD pathology. In the present review, we summarise the studies of AD using human brain organoids that recapitulate the major pathological components of AD including amyloid-β and tau aggregation, neuroinflammation, mitochondrial dysfunction, oxidative stress and synaptic and circuitry dysregulation. Additionally, the current challenges and future directions of the brain organoids modelling system are discussed.

Type
Review
Copyright
Copyright © The Author(s), 2022. Published by Cambridge University Press

Access options

Get access to the full version of this content by using one of the access options below. (Log in options will check for institutional or personal access. Content may require purchase if you do not have access.)

Footnotes

*

These authors contributed equally.

References

Nichols, E et al. (2022) Estimation of the global prevalence of dementia in 2019 and forecasted prevalence in 2050: an analysis for the global burden of disease study 2019. The Lancet Public Health 7, 105125.CrossRefGoogle Scholar
Wimo, A et al. (2017) The worldwide costs of dementia 2015 and comparisons with 2010. Alzheimer's & Dementia 13, 17.CrossRefGoogle ScholarPubMed
Luengo-Fernandez, R, Leal, J and Gray, A (2015) UK research spend in 2008 and 2012: comparing stroke, cancer, coronary heart disease and dementia. BMJ Open 5, e006648.CrossRefGoogle ScholarPubMed
DeTure, MA and Dickson, DW (2019) The neuropathological diagnosis of Alzheimer's disease. Molecular Neurodegeneration 14, 118.CrossRefGoogle ScholarPubMed
Friedrich, RP et al. (2010) Mechanism of amyloid plaque formation suggests an intracellular basis of Aβ pathogenicity. Proceedings of the National Academy of Sciences of the United States of America 107, 19421947.CrossRefGoogle ScholarPubMed
Reitz, C (2012) Alzheimer's disease and the amyloid cascade hypothesis: a critical review. International Journal of Alzheimer's Disease 2012, article ID 369808.CrossRefGoogle ScholarPubMed
Mucke, L et al. (2000) High-level neuronal expression of Aβ1–42 in wild-type human amyloid protein precursor transgenic mice: synaptotoxicity without plaque formation. The Journal of Neuroscience 20, 40504058.CrossRefGoogle Scholar
Kitazawa, M, Medeiros, R and LaFerla, FM (2012) Transgenic mouse models of Alzheimer disease: developing a better model as a tool for therapeutic interventions. Current Pharmaceutical Design 18, 11311147.CrossRefGoogle ScholarPubMed
Bloom, GS (2014) Amyloid-β and tau: the trigger and bullet in Alzheimer disease pathogenesis. JAMA Neurology 71, 505508.CrossRefGoogle ScholarPubMed
Hadjichrysanthou, C et al. (2020) The dynamics of biomarkers across the clinical spectrum of Alzheimer's disease. Alzheimer's Research & Therapy 12, 116.CrossRefGoogle ScholarPubMed
Cline, EN et al. (2018) The amyloid-β oligomer hypothesis: beginning of the third decade. Journal of Alzheimer's Disease 64, 567610.CrossRefGoogle ScholarPubMed
Resnick, SM et al. (2010) Longitudinal cognitive decline is associated with fibrillar amyloid-beta measured by [11C]PiB. Neurology 74, 807815.CrossRefGoogle ScholarPubMed
Shepherd, C et al. (2009) Variations in the neuropathology of familial Alzheimer's disease. Acta Neuropathologica 118, 3752.CrossRefGoogle ScholarPubMed
Janssen, JC et al. (2003) Early onset familial Alzheimer's disease. Neurology 60, 235239.CrossRefGoogle ScholarPubMed
Crous-Bou, M et al. (2017) Alzheimer's disease prevention: from risk factors to early intervention. Alzheimer's Research & Therapy 9, Article number: 71.CrossRefGoogle ScholarPubMed
Chen, C and Zissimopoulos, JM (2018) Racial and ethnic differences in trends in dementia prevalence and risk factors in the United States. Alzheimer's & Dementia: Translational Research & Clinical Interventions 4, 510.CrossRefGoogle ScholarPubMed
Yamazaki, Y et al. (2019) Apolipoprotein E and Alzheimer disease: pathobiology and targeting strategies. Nature Reviews Neurology 15, 501518.CrossRefGoogle ScholarPubMed
Efthymiou, AG and Goate, AM (2017) Late onset Alzheimer's disease genetics implicates microglial pathways in disease risk. Molecular Neurodegeneration 12, 112.CrossRefGoogle ScholarPubMed
Jonsson, T et al. (2013) Variant of TREM2 associated with the risk of Alzheimer's disease. The New England Journal of Medicine 368, 107116.CrossRefGoogle ScholarPubMed
Griciuc, A et al. (2013) Alzheimer's disease risk gene CD33 inhibits microglial uptake of amyloid beta. Neuron 78, 631643.CrossRefGoogle ScholarPubMed
Lambert, J-C et al. (2009) Genome-wide association study identifies variants at CLU and CR1 associated with Alzheimer's disease. Nature Genetics 41, 10941099.CrossRefGoogle ScholarPubMed
Tan, L et al. (2016) Effect of CLU genetic variants on cerebrospinal fluid and neuroimaging markers in healthy, mild cognitive impairment and Alzheimer's disease cohorts. Scientific Reports 6, Article number: 26027.Google ScholarPubMed
Sung, P-S et al. (2020) Neuroinflammation and neurogenesis in Alzheimer's disease and potential therapeutic approaches. International Journal of Molecular Sciences 21, 701721.CrossRefGoogle ScholarPubMed
Zhang, F and Jiang, L (2015) Neuroinflammation in Alzheimer's disease. Neuropsychiatric Disease and Treatment 11, 243.CrossRefGoogle ScholarPubMed
DiSabato, D, Quan, N and Godbout, JP (2016) Neuroinflammation: the devil is in the details. Journal of Neurochemistry 139, 136.CrossRefGoogle ScholarPubMed
Leng, F and Edison, P (2020) Neuroinflammation and microglial activation in Alzheimer disease: where do we go from here? Nature Reviews Neurology 17, 157172.CrossRefGoogle Scholar
Wang, Q et al. (2022) Baseline microglial activation correlates with brain amyloidosis and longitudinal cognitive decline in Alzheimer disease. Neurology – Neuroimmunology and Neuroinflammation 9, e1152.CrossRefGoogle ScholarPubMed
Pascoal, TA et al. (2021) Microglial activation and tau propagate jointly across Braak stages. Nature Medicine 27, 15921599.CrossRefGoogle ScholarPubMed
Liddelow, SA et al. (2017) Neurotoxic reactive astrocytes are induced by activated microglia HHS public access. Nature 10, 481487.CrossRefGoogle Scholar
González-Reyes, RE et al. (2017) Involvement of astrocytes in Alzheimer's disease from a neuroinflammatory and oxidative stress perspective. Frontiers in Molecular Neuroscience 10, 427.CrossRefGoogle ScholarPubMed
Swardfager, W et al. (2010) A meta-analysis of cytokines in Alzheimer's disease. Biological Psychiatry 68, 930941.CrossRefGoogle ScholarPubMed
Drummond, E and Wisniewski, T (2017) Alzheimer's disease: experimental models and reality. Acta Neuropathologica 133, 155.CrossRefGoogle Scholar
Elder, GA, Sosa, MAG and de Gasperi, R (2010) Transgenic mouse models of Alzheimer's disease. The Mount Sinai Journal of Medicine 77, 69.CrossRefGoogle ScholarPubMed
Gunawardena, S and Goldstein, L (2001) Disruption of axonal transport and neuronal viability by amyloid precursor protein mutations in drosophila. Neuron 32, 389401.CrossRefGoogle ScholarPubMed
Sasaguri, H et al. (2017) APP mouse models for Alzheimer's disease preclinical studies. The EMBO Journal 36, 24732487.CrossRefGoogle ScholarPubMed
Oddo, S et al. (2003) Triple-transgenic model of Alzheimer's disease with plaques and tangles: intracellular Abeta and synaptic dysfunction. Neuron 39, 409421.CrossRefGoogle ScholarPubMed
Hutton, M et al. (1998) Association of missense and 5′-splice-site mutations in tau with the inherited dementia FTDP-17. Nature 393, 702705.CrossRefGoogle ScholarPubMed
Smits, LM et al. (2019) Modeling Parkinson's disease in midbrain-like organoids. Parkinson's Disease 5, 18.Google ScholarPubMed
Sakaguchi, H et al. (2015) Generation of functional hippocampal neurons from self-organizing human embryonic stem cell-derived dorsomedial telencephalic tissue. Nature Communications 6, 111.CrossRefGoogle ScholarPubMed
Ballabio, C et al. (2020) Modeling medulloblastoma in vivo and with human cerebellar organoids. Nature Communications 11, 118.CrossRefGoogle ScholarPubMed
Penney, J, Ralvenius, WT and Tsai, L-H (2019) Modeling Alzheimer's disease with iPSC-derived brain cells. Molecular Psychiatry 25, 148167.CrossRefGoogle ScholarPubMed
Zhang, Y et al. (2013) Rapid single-step induction of functional neurons from human pluripotent stem cells. Neuron 78, 785798.CrossRefGoogle ScholarPubMed
Choi, SH et al. (2015) Recapitulating amyloid β and tau pathology in human neural cell culture models: clinical implications. US Neurology 11, 102.CrossRefGoogle ScholarPubMed
Slanzi, A et al. (2020) In vitro models of neurodegenerative diseases. Frontiers in Cell and Developmental Biology 8, 328.CrossRefGoogle ScholarPubMed
Lancaster, MA et al. (2013) Cerebral organoids model human brain development and microcephaly. Nature 501, 373379.CrossRefGoogle ScholarPubMed
Qian, X et al. (2016) Brain-region-specific organoids using mini-bioreactors for modeling ZIKV exposure. Cell 165, 12381254.CrossRefGoogle ScholarPubMed
Qian, X, Song, H and Ming, G-L (2019) Brain organoids: advances, applications and challenges. Development 146, dev166074.CrossRefGoogle Scholar
Lancaster, MA et al. (2017) Guided self-organization and cortical plate formation in human brain organoids. Nature Biotechnology 35, 659666.CrossRefGoogle ScholarPubMed
Kadoshima, T et al. (2013) Self-organization of axial polarity, inside-out layer pattern, and species-specific progenitor dynamics in human ES cell-derived neocortex. Proceedings of the National Academy of Sciences of the United States of America 110, 2028420289.CrossRefGoogle ScholarPubMed
Pavoni, S et al. (2018) Small-molecule induction of Aβ-42 peptide production in human cerebral organoids to model Alzheimer's disease associated phenotypes. PLoS ONE 13, e0209150.CrossRefGoogle ScholarPubMed
Bagley, JA et al. (2017) Fused cerebral organoids model interactions between brain regions. Nature Methods 14, 743751.CrossRefGoogle ScholarPubMed
Nascimento, JM et al. (2019) Human cerebral organoids and fetal brain tissue share proteomic similarities. Frontiers in Cell and Developmental Biology 7, 303.CrossRefGoogle ScholarPubMed
Oliveira, B, Çerağ Yahya, A and Novarino, G (2019) Modeling cell–cell interactions in the brain using cerebral organoids. Brain Research 1724, 146458.CrossRefGoogle ScholarPubMed
Abud, EM et al. (2018) iPSC-derived human microglia-like cells to study neurological diseases. 94, 278293.Google Scholar
Ormel, PR et al. (2018) Microglia innately develop within cerebral organoids. Nature Communications 9, Article number: 4167.CrossRefGoogle ScholarPubMed
Choi, SH et al. (2014) A three-dimensional human neural cell culture model of Alzheimer's disease. Nature 515, 274.CrossRefGoogle ScholarPubMed
Chen, G et al. (2017) Amyloid beta: structure, biology and structure-based therapeutic development. Acta Pharmacologica Sinica 38, 12051235.CrossRefGoogle ScholarPubMed
Gonzalez, C et al. (2018) Modeling amyloid beta and tau pathology in human cerebral organoids. Molecular Psychiatry 23, 23632374.CrossRefGoogle ScholarPubMed
Raja, WK et al. (2016) Self-organizing 3D human neural tissue derived from induced pluripotent stem cells recapitulate Alzheimer's disease phenotypes. PLoS ONE 11, e0161969.CrossRefGoogle ScholarPubMed
Alić, I et al. (2020) Patient-specific Alzheimer-like pathology in trisomy 21 cerebral organoids reveals BACE2 as a gene dose-sensitive AD suppressor in human brain. Molecular Psychiatry 26, 5766–5788.Google ScholarPubMed
Kwak, SS et al. (2020) Amyloid-β42/40 ratio drives tau pathology in 3D human neural cell culture models of Alzheimer's disease. Nature Communications 11, Article number: 1377.CrossRefGoogle ScholarPubMed
Pomeshchik, Y et al. (2020) Human iPSC-derived hippocampal spheroids: an innovative tool for stratifying Alzheimer disease patient-specific cellular phenotypes and developing therapies. Stem Cell Reports 15, 256273.CrossRefGoogle ScholarPubMed
Arber, C et al. (2021) Familial Alzheimer's disease mutations in PSEN1 lead to premature human stem cell neurogenesis. Cell Reports 34, 108615.CrossRefGoogle ScholarPubMed
Quartey, MO et al. (2021) The Aβ(1–38) peptide is a negative regulator of the Aβ(1–42) peptide implicated in Alzheimer disease progression. Scientific Reports 11, 117.CrossRefGoogle ScholarPubMed
Holland, AJ et al. (1998) Population-based study of the prevalence and presentation of dementia in adults with down's syndrome. The British Journal of Psychiatry 172, 493498.CrossRefGoogle ScholarPubMed
Yan, Y et al. (2018) Modeling neurodegenerative microenvironment using cortical organoids derived from human stem cells. Tissue Engineering 24, 11251137.CrossRefGoogle ScholarPubMed
Lin, Y-T et al. (2018) APOE4 causes widespread molecular and cellular alterations associated with Alzheimer's disease phenotypes in human iPSC-derived brain cell types. Neuron 98, 11411154.CrossRefGoogle ScholarPubMed
Nixon, RA (2017) Amyloid precursor protein and endosomal–lysosomal dysfunction in Alzheimer's disease: inseparable partners in a multifactorial disease. The FASEB Journal 31, 27292743.CrossRefGoogle ScholarPubMed
Zhao, J et al. (2020) APOE4 exacerbates synapse loss and neurodegeneration in Alzheimer's disease patient iPSC-derived cerebral organoids. Nature Communications 11, Article number: 5540.CrossRefGoogle ScholarPubMed
Barbier, P et al. (2019) Role of tau as a microtubule-associated protein: structural and functional aspects. Frontiers in Aging Neuroscience 11, 1–14.CrossRefGoogle ScholarPubMed
Arendt, T, Stieler, JT and Holzer, M (2016) Tau and tauopathies. Brain Research Bulletin 126, 238292.CrossRefGoogle ScholarPubMed
Cherry, JD et al. (2021) Tau isoforms are differentially expressed across the hippocampus in chronic traumatic encephalopathy and Alzheimer's disease. Acta Neuropathologica Communications 9, 117.CrossRefGoogle ScholarPubMed
Moore, S et al. (2015) APP metabolism regulates tau proteostasis in human cerebral cortex neurons. Cell Reports 11, 689696.CrossRefGoogle ScholarPubMed
Choi, H et al. (2020) Acetylation changes tau interactome to degrade tau in Alzheimer's disease animal and organoid models. Aging Cell 19, 113.CrossRefGoogle ScholarPubMed
Ding, H, Dolan, PJ and Johnson, GVW (2008) Histone deacetylase 6 interacts with the microtubule-associated protein tau. Journal of Neurochemistry 106, 2119.CrossRefGoogle ScholarPubMed
Baxter, PS et al. (2021) Microglial identity and inflammatory responses are controlled by the combined effects of neurons and astrocytes. Cell Reports 34, 108882.CrossRefGoogle ScholarPubMed
Papadimitriou, C et al. (2018) 3D culture method for Alzheimer's disease modeling reveals interleukin-4 rescues Aβ42-induced loss of human neural stem cell plasticity. Developmental Cell 46, 85101.CrossRefGoogle ScholarPubMed
Griffin, K et al. (2020) Human stem cell-derived aggregates of forebrain astroglia respond to amyloid beta oligomers. Tissue Engineering 26, 527542.CrossRefGoogle ScholarPubMed
Argaw, AT et al. (2012) Astrocyte-derived VEGF-A drives blood–brain barrier disruption in CNS inflammatory disease. The Journal of Clinical Investigation 122, 2454.CrossRefGoogle ScholarPubMed
Park, J et al. (2018) A 3D human triculture system modeling neurodegeneration and neuroinflammation in Alzheimer's disease. Nature Neuroscience 21, 941951.CrossRefGoogle ScholarPubMed
TCW, J et al. (2022) Cholesterol and matrisome pathways dysregulated in astrocytes and microglia. Cell 185, 22132233.e25.CrossRefGoogle ScholarPubMed
Zhang, W et al. (2022) Impairment of the autophagy–lysosomal pathway in Alzheimer's diseases: pathogenic mechanisms and therapeutic potential. Acta Pharmaceutica Sinica B 12, 10191040.CrossRefGoogle ScholarPubMed
Wang, W et al. (2020) Mitochondria dysfunction in the pathogenesis of Alzheimer's disease: recent advances. Molecular Neurodegeneration 15, 122.CrossRefGoogle ScholarPubMed
Eckert, A, Schmitt, K and Götz, J (2011) Mitochondrial dysfunction – the beginning of the end in Alzheimer's disease? Separate and synergistic modes of tau and amyloid-β toxicity. Alzheimer's Research & Therapy 3, 15.CrossRefGoogle ScholarPubMed
Brunetti, D et al. (2021) Role of PITRM1 in mitochondrial dysfunction and neurodegeneration. Biomedicines 9, 833.CrossRefGoogle ScholarPubMed
Pérez, MJ et al. (2020) Loss of function of the mitochondrial peptidase PITRM1 induces proteotoxic stress and Alzheimer's disease-like pathology in human cerebral organoids. Molecular Psychiatry 2020 118.Google Scholar
Marsh, J and Alifragis, P (2018) Synaptic dysfunction in Alzheimer's disease: the effects of amyloid beta on synaptic vesicle dynamics as a novel target for therapeutic intervention. Neural Regeneration Research 13, 616623.Google ScholarPubMed
Ghatak, S et al. (2019) Mechanisms of hyperexcitability in Alzheimer's disease hiPSC-derived neurons and cerebral organoids versus isogenic control. eLife 8, 122.CrossRefGoogle Scholar
Wang, R and Reddy, PH (2017) Role of glutamate and NMDA receptors in Alzheimer's disease. Journal of Alzheimer's disease 57, 10411048.CrossRefGoogle ScholarPubMed
Hynd, M, Scott, H and Dodd, P (2004) Glutamate-mediated excitotoxicity and neurodegeneration in Alzheimer's disease. Neurochemistry International 45, 583595.CrossRefGoogle ScholarPubMed
Ghatak, S et al. (2020) NitroSynapsin ameliorates hypersynchronous neural network activity in Alzheimer hiPSC models. Molecular Psychiatry 26, 57515765.CrossRefGoogle ScholarPubMed
Yin, J and Vandongen, AM (2021) Enhanced neuronal activity and asynchronous calcium transients revealed in a 3D organoid model of Alzheimer's disease. ACS Biomaterials Science and Engineering 7, 254264.CrossRefGoogle Scholar
Bi, F-C et al. (2021) Optimization of cerebral organoids: a more qualified model for Alzheimer's disease research. Translational Neurodegeneration 10, 113.CrossRefGoogle ScholarPubMed
Sabogal-Guáqueta, AM et al. (2020) Microglia alterations in neurodegenerative diseases and their modeling with human induced pluripotent stem cell and other platforms. Progress in Neurobiology 190, 101805.CrossRefGoogle ScholarPubMed
Muffat, J et al. (2016) Efficient derivation of microglia-like cells from human pluripotent stem cells. Nature Medicine 22, 13581367.CrossRefGoogle ScholarPubMed
Douvaras, P et al. (2017) Directed differentiation of human pluripotent stem cells to microglia. Stem Cell Reports 8, 15161524.CrossRefGoogle ScholarPubMed
Haenseler, W et al. (2017) A highly efficient human pluripotent stem cell microglia model displays a neuronal-co-culture-specific expression profile and inflammatory response. Stem Cell Reports 8, 17271742.CrossRefGoogle ScholarPubMed
Pandya, H et al. (2017) Differentiation of human and murine induced pluripotent stem cells to microglia-like cells. Nature Neuroscience 20, 753759.CrossRefGoogle ScholarPubMed
Takata, K et al. (2017) Induced-pluripotent-stem-cell-derived primitive macrophages provide a platform for modeling tissue-resident macrophage differentiation and function. Immunity 47, 183198.CrossRefGoogle ScholarPubMed
Brownjohn, PW et al. (2018) Functional studies of missense TREM2 mutations in human stem cell-derived microglia. Stem Cell Reports 10, 12941307.CrossRefGoogle ScholarPubMed
Garcia-Reitboeck, P et al. (2018) Human induced pluripotent stem cell-derived microglia-like cells harboring TREM2 missense mutations show specific deficits in phagocytosis. Cell Reports 24, 23002311.CrossRefGoogle ScholarPubMed
McQuade, A et al. (2018) Development and validation of a simplified method to generate human microglia from pluripotent stem cells. Molecular Neurodegeneration 13, 113.CrossRefGoogle ScholarPubMed
Papaspyropoulos, A et al. (2020) Modeling and targeting Alzheimer's disease with organoids. Frontiers in Pharmacology 11, 396.CrossRefGoogle ScholarPubMed
Montagne, A, Zhao, Z and Zlokovic, BV (2017) Alzheimer's disease: a matter of blood–brain barrier dysfunction? The Journal of Experimental Medicine 214, 3151.CrossRefGoogle ScholarPubMed
Cheah, P-S, Mason, JO and Ling, KH (2019) Challenges and future perspectives for 3D cerebral organoids as a model for complex brain disorders. Neuroscience Research Notes 2, 16.CrossRefGoogle Scholar
Bergmann, S et al. (2018) Blood–brain-barrier organoids for investigating the permeability of CNS therapeutics. Nature Protocols 13, 28272843.CrossRefGoogle ScholarPubMed
Lauschke, K, Frederiksen, L and Hall, VJ (2017) Paving the way toward complex blood–brain barrier models using pluripotent stem cells. Stem Cells Development 26, 857874.CrossRefGoogle ScholarPubMed
Cakir, B et al. (2019) Engineering of human brain organoids with a functional vascular-like system. Nature Methods 16, 11691175.CrossRefGoogle ScholarPubMed
Shi, Y et al. (2020) Vascularized human cortical organoids (vOrganoids) model cortical development in vivo. PLoS Biology 18, e3000705.CrossRefGoogle ScholarPubMed
Matsui, TK et al. (2021) Vascularization of human brain organoids. Stem Cells 39, 10171024.CrossRefGoogle ScholarPubMed
Camp, JG et al. (2015) Human cerebral organoids recapitulate gene expression programs of fetal neocortex development. Proceedings of the National Academy of Sciences of the United States of America 112, 1567215677.CrossRefGoogle ScholarPubMed
Kelava, I and Lancaster, MA (2016) Dishing out mini-brains: current progress and future prospects in brain organoid research. Developmental Biology 420, 199209.CrossRefGoogle ScholarPubMed
Vera, E, Bosco, N and Studer, L (2016) Generating late-onset human iPSC-based disease models by inducing neuronal age-related phenotypes through telomerase manipulation. Cell Reports 17, 11841192.CrossRefGoogle ScholarPubMed
Miller, JD et al. (2013) Human iPSC-based modeling of late-onset disease via progerin-induced aging. Cell Stem Cell 13, 691705.CrossRefGoogle ScholarPubMed
Szebényi, K et al. (2021) Human ALS/FTD brain organoid slice cultures display distinct early astrocyte and targetable neuronal pathology. Nature Neuroscience 24, 15421554.CrossRefGoogle ScholarPubMed
Song, G et al. (2021) The application of brain organoid technology in stroke research: challenges and prospects. Frontiers in Cellular Neuroscience 15, 203.CrossRefGoogle ScholarPubMed
Daviaud, N, Friedel, RH and Zou, H (2018) Vascularization and engraftment of transplanted human cerebral organoids in mouse cortex. eNeuro 5, ENEURO.0219.CrossRefGoogle ScholarPubMed
Mansour, AA et al. (2018) An in vivo model of functional and vascularized human brain organoids. Nature Biotechnology 36, 432441.CrossRefGoogle Scholar
Wang, Z et al. (2020) Cerebral organoids transplantation improves neurological motor function in rat brain injury. CNS Neuroscience & Therapeutics 26, 682697.CrossRefGoogle ScholarPubMed
Dong, X et al. (2020) Human cerebral organoids establish subcortical projections in the mouse brain after transplantation. Molecular Psychiatry 26, 29642976.CrossRefGoogle ScholarPubMed
Wang, SN et al. (2020) Cerebral organoids repair ischemic stroke brain injury. Translational Stroke Research 11, 9831000.CrossRefGoogle ScholarPubMed
Ooi, L et al. (2020) If human brain organoids are the answer to understanding dementia, what are the questions? Neuroscientist 26, 438454.CrossRefGoogle ScholarPubMed
University of Cambridge (2021) Lab-grown ‘mini brains’ hint at treatments for neurodegenerative diseases, University of Cambridge. Available at https://www.cam.ac.uk/research/news/lab-grown-mini-brains-hint-at-treatments-for-neurodegenerative-diseases (Accessed 27 October 2021).Google Scholar