Human-Induced Pluripotent Stem Cell Technology: Toward the Future of Personalized Psychiatry
Abstract
:1. The Limitations of Personalized Medicine in Psychiatry
2. The Induced Pluripotent Stem Cells
2.1. Definition and Development
2.2. Three-Dimensional (3D) Brain iPSC Culture Models (Organoids)
3. The Potential of Induced Pluripotent Stem Cells in Personalized Psychiatry
3.1. Schizophrenia
3.2. Major Depressive Disorder
3.3. Bipolar Disorders
3.4. Autism Spectrum Disorder
3.5. Conclusions
4. Limitations of iPSC Technology and Future Perspectives
Author Contributions
Funding
Institutional Review Board Statement
Informed Consent Statement
Data Availability Statement
Conflicts of Interest
References
- Perna, G.; Nemeroff, C.B. Personalized Medicine in Psychiatry: Back to the Future. Pers. Med. Psychiatry 2017, 1, 1. [Google Scholar] [CrossRef]
- Sugeir, S.; Naylor, S. Critical Care and Personalized or Precision Medicine: Who Needs Whom? J. Crit. Care 2018, 43, 401–405. [Google Scholar] [CrossRef] [PubMed]
- Rocca, A.; Kholodenko, B.N. Can Systems Biology Advance Clinical Precision Oncology? Cancers 2021, 13, 6312. [Google Scholar] [CrossRef]
- Levchenko, A.; Nurgaliev, T.; Kanapin, A.; Samsonova, A.; Gainetdinov, R.R. Current Challenges and Possible Future Developments in Personalized Psychiatry with an Emphasis on Psychotic Disorders. Heliyon 2020, 6, e03990. [Google Scholar] [CrossRef] [PubMed]
- American Psychiatric Association. Diagnostic and Statistical Manual of Mental Disorders (DSM-5®), 5th ed.; American Psychiatric Association: Arlington, VA, USA; Washington DC, USA, 2013. [Google Scholar]
- World Health Organization. World Health Organisation International Statistical Classification of Diseases and Related Health Problems, 11th ed.; World Health Organization: Geneva, Switzerland, 2004.
- Clark, L.A.; Cuthbert, B.; Lewis-Fernández, R.; Narrow, W.E.; Reed, G.M. Three Approaches to Understanding and Classifying Mental Disorder: ICD-11, DSM-5, and the National Institute of Mental Health’s Research Domain Criteria (RDoC). Psychol. Sci. Public Interes. 2017, 18, 72–145. [Google Scholar] [CrossRef] [PubMed]
- Smoller, J.W.; Andreassen, O.A.; Edenberg, H.J.; Faraone, S.V.; Glatt, S.J.; Kendler, K.S. Psychiatric Genetics and the Structure of Psychopathology. Mol. Psychiatry 2019, 24, 409–420. [Google Scholar] [CrossRef] [PubMed]
- Craddock, N.; Owen, M.J. The Beginning of the End for the Kraepelinian Dichotomy. Br. J. Psychiatry 2005, 186, 364–366. [Google Scholar] [CrossRef]
- Zimmerman, M. Why Hierarchical Dimensional Approaches to Classification Will Fail to Transform Diagnosis in Psychiatry. World Psychiatry 2021, 20, 70–71. [Google Scholar] [CrossRef]
- Howland, J.G.; Greenshaw, A.J.; Winship, I.R. Practical Aspects of Animal Models of Psychiatric Disorders. Can. J. Psychiatry 2019, 64, 3–4. [Google Scholar] [CrossRef]
- Hodge, R.D.; Bakken, T.E.; Miller, J.A.; Smith, K.A.; Barkan, E.R.; Graybuck, L.T.; Close, J.L.; Long, B.; Johansen, N.; Penn, O.; et al. Conserved Cell Types with Divergent Features in Human versus Mouse Cortex. Nature 2019, 573, 61–68. [Google Scholar] [CrossRef]
- Bakken, T.E.; Miller, J.A.; Ding, S.L.; Sunkin, S.M.; Smith, K.A.; Ng, L.; Szafer, A.; Dalley, R.A.; Royall, J.J.; Lemon, T.; et al. A Comprehensive Transcriptional Map of Primate Brain Development. Nature 2016, 535, 367–375. [Google Scholar] [CrossRef] [PubMed]
- Mccullumsmith, R.E.; Hammond, J.H.; Shan, D.; Meador-Woodruff, J.H. Postmortem Brain: An Underutilized Substrate for Studying Severe Mental Illness. Neuropsychopharmacology 2014, 39, 65–87. [Google Scholar] [CrossRef] [PubMed]
- Takahashi, K.; Yamanaka, S. Induction of Pluripotent Stem Cells from Mouse Embryonic and Adult Fibroblast Cultures by Defined Factors. Cell 2006, 126, 663–676. [Google Scholar] [CrossRef] [PubMed]
- Yu, J.; Vodyanik, M.A.; Smuga-Otto, K.; Antosiewicz-Bourget, J.; Frane, J.L.; Tian, S.; Nie, J.; Jonsdottir, G.A.; Ruotti, V.; Stewart, R.; et al. Induced Pluripotent Stem Cell Lines Derived from Human Somatic Cells. Science 2007, 318, 1917–1920. [Google Scholar] [CrossRef]
- Abdullah, A.I.; Pollock, A.; Sun, T. The path from skin to brain: Generation of functional neurons from fibroblasts. Mol. Neurobiol. 2012, 45, 586–595. [Google Scholar] [CrossRef]
- Park, B.; Yoo, K.H.; Kim, C. Hematopoietic Stem Cell Expansion and Generation: The Ways to Make a Breakthrough. Blood Res. 2015, 50, 194–203. [Google Scholar] [CrossRef]
- Liu, G.; David, B.T.; Trawczynski, M.; Fessler, R.G. Advances in Pluripotent Stem Cells: History, Mechanisms, Technologies, and Applications. Stem Cell Rev. Rep. 2020, 16, 3–32. [Google Scholar] [CrossRef]
- Nelson, T.J.; Martinez-Fernandez, A.; Yamada, S.; Perez-Terzic, C.; Ikeda, Y.; Terzic, A. Repair of Acute Myocardial Infarction by Human Stemness Factors Induced Pluripotent Stem Cells. Circulation 2009, 120, 408–416. [Google Scholar] [CrossRef]
- Zhang, D.; Jiang, W.; Liu, M.; Sui, X.; Yin, X.; Chen, S.; Shi, Y.; Deng, H. Highly Efficient Differentiation of Human ES Cells and IPS Cells into Mature Pancreatic Insulin-Producing Cells. Cell Res. 2009, 19, 429–438. [Google Scholar] [CrossRef]
- Okita, K.; Ichisaka, T.; Yamanaka, S. Generation of Germline-Competent Induced Pluripotent Stem Cells. Nature 2007, 448, 313–317. [Google Scholar] [CrossRef]
- Montini, E.; Cesana, D.; Schmidt, M.; Sanvito, F.; Ponzoni, M.; Bartholomae, C.; Sergi Sergi, L.; Benedicenti, F.; Ambrosi, A.; Di Serio, C.; et al. Hematopoietic Stem Cell Gene Transfer in a Tumor-Prone Mouse Model Uncovers Low Genotoxicity of Lentiviral Vector Integration. Nat. Biotechnol. 2006, 24, 687–696. [Google Scholar] [CrossRef] [PubMed]
- Fusaki, N.; Ban, H.; Nishiyama, A.; Saeki, K.; Hasegawa, M. Efficient Induction of Transgene-Free Human Pluripotent Stem Cells Using a Vector Based on Sendai Virus, an RNA Virus That Does Not Integrate into the Host Genome. Proc. Jpn. Acad. Ser. B Phys. Biol. Sci. 2009, 85, 348–362. [Google Scholar] [CrossRef] [PubMed]
- Sayed, N.; Wong, W.T.; Ospino, F.; Meng, S.; Lee, J.; Jha, A.; Dexheimer, P.; Aronow, B.J.; Cooke, J.P. Transdifferentiation of Human Fibroblasts to Endothelial Cells: Role of Innate Immunity. Circulation 2015, 131, 300–309. [Google Scholar] [CrossRef] [PubMed]
- De Masi, C.; Spitalieri, P.; Murdocca, M.; Novelli, G.; Sangiuolo, F. Application of CRISPR/Cas9 to Human-Induced Pluripotent Stem Cells: From Gene Editing to Drug Discovery. Hum. Genom. 2020, 14, 25. [Google Scholar] [CrossRef]
- Duarte, F.; Déglon, N. Genome Editing for CNS Disorders. Front. Neurosci. 2020, 14, 10–22. [Google Scholar] [CrossRef]
- Bassett, A.R. Editing the Genome of HiPSC with CRISPR/Cas9: Disease Models. Mamm. Genome 2017, 28, 348–364. [Google Scholar] [CrossRef]
- Grath, A.; Dai, G. Direct Cell Reprogramming for Tissue Engineering and Regenerative Medicine. J. Biol. Eng. 2019, 13, 14. [Google Scholar] [CrossRef]
- Meng, F.; Chen, S.; Miao, Q.; Zhou, K.; Lao, Q.; Zhang, X.; Guo, W.; Jiao, J. Induction of Fibroblasts to Neurons through Adenoviral Gene Delivery. Cell Res. 2012, 22, 436–440. [Google Scholar] [CrossRef]
- Rubio, A.; Luoni, M.; Giannelli, S.G.; Radice, I.; Iannielli, A.; Cancellieri, C.; Di Berardino, C.; Regalia, G.; Lazzari, G.; Menegon, A.; et al. Rapid and Efficient CRISPR/Cas9 Gene Inactivation in Human Neurons during Human Pluripotent Stem Cell Differentiation and Direct Reprogramming. Sci. Rep. 2016, 6, 37540. [Google Scholar] [CrossRef]
- Frederiksen, H.R.; Doehn, U.; Tveden-nyborg, P. Non-Immunogenic Induced Pluripotent Stem Cells, a Promising Way Forward for Allogenic Transplantations for Neurological Disorders. Front. Genome Ed. 2021, 2, 623717. [Google Scholar] [CrossRef]
- Lancaster, M.A.; Renner, M.; Martin, C.A.; Wenzel, D.; Bicknell, L.S.; Hurles, M.E.; Homfray, T.; Penninger, J.M.; Jackson, A.P.; Knoblich, J.A. Cerebral Organoids Model Human Brain Development and Microcephaly. Nature 2013, 501, 373–379. [Google Scholar] [CrossRef] [PubMed]
- Eglen, R.M.; Reisine, T. Human IPS Cell-Derived Patient Tissues and 3D Cell Culture Part 1: Target Identification and Lead Optimization. SLAS Technol. 2019, 24, 3–17. [Google Scholar] [CrossRef] [PubMed]
- Srikanth, P.; Young-Pearse, T.L. Stem Cells on the Brain: Modeling Neurodevelopmental and Neurodegenerative Diseases Using Human Induced Pluripotent Stem Cells. J. Neurogenet. 2014, 28, 5–29. [Google Scholar] [CrossRef] [PubMed]
- Qian, X.; Nguyen, H.N.; Song, M.M.; Hadiono, C.; Ogden, S.C.; Hammack, C.; Yao, B.; Hamersky, G.R.; Jacob, F.; Zhong, C.; et al. Brain-Region-Specific Organoids Using Mini-Bioreactors for Modeling ZIKV Exposure. Cell 2016, 165, 1238–1254. [Google Scholar] [CrossRef]
- Sun, N.; Meng, X.; Liu, Y.; Song, D.; Jiang, C.; Cai, J. Applications of Brain Organoids in Neurodevelopment and Neurological Diseases. J. Biomed. Sci. 2021, 28, 30. [Google Scholar] [CrossRef]
- Brennand, K.J.; Simone, A.; Jou, J.; Gelboin-Burkhart, C.; Tran, N.; Sangar, S.; Li, Y.; Mu, Y.; Chen, G.; Yu, D.; et al. Modelling Schizophrenia Using Human Induced Pluripotent Stem Cells. Nature 2011, 473, 221–225. [Google Scholar] [CrossRef]
- Lee, J.; Song, S.; Lee, J.; Kang, J.; Choe, E.K.; Lee, T.Y.; Chon, M.W.; Kim, M.; Kim, S.W.; Chun, M.S.; et al. Impaired migration of autologous induced neural stem cells from patients with schizophrenia and implications for genetic risk for psychosis. Schizophr. Res. 2022, 246, 225–234. [Google Scholar] [CrossRef]
- Narla, S.T.; Lee, Y.W.; Benson, C.A.; Sarder, P.; Brennand, K.J.; Stachowiak, E.K.; Stachowiak, M.K. Common Developmental Genome Deprogramming in Schizophrenia—Role of Integrative Nuclear FGFR1 Signaling (INFS). Schizophr. Res. 2017, 185, 17–32. [Google Scholar] [CrossRef]
- Yu, D.X.; Di Giorgio, F.P.; Yao, J.; Marchetto, M.C.; Brennand, K.; Wright, R.; Mei, A.; McHenry, L.; Lisuk, D.; Grasmick, J.M.; et al. Modeling Hippocampal Neurogenesis Using Human Pluripotent Stem Cells. Stem Cell Rep. 2014, 2, 295–310. [Google Scholar] [CrossRef]
- Sarkar, A.; Mei, A.; Paquola, A.; Stern, S.; Bardy, C.; Klug, J.R.; Kim, S.; Neshat, N.; Kim, H.J.; Ku, M.; et al. Efficient Generation of CA3 Neurons from Human Pluripotent Stem Cells Enables Modeling of Hippocampal Connectivity In Vitro. Cell Stem Cell 2018, 22, 684–697.e9. [Google Scholar] [CrossRef]
- Brennand, K.; Savas, J.N.; Kim, Y.; Tran, N.; Simone, A.; Hashimoto-Torii, K.; Beaumont, K.G.; Kim, H.J.; Topol, A.; Ladran, I.; et al. Phenotypic Differences in HiPSC NPCs Derived from Patients with Schizophrenia. Mol. Psychiatry 2015, 20, 361–368. [Google Scholar] [CrossRef] [PubMed]
- Hashimoto-torii, K.; Torii, M.; Fujimoto, M.; Nakai, A. Roles of Heat Shock Factor 1 in neuronal response to fetal environmental risks and its relevance to brain disorders. Neuron 2015, 82, 560–572. [Google Scholar] [CrossRef] [PubMed]
- Stachowiak, E.K.; Benson, C.A.; Narla, S.T.; Dimitri, A.; Chuye, L.E.B.; Dhiman, S.; Harikrishnan, K.; Elahi, S.; Freedman, D.; Brennand, K.J.; et al. Cerebral Organoids Reveal Early Cortical Maldevelopment in Schizophrenia—Computational Anatomy and Genomics, Role of FGFR1. Transl. Psychiatry 2017, 7, 6. [Google Scholar] [CrossRef] [PubMed]
- Benson, C.A.; Powell, H.R.; Liput, M.; Dinham, S.; Freedman, D.A.; Ignatowski, T.A.; Stachowiak, E.K.; Stachowiak, M.K. Immune Factor, TNFα, Disrupts Human Brain Organoid Development Similar to Schizophrenia—Schizophrenia Increases Developmental Vulnerability to TNFα. Front. Cell. Neurosci. 2020, 14, 233. [Google Scholar] [CrossRef] [PubMed]
- Kathuria, A.; Lopez-Lengowski, K.; Jagtap, S.S.; McPhie, D.; Perlis, R.H.; Cohen, B.M.; Karmacharya, R. Transcriptomic Landscape and Functional Characterization of Induced Stem Cell–Derived Cerebral Organoids In. JAMA Psychiatry 2020, 77, 745. [Google Scholar] [CrossRef] [PubMed]
- Notaras, M.; Lodhi, A.; Fang, H.; Greening, D.; Colak, D. The Proteomic Architecture of Schizophrenia IPSC-Derived Cerebral Organoids Reveals Alterations in GWAS and Neuronal Development Factors. Transl. Psychiatry 2021, 11, 541. [Google Scholar] [CrossRef] [PubMed]
- Notaras, M.; Lodhi, A.; Dündar, F.; Collier, P.; Sayles, N.M.; Tilgner, H.; Greening, D.; Colak, D. Schizophrenia is defined by cell-specific neuropathology and multiple neurodevelopmental mechanisms in patient-derived cerebral organoids. Mol. Psychiatry 2022, 27, 1416–1434. [Google Scholar] [CrossRef] [PubMed]
- Page, S.C.; Sripathy, S.R.; Farinelli, F.; Ye, Z.; Wang, Y.; Hiler, D.J.; Pattie, E.A.; Nguyen, C.V.; Tippani, M.; Moses, R.L.; et al. Electrophysiological measures from human iPSC-derived neurons are associated with schizophrenia clinical status and predict individual cognitive performance. Proc. Natl. Acad. Sci. USA 2022, 119, e2109395119. [Google Scholar] [CrossRef]
- Hagihara, H.; Takao, K.; Walton, N.M.; Matsumoto, M.; Miyakawa, T. Immature Dentate Gyrus: An Endophenotype of Neuropsychiatric Disorders. Neural Plast. 2013, 2013, 318596. [Google Scholar] [CrossRef]
- Ahmad, R.; Sportelli, V.; Ziller, M.; Spengler, D.; Hoffmann, A. Tracing Early Neurodevelopment in Schizophrenia with Induced Pluripotent Stem Cells. Cells 2018, 7, 140. [Google Scholar] [CrossRef]
- Schrode, N.; Ho, S.M.; Yamamuro, K.; Dobbyn, A.; Huckins, L.; Matos, M.R.; Cheng, E.; Deans, P.J.M.; Flaherty, E.; Barretto, N.; et al. Synergistic effects of common schizophrenia risk variants. Nat. Genet. 2019, 51, 1475–1485. [Google Scholar] [CrossRef] [PubMed]
- Srikanth, P.; Han, K.; Callahan, D.G.; Makovkina, E.; Muratore, C.R.; Lalli, M.A.; Zhou, H.; Boyd, J.D.; Kosik, K.S.; Selkoe, D.J.; et al. Genomic DISC1 Disruption in HiPSCs Alters Wnt Signaling and Neural Cell Fate. Cell Rep. 2015, 12, 1414–1429. [Google Scholar] [CrossRef] [PubMed]
- Srikanth, P.; Lagomarsino, V.N.; Muratore, C.R.; Ryu, S.C.; He, A.; Taylor, W.M.; Zhou, C.; Arellano, M.; Young-Pearse, T.L. Shared Effects of DISC1 Disruption and Elevated WNT Signaling in Human Cerebral Organoids. Transl. Psychiatry 2018, 8, 77. [Google Scholar] [CrossRef] [PubMed]
- Warden, D.; Rush, A.J.; Trivedi, M.H.; Fava, M.; Wisniewski, S.R. The STAR*D Project Results: A Comprehensive Review of Findings. Curr. Psychiatry Rep. 2007, 9, 449–459. [Google Scholar] [CrossRef]
- Vadodaria, K.C.; Ji, Y.; Skime, M.; Paquola, A.; Nelson, T.; Hall-Flavin, D.; Fredlender, C.; Heard, K.J.; Deng, Y.; Le, A.T.; et al. Serotonin-Induced Hyperactivity in SSRI-Resistant Major Depressive Disorder Patient-Derived Neurons. Mol. Psychiatry 2019, 24, 795–807. [Google Scholar] [CrossRef]
- Vadodaria, K.C.; Ji, Y.; Skime, M.; Paquola, A.C.; Nelson, T.; Hall-Flavin, D.; Heard, K.J.; Fredlender, C.; Deng, Y.; Elkins, J.; et al. Altered Serotonergic Circuitry in SSRI-Resistant Major Depressive Disorder Patient-Derived Neurons. Mol. Psychiatry 2019, 24, 808–818. [Google Scholar] [CrossRef]
- Avior, Y.; Ron, S.; Kroitorou, D.; Albeldas, C.; Lerner, V.; Corneo, B.; Nitzan, E.; Laifenfeld, D.; Cohen Solal, T. Depression patient-derived cortical neurons reveal potential biomarkers for antidepressant response. Transl. Psychiatry. 2021, 11, 201. [Google Scholar] [CrossRef]
- Cavalleri, L.; Merlo Pich, E.; Millan, M.J.; Chiamulera, C.; Kunath, T.; Spano, P.F.; Collo, G. Ketamine Enhances Structural Plasticity in Mouse Mesencephalic and Human IPSC-Derived Dopaminergic Neurons via AMPAR-Driven BDNF and MTOR Signaling. Mol. Psychiatry 2018, 23, 812–823. [Google Scholar] [CrossRef]
- Collo, G.; Cavalleri, L.; Chiamulera, C.; Merlo Pich, E. (2 R,6 R)-Hydroxynorketamine Promotes Dendrite Outgrowth in Human Inducible Pluripotent Stem Cell-Derived Neurons through AMPA Receptor with Timing and Exposure Compatible with Ketamine Infusion Pharmacokinetics in Humans. Neuroreport 2018, 29, 1425–1430. [Google Scholar] [CrossRef]
- Katori, S.; Noguchi-Katori, Y.; Okayama, A.; Kawamura, Y.; Luo, W.; Sakimura, K.; Hirabayashi, T.; Iwasato, T.; Yagi, T. Protocadherin-AC2 Is Required for Diffuse Projections of Serotonergic Axons. Sci. Rep. 2017, 7, 15908. [Google Scholar] [CrossRef]
- Chen, H.M.; DeLong, C.J.; Bame, M.; Rajapakse, I.; Herron, T.J.; McInnis, M.G.; O’Shea, K.S. Transcripts Involved in Calcium Signaling and Telencephalic Neuronal Fate Are Altered in Induced Pluripotent Stem Cells from Bipolar Disorder Patients. Transl. Psychiatry 2014, 4, e375. [Google Scholar] [CrossRef] [PubMed]
- Tobe, B.T.D.; Crain, A.M.; Winquist, A.M.; Calabrese, B.; Makihara, H.; Zhao, W.N.; Lalonde, J.; Nakamura, H.; Konopaske, G.; Sidor, M.; et al. Probing the lithium-response pathway in hiPSCs implicates the phosphoregulatory set-point for a cytoskeletal modulator in bipolar pathogenesis. Proc. Natl. Acad. Sci. USA 2017, 114, E4462–E4471. [Google Scholar] [CrossRef] [PubMed]
- Bavamian, S.; Mellios, N.; Lalonde, J.; Fass, D.M.; Wang, J.; Sheridan, S.D.; Madison, J.M.; Zhou, F.; Rueckert, E.H.; Barker, D.; et al. Dysregulation of MiR-34a Links Neuronal Development to Genetic Risk Factors for Bipolar Disorder. Mol. Psychiatry 2015, 20, 573–584. [Google Scholar] [CrossRef] [PubMed]
- Ishii, T.; Ishikawa, M.; Fujimori, K.; Maeda, T.; Kushima, I.; Arioka, Y.; Mori, D.; Nakatake, Y.; Yamagata, B.; Nio, S.; et al. In Vitro Modeling of the Bipolar Disorder and Schizophrenia Using Patient-Derived Induced Pluripotent Stem Cells with Copy Number Variations of PCDH15 and RELN. eNeuro 2019, 6. [Google Scholar] [CrossRef]
- Madison, J.M.; Zhou, F.; Nigam, A.; Hussain, A.; Barker, D.D.; Nehme, R.; Van Der Ven, K.; Hsu, J.; Wolf, P.; Fleishman, M.; et al. Characterization of Bipolar Disorder Patient-Specific Induced Pluripotent Stem Cells from a Family Reveals Neurodevelopmental and MRNA Expression Abnormalities. Mol. Psychiatry 2015, 20, 703–717. [Google Scholar] [CrossRef]
- Kim, K.H.; Liu, J.; Galvin, R.J.S.; Dage, J.L.; Egeland, J.A.; Smith, R.C.; Merchant, K.M.; Paul, S.M. Transcriptomic Analysis of Induced Pluripotent Stem Cells Derived from Patients with Bipolar Disorder from an Old Order Amish Pedigree. PLoS ONE 2015, 10, e0142693. [Google Scholar] [CrossRef]
- Vadodaria, K.C.; Mendes, A.P.D.; Mei, A.; Racha, V.; Erikson, G.; Shokhirev, M.N.; Oefner, R.; Heard, K.J.; McCarthy, M.J.; Eyler, L.; et al. Altered Neuronal Support and Inflammatory Response in Bipolar Disorder Patient-Derived Astrocytes. Stem Cell Rep. 2021, 16, 825–835. [Google Scholar] [CrossRef]
- Mertens, J.; Wang, Q.W.; Kim, Y.; Yu, D.X.; Pham, S.; Yang, B.; Zheng, Y.; Diffenderfer, K.E.; Zhang, J.; Soltani, S.; et al. Differential Responses to Lithium in Hyperexcitable Neurons from Patients with Bipolar Disorder. Nature 2015, 527, 95–99. [Google Scholar] [CrossRef]
- Osete, J.R.; Akkouh, I.A.; de Assis, D.R.; Szabo, A.; Frei, E.; Hughes, T.; Smeland, O.B.; Steen, N.E.; Andreassen, O.A.; Djurovic, S. Lithium increases mitochondrial respiration in iPSC-derived neural precursor cells from lithium responders. Mol. Psychiatry 2021, 26, 6789–6805. [Google Scholar] [CrossRef]
- Stern, S.; Santos, R.; Marchetto, M.C.; Mendes, A.P.D.; Rouleau, G.A.; Biesmans, S.; Wang, Q.W.; Yao, J.; Charnay, P.; Bang, A.G.; et al. Neurons Derived from Patients with Bipolar Disorder Divide into Intrinsically Different Sub-Populations of Neurons, Predicting the Patients’ Responsiveness to Lithium. Mol. Psychiatry 2018, 23, 1453–1465. [Google Scholar] [CrossRef]
- Hunsberger, J.G.; Fessler, E.B.; Chibane, F.L.; Leng, Y.; Maric, D.; Elkahloun, A.G.; Chuang, D.M. Mood Stabilizer-Regulated MiRNAs in Neuropsychiatric and Neurodegenerative Diseases: Identifying Associations and Functions. Am. J. Transl. Res. 2013, 5, 450–464. [Google Scholar] [PubMed]
- Palladino, V.S.; Subrata, N.O.C.; Geburtig-Chiocchetti, A.; McNeill, R.; Hoffmann, P.; Reif, A.; Kittel-Schneider, S. Generation of Human Induced Pluripotent Stem Cell Lines (HiPSC) from One Bipolar Disorder Patient Carrier of a DGKH Risk Haplotype and One Non-Risk-Variant-Carrier Bipolar Disorder Patient. Stem Cell Res. 2018, 32, 104–109. [Google Scholar] [CrossRef] [PubMed]
- Freitas, B.C.; Mei, A.; Mendes, A.P.D.; Beltrão-Braga, P.C.B.; Marchetto, M.C. Modeling Inflammation in Autism Spectrum Disorders Using Stem Cells. Front. Pediatr. 2018, 6, 394. [Google Scholar] [CrossRef] [PubMed]
- Marchetto, M.C.; Belinson, H.; Tian, Y.; Freitas, B.C.; Fu, C.; Beltrao-braga, P.; Trujillo, C.A.; Mendes, A.P.D.; Nunez, Y.; Ou, J.; et al. Altered proliferation and networks in neural cells derived from idiopathic autistic individuals. Mol. Psychiatry 2017, 22, 820–835. [Google Scholar] [CrossRef] [PubMed]
- Schafer, S.T.; Paquola, A.C.M.; Stern, S.; Gosselin, D.; Ku, M.; Pena, M.; Kuret, T.J.M.; Liyanage, M.; Mansour, A.A.; Jaeger, N.; et al. Pathological priming causes developmental gene network heterochronicity in autistic subject-derived neurons. Nat. Neurosci. 2019, 22, 243–255. [Google Scholar] [CrossRef]
- Mariani, J.; Coppola, G.; Zhang, P.; Abyzov, A.; Tomasini, L.; Amenduni, M.; Szekely, A.; Palejev, D.; Wilson, M.; Gerstein, M.; et al. FOXG1-Dependent Dysregulation of GABA/Glutamate Neuron Differentiation in Autism Spectrum Disorders. Cell 2016, 162, 375–390. [Google Scholar] [CrossRef]
- Adhya, D.; Swarup, V.; Nagy, V.R.; Dutan, L.; Shum, C.; Valencia-Alarcón, E.P.; Jozwik, K.M.; Mendez, M.A.; Horder, J.; Loth, E.; et al. Atypical neurogenesis in induced pluripotent stem cells from autistic individuals. Biol. Psychiatry 2021, 89, 486–496. [Google Scholar] [CrossRef]
- Sunamura, N.; Iwashita, S.; Enomoto, K.; Kadoshima, T.; Isono, F. Loss of the Fragile X Mental Retardation Protein Causes Aberrant Differentiation in Human Neural Progenitor Cells. Sci. Rep. 2018, 8, 11585. [Google Scholar] [CrossRef]
- Brighi, C.; Salaris, F.; Soloperto, A.; Cordella, F.; Ghirga, S.; de Turris, V.; Rosito, M.; Porceddu, P.F.; D’Antoni, C.; Reggiani, A.; et al. Novel Fragile X Syndrome 2D and 3D Brain Models Based on Human Isogenic FMRP-KO IPSCs. Cell Death Dis. 2021, 12, 498. [Google Scholar] [CrossRef]
- Kurosaki, T.; Imamachi, N.; Pröschel, C.; Mitsutomi, S.; Nagao, R.; Akimitsu, N.; Maquat, L.E. Loss of the fragile X syndrome protein FMRP results in misregulation of nonsense-mediated mRNA decay. Nat. Cell Biol. 2021, 3, 40–48. [Google Scholar] [CrossRef]
- Gordon, A.; Geschwind, D.H. Human In Vitro Models for Understanding Mechanisms of Autism Spectrum Disorder. Mol. Autism 2020, 11, 26. [Google Scholar] [CrossRef] [PubMed]
- Huang, C.Y.; Liu, C.L.; Ting, C.Y.; Chiu, Y.T.; Cheng, Y.C.; Nicholson, M.W.; Hsieh, P.C.H. Human IPSC Banking: Barriers and Opportunities. J. Biomed. Sci. 2019, 26, 87. [Google Scholar] [CrossRef] [PubMed]
- Studer, L.; Vera, E.; Cornacchia, D. Programming and reprogramming cellular age in the era of induced pluripotency. Cell Stem Cell 2015, 16, 591. [Google Scholar] [CrossRef] [PubMed]
- Zhang, Y.; Xie, X.; Hu, J.; Afreen, K.S.; Zhang, C.L.; Zhuge, Q.; Yang, J. Prospects of Directly Reprogrammed Adult Human Neurons for Neurodegenerative Disease Modeling and Drug Discovery: IN vs. IPSCs Models. Front. Neurosci. 2020, 14, 546484. [Google Scholar] [CrossRef]
- Lee, C.T.; Bendriem, R.M.; Wu, W.W.; Shen, R.F. 3D brain Organoids derived from pluripotent stem cells: Promising experimental models for brain development and neurodegenerative disorders. J. Biomed. Sci. 2017, 24, 59. [Google Scholar] [CrossRef]
- Sloan, S.A.; Andersen, J.; Pașca, A.M.; Birey, F.; Pașca, S.P. Generation and assembly of human brain region-specific three-dimensional cultures. Nat. Protoc. 2018, 13, 2062–2085. [Google Scholar] [CrossRef]
- Schweitzer, J.S.; Song, B.; Herrington, T.M.; Park, T.Y.; Lee, N.; Ko, S.; Jeon, J.; Cha, Y.; Kim, K.; Li, Q.; et al. Personalized iPSC-Derived Dopamine Progenitor Cells for Parkinson’s Disease. N. Engl. J. Med. 2020, 382, 1926–1932. [Google Scholar] [CrossRef]
- Van Den Berg, A.; Mummery, C.L.; Passier, R.; Van der Meer, A.D. Personalised Organs-on-Chips: Functional Testing for Precision Medicine. Lab Chip 2019, 19, 198–205. [Google Scholar] [CrossRef]
In Vitro Model | Observation (s) | Implication (s) | References |
---|---|---|---|
The iPSCs differentiated to NPCs, glutamatergic neurons | Diminished PSD95 synaptic protein levels, altered expression of Wnt signaling pathway genes | Diminished neuronal connectivity, neurite number, altered Wnt signaling pathways, and glutamate receptors | Brennand et al., 2011 [38] |
Genetic high risk (GHR) individuals and SZ-iNSC-derived without genetic modification | The migration rate of SZ-iNSCs was significantly slower than that of HC-iNSCs | Migration capacity was impaired in SZ-iNSCs compared to iNSCs from GHR individuals or controls. iNSCs from a GHR individual who later developed SZ showed migratory impairment similar to SZ-iNSCs. | Lee et al., 2022 [39] |
The iPSCs differentiated to neuron committed cells (NCCs) | Altered nFGFR1 signaling which integrates diverse pathways with schizophrenia-linked mutations | Global dysregulation of developmental genome | Narla et al., 2017 [40] |
The iPSC-derived DG hippocampal neuron | Deficits in the generation of DG granule neurons | Reduced neuronal activity and spontaneous neurotransmitter release | Yu et al., 2014 [41] |
The iPSCs, differentiated to CA3 neurons and DG neurons | Reduced activity in DG-CA3 co-culture and deficits in spontaneous and evoked activity in CA3 neurons | iPSC-derived DG-CA3 co-cultures present deficits in hippocampal activity | Sarkar et al., 2018 [42] |
The iPSC-derived neurons neural progenitor cells (NPCs). | Abnormal gene expression and protein levels related to cytoskeletal remodeling and oxidative stress | Aberrant migration and increased oxidative stress | Brennand et al., 2015 [43] |
The NPCs derived from SZ-iPSCs and HC | The embryo exposure to environmental factors activates HSF1 in cerebral cortical cells. SZ-NPCs showed higher variability in the levels of HSF1 activation than HC | HSF1 plays a role in the response of brain cells to prenatal environmental insults | Hashimoto-Torii et al., 2015 [44] |
Embryonic stem cell (hESC) and iPSC-derived BO | The nFGFR1, that controls the brain development by integrating signals from diverse development–initiating factors, was lost from the nuclei of differentiating cortical neurons in SZ-BO | The loss of nFGFR1 likely underlies the SZ impaired cortical neuronal differentiation | Stachowiak et al., 2017 [45] |
Embryonic stem cells (ESC), SZ- and HC-iPSC-derived BO | In HC-BO iPSC the transient exposition to TNF produced malformations similar to those of the SZ-BO. Both SZ- and TNF-induced malformations were associated with the loss of nFGFR1 in the cortical zone | Maternal infection during pregnancy exposes the fetus to elevated TNF levels. The TNF, similar to SZ, alters nFGFR1 signaling, NPCs, and neurons in developing BO | Benson et al., 2020 [46] |
An SZ-iPSC and HC-iPSC derived BO | Downregulation of genes involved in cell adhesion, neurodevelopment, and synaptic biology along with the upregulation of genes involved in immune signaling in SZ-BO compared to HC-BO | The SZ-BO showed differences in expression of genes involved in mitochondrial function, excitation/inhibition balance modulation, synapse biology, neurodevelopment, and immune response | Kathuria et al., 2020 [47] |
An SZ-iPSC and HC-iPSC derived BO | About 2.62% of global proteome was differentially regulated in SZ organoids (43 proteins up-regulated and 54 down-regulated) with depletion of factors that support neuronal development and differentiation | Key pathways regulating nervous system development was perturbed in SZ-derived organoids | Notaras et al., 2021 [48] |
An SZ-iPSC and HC-iPSC derived BO | The SZ organoids exhibited progenitors depleted of neuronal programming factors leading to altered developing cortical areas | Multiple mechanisms in SZ-derived BO converge upon brain developmental pathways and contribute to raising the risk of developing SZ | Notaras et al., 2022 [49] |
The iPSC-derived neural progenitors and cortical neurons | Electrophysiological measures in iPSC-derived neurons showed altered Na+ channel function, action potential interspike interval, and GABAergic neurotransmission | Electrophysiological measures predicted cardinal clinical and cognitive features | Page et al., 2022 [50] |
iPSC Model (s) | Observation (s) | Implication (s) | References |
---|---|---|---|
The iPSCs from SSRI-remitters and differentiated to serotonergic neurons | The SSRI-non-remitters patient-derived neurons displayed 5-HT-induced hyperactivity via upregulated 5-HT2A and 5-HT7 receptors | The SSRI-resistant patients represent a subset of MDD patients with differential responses to 5-HT | Vadodaria et al., 2019 [57] |
The iPSCs from SSRI-remitters and SSRI-non-remitters differentiated to serotonergic neurons | Altered morphological phenotypes associated with SSRI-non-remission downstream of PCDHA6 and PCDHA8 genes | The differences in serotonergic neuron morphology may contribute to SSRI-non-remission in MDD patients | Vadodaria et al. [58] |
The MDD BP-responders’ and- non-responders’ LCLs, reprogrammed to iPSCs, and differentiated to mature prefrontal cortex neurons | Following BP treatment, cortical neurons of BP-responders, compared with vehicle treatment, had significant differences, including enhanced colocalization of synaptic markers and spine morphology | The enhanced colocalization of synaptic markers observed in BP-responders’ derived cells could be indicative of enhanced connectivity following effective antidepressant treatment. The synaptic changes could be used as a biomarker for BP effects in vitro | Avior et al., 2021 [59] |
The iPSC-derived dopaminergic neurons | Ketamine elicits structural plasticity in iPSCs-derived DA neurons by recruitment of AMPA, mTOR and BDNF signaling | The transient exposure to ketamine dose-dependently promotes structural plasticity as determined by enhanced dendritic outgrowth and increased soma size | Cavalleri et al., 2018 [60] |
The iPSCs from healthy donors differentiated to dopaminergic neurons | Exposure to the ketamine metabolite HNK for 6 h produces AMPA receptor-mediated and mTOR-mediated dendrite outgrowth when measured 3 days after exposure | The prolonged antidepressant. Effect after a single infusion of ketamine is based on the triggering of long-lasting neuroplasticity | Collo et al., 2018 [61] |
iPSC Model (s) | Observation (s) | Implication (s) | References |
---|---|---|---|
The iPSCs differentiated to neurons | Increased expression of transcripts for membrane-bound receptors and ion channels in BD-derived neurons than in controls | Neurons from BD iPSC were significantly different in their gene expression, particularly transcripts involved in calcium signaling, from those derived from control iPSC | Chen et al., 2014 [63] |
The iPSCs differentiated to neurons | Lithium alters the phosphorylation state of collapsin response mediator protein-2 (CRMP2) | The molecular lithium-response pathway in BD acts via CRMP2 to alter neuronal cytoskeletal dynamics, mainly dendrite and dendritic spine formation/function | Tobe et al., 2017 [64] |
The iPSC-derived NPCs | The miR-34a is upregulated and targets BD risk genes ANK3, CACNB3, and DDN | The miR-34a overexpression impairs neuronal differentiation, expression of synaptic proteins and neuronal morphology | Bavamian et al., 2015 [65] |
The iPSCs from patients who carried CNVs differentiated to neurons | Exonic deletion of Protocadherin 15 (PCDH15) is associated with shorter dendrites and decreased number of synapses than controls in both glutamatergic and GABAergic neurons | Neurons derived from iPSCs exhibited dendrite shortening and decreased synapse numbers | Ishi et al., 2019 [66] |
The CXCR4 expressing NPCs | Phenotypic differences at the level of neurogenesis and expression of genes critical for neuroplasticity, including WNT pathway components and ion channels | Abnormalities in early steps in NPC formation, WNT/ GSK3 signaling and expression of ion channels in the BD patient-derived NPCs and neurons | Madison et al., 2015 [67] |
The iPSCs differentiated to NPCs and then to neurons from Old Order Amish of Lancaster County | A total of 328 genes were differentially expressed between BPD and control L neurons including GAD1, glutamate decarboxylase 1, and SCN4B, the voltage gated type IV sodium channel beta subunit | The differentially expressed genes suggested that the alterations in RNA metabolic processes, protein trafficking, and receptor-mediated signaling contribute to BD biology | Kim et al., 2015 [68] |
The GPCs from BD- and HC-iPSCs | The BD astrocytes induced a reduction in neuronal activity when co-cultured with neurons, mainly after cytokine stimulation | The BD astrocytes are functionally less supportive of neuronal activity and this effect is partially mediated by cytokines | Vadodoria et al., 2021 [69] |
The iPSCs differentiated to hippocampal DG-like neurons | Using both patch-clamp recording and somatic Ca2+ imaging, hyperactive action-potential firing was observed | The hyperexcitability phenotype of BD neurons was reversed by lithium in neurons derived from patients who also responded to lithium treatment | Mertens et al., 2015 [70] |
The iPSC-derived NPCs | Gene expression signature associated with the three most commonly used mood stabilizers (Li, VPA, and LTG) | The findings support the involvement of mitochondrial functions in the molecular mechanisms of mood stabilizers | Osere et al., 2021 [71] |
The BD-iPSCs derived from EBV immortalized lymphocytes differentiated to DG granule neurons | Endophenotype of hyperexcitability is shared by BD DG neurons. Functional analysis showed that intrinsic cell parameters are very different between neurons derived from Li-responders and those derived from Li-non-responders | Chronic Li treatment reduced the hyperexcitability in the lymphoblast-derived Li-responders but not in the Li-non-responders | Stern et al., 2018 [72] |
iPSC Model (s) | Observation (s) | Implication (s) | References |
---|---|---|---|
The iPSCs differentiated to NPCs and neurons from ASD individuals with early brain overgrowth and non-ASD controls with normal brain size | The ASD-derived NPCs display faster proliferation than control-derived NPCs due to dysregulation of a β-catenin/BRN2 transcriptional cascade | Abnormal neurogenesis and synaptogenesis leading to functional defects in neuronal networks that could be rescued by the neurotrophic factor IGF-1 | Marchetto et al. [76] |
The iPSCs from ASD with macrocephaly differentiated to NSCs; iPSCs directly converted into iNs; generation of cerebral organoids | Dysregulation of specific transcriptional networks that caused aberrant neuronal maturation of ASD cortical neurons | The ASD-associated neurodevelopmental aberrations are triggered by a pathological priming of gene regulatory networks during early neural development | Shafer et al., 2019 [77] |
Organoids derived from ASD-iPSCs | The ASD-derived organoids exhibit an overproduction of GABAergic inhibitory neurons, due to the overexpression of the transcription factor FOXG1 | Cortical organoids of ASD patients show exuberant GABAergic differentiation and no change in glutamate neuron types, which together cause an imbalance in glutamate/GABA neuron ratio | Mariani et al., 2016 [78] |
The iPSCs from ASD without macrocephaly differentiated to cortical and midbrain neurons | The ASD-iPSCs differentiated to cortical neurons displayed impaired neural differentiation. These cellular phenotypes occurred in the absence of alterations in cell proliferation during cortical differentiation, in contrast to previous studies | Patients with ASD but without macrocephaly exhibited impairments in neurogenesis compared with those from neurotypical individuals | Adhya et al., 2021 [79] |
The NPCs derived from FMR1-knockout iPSCs as a model for studying FMRP functions and FXS pathology | Altered expression of neural differentiation markers, MRP-deficient neurons showed less spontaneous calcium bursts, corrected by the protein kinase inhibitor LX7101 | Loss of FMRP resulted in abnormal differentiation accompanied by impaired neuronal activity | Sunamura et al. [80] |
Both 2D and 3D FXS models based on isogenic FMR1 knock-out mutant and wild-type human iPSC lines | Cortical neurons derived from FMRP-deficient iPSCs exhibit altered gene expression and impaired differentiation when compared with the healthy counterpart | The FMRP is required to correctly support neuronal and glial cell proliferation, and to set the correct excitation/inhibition ratio in the developing brain | Brighi et al. [81] |
Human SH-SY5Y neuroblastoma cells and FXS fibroblast-derived iPSCs | The FMRP deficiency results in hyperactivated nonsense-mediated mRNA decay (NMD). The key NMD factor UPF1 binds directly to FMRP, promoting FMRP binding to NMD targets | The FMRP acts as an NMD repressor. In the absence of FMRP, NMD targets are relieved from FMRP-mediated repression. Many abnormalities in FMRP-deficient cells are attributable, either directly or indirectly, to misregulated NMD | Kurosaky et al. [82] |
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations. |
© 2022 by the authors. Licensee MDPI, Basel, Switzerland. This article is an open access article distributed under the terms and conditions of the Creative Commons Attribution (CC BY) license (https://creativecommons.org/licenses/by/4.0/).
Share and Cite
Alciati, A.; Reggiani, A.; Caldirola, D.; Perna, G. Human-Induced Pluripotent Stem Cell Technology: Toward the Future of Personalized Psychiatry. J. Pers. Med. 2022, 12, 1340. https://doi.org/10.3390/jpm12081340
Alciati A, Reggiani A, Caldirola D, Perna G. Human-Induced Pluripotent Stem Cell Technology: Toward the Future of Personalized Psychiatry. Journal of Personalized Medicine. 2022; 12(8):1340. https://doi.org/10.3390/jpm12081340
Chicago/Turabian StyleAlciati, Alessandra, Angelo Reggiani, Daniela Caldirola, and Giampaolo Perna. 2022. "Human-Induced Pluripotent Stem Cell Technology: Toward the Future of Personalized Psychiatry" Journal of Personalized Medicine 12, no. 8: 1340. https://doi.org/10.3390/jpm12081340
APA StyleAlciati, A., Reggiani, A., Caldirola, D., & Perna, G. (2022). Human-Induced Pluripotent Stem Cell Technology: Toward the Future of Personalized Psychiatry. Journal of Personalized Medicine, 12(8), 1340. https://doi.org/10.3390/jpm12081340