Quick viewing(Text Mode)

Mutations in Thyroid Hormone Receptor Α1 Cause Premature Neurogenesis and Progenitor Cell Depletion in Human Cortical Development,” by Teresa G

Mutations in Thyroid Hormone Receptor Α1 Cause Premature Neurogenesis and Progenitor Cell Depletion in Human Cortical Development,” by Teresa G

Correction

MEDICAL SCIENCES Correction for “ in hormone α1 cause premature neurogenesis and progenitor cell depletion in human cortical development,” by Teresa G. Krieger, Carla M. Moran, Alberto Frangini, W. Edward Visser, Erik Schoenmakers, Francesco Muntoni, Chris A. Clark, David Gadian, Wui K. Chong, Adam Kuczynski, Mehul Dattani, Greta Lyons, Alexandra Efthymiadou, Faraneh Varga-Khadem, Benjamin D. Simons, Krishna Chatterjee, and Frederick J. Livesey, which was first published October 18, 2019; 10.1073/pnas.1908762116 (Proc. Natl. Acad. Sci. U.S.A. 116,22754–22763). The authors note that “In Fig. 4D, the horizontal line that highlights the comparison between neuronal activity in the three control lines and one of the three THRA mutant lines was drawn to end above the TRα1-FS382 , rather than the TRα1- FS397 mutation. The placement of the horizontal line incorrectly indicated that neuronal activity in TRα1-FS382 mutation neurons was significantly lower than that of controls, whereas activity in TRα1-FS397 neurons was not. The figure has been amended, and now the horizontal line correctly indicates that neuronal activity in CORRECTION TRα1-FS397 neurons was significantly lower than that of controls, whereas activity in TRα1-FS382 neurons was not. We apologize for any confusion this labeling error may have caused.” The cor- rected Fig. 4 and its legend appear below. The authors also note that the author name Faraneh Varga- Khadem should instead appear as Faraneh Vargha-Khadem. The corrected author line appears below. The online version has been corrected.

Teresa G. Krieger, Carla M. Moran, Alberto Frangini, W. Edward Visser, Erik Schoenmakers, Francesco Muntoni, Chris A. Clark, David Gadian, Wui K. Chong, Adam Kuczynski, Mehul Dattani, Greta Lyons, Alexandra Efthymiadou, Faraneh Vargha-Khadem, Benjamin D. Simons, Krishna Chatterjee, and Frederick J. Livesey

www.pnas.org PNAS | March 31, 2020 | vol. 117 | no. 13 | 7537–7538 Downloaded by guest on September 27, 2021 Fig. 4. FACS-enriched THRA mutation-containing cortical progenitor cells generate functional neuronal networks. (A) Representative FACS results for WT and THRA mutation-containing cells at day 18. Immunocytochemistry confirmed the cortical identity of FACS- selected progenitor cells. (Scale bars: 50 μm.) (B) Neurons produced by FACS-purified progenitor cells expressed the cortical layer-specific factors TBR1 (layer VI) and CTIP2 (layer V), as well as the vesicular glutamate transporter vGlut1, at day 40. (Scale bars: 50 μm.) (C) Representative traces showing action potential firing in response to stepwise current injection in THRA mutation-containing and control neurons. Numbers indicate the proportion of cells that fired action potentials. (D) Calcium indicator Oregon Green BAPTA was used as a proxy for action potential firing to measure spontaneous neuronal activity. Fluorescence images show control and THRA

mutant cultures after loading. Heat maps highlight different levels of activity across the cultures. Total calcium activity across the field of view, ACa, was quantified based on 3–6 videos from each cell line (403X and FS382, n = 6 samples from 2 independent inductions; all others, n = 3 samples from 1 induction; *P < 0.05, Student’s t test). Error bars indicate SEM.

Published under the PNAS license. First published March 23, 2020.

www.pnas.org/cgi/doi/10.1073/pnas.2003767117

7538 | www.pnas.org Downloaded by guest on September 27, 2021 Mutations in thyroid α1 cause premature neurogenesis and progenitor cell depletion in human cortical development

Teresa G. Kriegera,b,1, Carla M. Moranc, Alberto Franginia, W. Edward Visserc,2, Erik Schoenmakersc, Francesco Muntonid, Chris A. Clarke, David Gadiane, Wui K. Chongf, Adam Kuczynskig, Mehul Dattanih, Greta Lyonsc, Alexandra Efthymiadoui, Faraneh Varga-Khademg,j, Benjamin D. Simonsa,b, Krishna Chatterjeec, and Frederick J. Liveseya,k,3 aGurdon Institute, University of Cambridge, Cambridge CB2 1QN, United Kingdom; bCavendish Laboratory, University of Cambridge, Cambridge CB3 0HE, United Kingdom; cWellcome Trust-MRC Institute of Metabolic Science, University of Cambridge, Cambridge CB2 0QQ, United Kingdom; dDubowitz Neuromuscular Centre and National Institute for Health Research (NIHR) Great Ormond Street (GOS) Hospital Biomedical Research Centre, London WC1N 1EH, United Kingdom; eDevelopmental Imaging and Biophysics Section, University College London (UCL) GOS Institute of Child Health, London WC1N 1EH, United Kingdom; fDepartment of Radiology, Great Ormond Street Children’s Hospital, London WC1N 3JH, United Kingdom; gDepartment of Neuropsychology, Great Ormond Street Children’s Hospital, London WC1N 1EH, United Kingdom; hDepartment of Endocrinology, Great Ormond Street Children’s Hospital and Genetics and Genomic Medicine Programme, UCL GOS Institute of Child Health, London WC1N 1EH, United Kingdom; iDepartment of Endocrinology, University of Ioannina, 45110 Ioannina, Greece; jCognitive Neuroscience and Neuropsychiatry Section, UCL GOS Institute of Child Health, London WC1N 1EH, United Kingdom; and kUCL Great Ormond Street Institute of Child Health, London WC1N 1EH, United Kingdom

Edited by Janet Rossant, Hospital for Sick Children, University of Toronto, Toronto, ON, Canada, and approved September 17, 2019 (received for review June 5, 2019) Mutations in the thyroid hormone receptor α 1gene(THRA)have the complex together with recruitment recently been identified as a cause of intellectual deficit in humans. and transcriptional activation (11, 12). Patients present with structural abnormalities including microencephaly, We reported the first human THRA mutation in 2012 (13), reduced cerebellar volume and decreased axonal density. Here, we after which approximately 29 other patients have been identified show that directed differentiation of THRA mutant patient-derived in- with shared phenotypic features defining the disorder resistance MEDICAL SCIENCES duced pluripotent stem cells to forebrain neural progenitors is mark- to thyroid hormone α (RTHα) (14–18). All the patients carry edly reduced, but mutant progenitor cells can generate deep and heterozygous missense or truncating mutations in the - upper cortical layer neurons and form functional neuronal net- binding domain of TRα1 that disrupt its ability to bind T3, works. Quantitative lineage tracing shows that THRA mutation- impairing corepressor dissociation and coactivator recruitment containing progenitor cells exit the cell cycle prematurely, resulting (13, 16). When coexpressed, mutant TRα1 inhibits the function in reduced clonal output. Using a micropatterned chip assay, we find of its wild-type (WT) counterpart in a dominant-negative manner that spatial self-organization of mutation-containing progenitor cells (13). In addition to growth retardation and skeletal dysplasia, in vitro is impaired, consistent with down-regulated expression of – cell cell adhesion genes. These results reveal that thyroid hormone Significance receptor α1 is required for normal neural progenitor cell proliferation in human cerebral cortical development. They also exemplify quan- Thyroid hormone deficiencies are the most common preventable titative approaches for studying neurodevelopmental disorders us- causes of intellectual disability. We report that mutations in the ing patient-derived cells in vitro. thyroid hormone receptor α1gene(THRA) that result in in- tellectual disability also reduce brain size. Using human THRA thyroid hormone | brain development | iPSCs mutation stem cell models, we studied the impact of THRA mu- tations on human brain development by combining quantitative he human cerebral cortex mediates higher cognitive and lineage analysis, gene expression analyses, and novel assays of Tsensorimotor functions, with thyroid hormone (TH) deficiency neuroepithelium formation. We found that THRA regulates the during pregnancy or the neonatal period recognized as the most balance between progenitor self-renewal and neurogenesis, and common preventable cause of intellectual disability worldwide (1). thus overall brain size. Importantly, these in vitro results are Defects in progenitor cell proliferation, synaptogenesis, and den- consistent with in vivo evidence from magnetic resonance imag- dritic arborization, neuronal migration, and cell survival have been ing of people with these mutations, advancing our understanding observed in the cerebral cortex of the progeny of hypothyroid of thyroid hormone action in human brain development. rodents (2–4). Aberrant behavior and cortical cytoarchitecture are Author contributions: T.G.K., B.D.S., and F.J.L. designed research; T.G.K., C.M.M., A.F., observed even following transient TH deficiency during the first W.E.V., E.S., F.M., C.A.C., D.G., W.K.C., A.K., G.L., A.E., F.V.-K., and K.C. performed re- half of gestation, emphasizing the critical role of THs in early search; T.G.K., C.M.M., F.M., C.A.C., D.G., M.D., and B.D.S. analyzed data; and T.G.K., brain development (5). However, in humans, the actions of THs C.M.M., K.C., and F.J.L. wrote the paper. on cells of the central nervous system (CNS) remain poorly de- The authors declare no competing interest. fined (6). In the absence of appropriate in vitro models, it has This article is a PNAS Direct Submission. been difficult to study TH action in specific cells or tissues sepa- This open access article is distributed under Creative Commons Attribution License 4.0 rate from its global effects, which are likely mediated by a range of (CC BY). 1 tissues and cell types (7). Present address: Digital Health Center, Berlin Institute of Health (BIH)/Charité-Univer- sitätsmedizin Berlin, 13353 Berlin, Germany. During cerebral cortex development, THs (thyroxine, T4; tri- 2 α Present address: Department of Internal Medicine, Academic Center for Thyroid Dis- iodothyronine, T3) act via a (TR 1) encoded by eases, Erasmus Medical Center, 3015 Rotterdam, The Netherlands. THRA the gene, to regulate transcription of target genes in a 3To whom correspondence may be addressed. Email: [email protected]. – ligand-dependent manner (8 10). Unliganded TH receptors This article contains supporting information online at www.pnas.org/lookup/suppl/doi:10. (TRs) recruit a corepressor complex to inhibit target gene tran- 1073/pnas.1908762116/-/DCSupplemental. scription (11); hormone (T3) occupancy promotes dissociation of www.pnas.org/cgi/doi/10.1073/pnas.1908762116 PNAS Latest Articles | 1of10 patients with RTHα exhibit mild-to-moderate intellectual dis- ability, notably affecting nonverbal IQ and sensorimotor process- A ing, and 1 adult woman has experienced epileptic seizures that began in infancy (16). These findings suggest a crucial role for TRα1 in human cortical neurogenesis, consistent with previous studies reporting a range of CNS abnormalities in mice mutant for TRα1 (19). However, the cellular mechanisms underlying aberrant neural development in patients with RTHα remain unknown. Here we have delineated the neurologic and neurocognitive phenotypes and undertaken structural (magnetic resonance im- aging [MRI], tractography) neuroimaging and proton magnetic resonance spectroscopy (MRS) in the first 4 RTHα patients reported, harboring frameshift/premature stop THRA mutations that are representative of the type of receptor defect found in 50% of the worldwide RTHα cohort (20). We directed differ- B entiation of THRA mutant patient-derived induced pluripotent stem cells (iPSCs) to a cortical excitatory neuronal fate, using an established in vitro system that recapitulates development from early neuroepithelium to functional neuronal circuits (21, 22). Based on quantitative analysis of lineage tracing data, we found that THRA mutation-containing cortical progenitor cells are bi- ased toward early differentiation, leading to premature neuro- genesis and depletion of the progenitor cell pool. They also exhibit impaired self-organization into cortical rosette-like struc- tures in vitro. Defects in neural progenitor proliferation, cell po- larity, and apical adhesion may thus contribute to the structural abnormalities and to the sensorimotor and neurocognitive phe- notypes seen in patients with RTHα. Results Neurologic, Neurocognitive, and Neuroimaging Abnormalities in Patients with THRA Mutation. We assessed neurologic, neurocognitive, and C neuroimaging phenotypes in the first 4 RTHα cases reported (SI D Appendix,TableS1), all harboring mutations that disrupt the carboxyterminal alpha helix of TRα1(Fig.1A).Thesubjectsarea9-y- old female (E403X TRα1 mutation; referred to as P1 below), a 13-y-old female and her 49-y-old father (F397fs406X TRα1 mutation; P2 and P3), and a 47-y-old female (A382PfsX7 TRα1mutation; P4). All 4 patients exhibited delayed developmental milestones (SI Appendix,TableS2) and neurologic abnormalities, including slow initiation of movement, ataxic gait, dysarthria, and fine and gross motor incoordination (SI Appendix,TableS3). Neuropsychological examinations showed significantly reduced nonverbal IQ in all cases, with scores ranging from 2 (P1, P2, and P3) to 3.4 (P4) SDs below the population mean (X = 100; SD, 1.5). Furthermore, all patients showed severe impairments in motor coordination, visual motor integration, and finger dexterity of both the dominant and nondominant hands. P4 showed intellectual disability affecting verbal as well as nonverbal abilities, whereas verbal abilities were relatively preserved in P1, P2, and P3. Per- formance on a test of visual perception was within the low average to average range in P1-P3 (SI Appendix, Table S4). Magnetic resonance imaging (MRI) revealed reduced cere- Fig. 1. THRA mutations are associated with structural abnormalities in A α bellar volumes in all 4 patients (Fig. 1B and SI Appendix, Fig. the brain. ( ) Structural modeling of the ligand-binding domain of TR 1 showing the position of mutations (red). The mutations in patients P4 S1), with microencephaly in the adult cases (P3 and P4), despite (A382PfsX7), P2 and P3 (F397fs406X), and P1 (E403X) disrupt or truncate the their increased skull volume (macrocephaly) denoted by known + carboxyterminal alpha helix (H12; black) of the receptor, exposing a hy- increased head circumference (P1, 3 SDs from the population drophobic cleft that facilitates corepressor (CoR; blue) binding by unliganded mean; P2, +4 SD; P3, +5 SD; P4, +9 SD) (Fig. 1C). Diffusion receptor (Left) and removing or changing amino acids required for T3 (cyan) tensor imaging (DTI) performed in P1 and P2 showed a global binding and coactivator (green) recruitment (Right). (B) MRI scans of patients increase in the mean diffusivity of water in white matter tracts, P1 and P2 and a control subject (female, age 10 y 8 mo) with T2-weighted denoting a reduction in their density (Fig. 1D). On MRS, there axial images (Top) and T1-weighted coronal images (Bottom), showing in- was a reduced ratio of N-acetylaspartate to the creatine plus creased CSF space (arrows) around the cerebellum and between folia, C phosphocreatine ratio in the frontal white matter and thalamus denoting reduced cerebellar size. ( ) Sagittal images from MRI brain scans of SI adult cases P3 and P4 and a control subject (female, age 52 y) showing in P2, P3, and P4, suggestive of neuronal loss or dysfunction ( D Appendix microencephaly. ( ) Tract-based spatial statistics analysis of DTI data in pa- , Table S5). These results suggest that the observed tients P1 and P2. Tracts highlighted in blue signify greater mean diffusivity neurocognitive and sensorimotor deficits in the patients are re- (MD) of water than in controls (n = 20 age- and sex-matched subjects), and lated to structural abnormalities in several brain regions, in- tracts highlighted in green denote not significantly different MD compared cluding the cerebellum and cerebral cortex. with controls.

2of10 | www.pnas.org/cgi/doi/10.1073/pnas.1908762116 Krieger et al. Reduced Differentiation of THRA Mutant iPSCs to Forebrain that were detected above negligible levels in all samples, we con- Neuroepithelial Cells. To study the cellular developmental origin sidered 478 transcripts whose expression differed significantly be- of cortical abnormalities in THRA mutant patients, iPSC lines tween THRA mutation-containing and control lines (false discovery were derived from P1, P2, and P4 by reprogramming with non- rate <0.25, independent of the sign of the fold change). integrating Sendai virus (SI Appendix, Fig. S2A) (23). THRA Hierarchical clustering analysis revealed that the expression mutant and control iPSCs were then differentiated to cerebral profiles of replicate inductions from each THRA mutation- cortex neural progenitor cells using an established protocol containing line clustered together, suggesting that the cortical based on dual SMAD inhibition (Fig. 2A) (21, 22). After 12 d of induction process in vitro preserves mutation-specific differences induction, a neuroepithelial sheet was formed in control cultures, in transcription (Fig. 3A). The differentially expressed genes and expression of PAX6 and FOXG1 mRNA confirmed cortical clustered into 4 distinct groups, with the largest subset repre- SI Appendix B identity ( , Fig. S2 ). Neural progenitor cells pro- senting genes whose expression is down-regulated in THRA ceeded to generate neurons as described previously (22). In mutation-containing vs. control cells, consistent with the known control cultures, TRα1 was expressed in stem cells before cor- + repressive function of TRα1 (209 genes in group C). Gene On- tical induction, in cycling (Ki67 ) progenitor cells, and in post- + B SI Appendix C tology (GO) analysis showed that many of these down-regulated mitotic (Tuj1 ) neurons (Fig. 2 and , Fig. S2 ), genes relate to DNA-templated transcription, ectoderm and nervous confirming that the receptor is present throughout the in vitro system development, or cell–cell adhesion via membrane adhesion differentiation process. molecules (Fig. 3B), with substantial overlap (Fig. 3C). Genes When THRA mutation-containing iPSCs from the 3 patients relating to both transcription and neural development include were subjected to the same differentiation protocol, cortical in- bHLH-transcription factors such as NEUROD1, NEUROD4 duction was found to be reduced in efficiency. While cortical (MATH3), and NEUROG2, which are involved in normal cere- transcription factors were also expressed in THRA mutant cultures bral cortex development (24), as well as LHX2, which is known (SI Appendix,Fig.S2B), expression of PAX6 was highly variable SI Appendix D to regulate the specification of cortical regional fates (25). The between cells ( ,Fig.S2 ). At day 15, following manual – fragmentation of the neuroepithelial sheet, PAX6 expression in genes involved in both neural development and cell cell adhe- THRA C D sion are 6 protocadherins (part of the cadherin family), the autism mutation-containing cells was decreased (Fig. 2 and ). NRXN1 NPTN PAX6 expression was mostly higher in, but not confined to, areas susceptibility gene (neurexin), and (neuroplastin), with rosette-like morphology (Fig. 2D). which has been associated with cortical thickness and intellectual ability in humans (26, 27). Altered Expression of Neural Development and Cell–Cell Adhesion Thus, our RNA-seq data show that THRA mutations are as- MEDICAL SCIENCES Genes during Neural Differentiation of THRA Mutant iPSCs. To in- sociated with the down-regulation of genes involved in neural vestigate differences in induction efficiency in an unbiased manner, development. In addition, the data suggest compromised cell– we performed RNA sequencing of 3 independent neural induc- cell adhesion in THRA mutation-containing cells, possibly con- tions from each THRA mutation-containing iPSC line (TRα1- tributing to impaired cortical growth. Notably, within up-regulated 403X carrying the E403X mutation, TRα1-FS382 with the (group B) genes, enrichment for nonneural differentiation fates A382PfsX7 mutation, and TRα1-FS397 with the F397fs406X mu- was not detected, suggesting that the reduced efficiency of neural tation) and 1 neural induction from each control line (H9, NAS6, induction seen in THRA mutation-containing cells is not due to and NDC), at day 12 of cortical induction. Out of 16,773 transcripts differentiation down nonneural paths.

Neuro- Cortical Deep-layer neurons ABprogenitor Upper-layer neurons

iPSCs epithelium DAPI rosettes Ki67 10-50 12 days 4 days days TRα1 Dual SMAD FGF2

in vitro inhibition + retinoids DAPI Tuj1 Tuj1 in vivo

apical side TRα1

C Control TRα1-403X TRα1-FS397 TRα1-FS382 D * * Phase DAPI Pax6/DAPI expression

Pax6 ()*+,-./0*-11-*Inside 231*+,-./0*-11-*Outside rosettes rosettes

Fig. 2. Inefficient cortical induction of THRA mutation-containing iPSCs. (A) Schematic of the in vitro cortical differentiation protocol. (B)TRα1 is expressed in cycling Ki67+ progenitors as well as postmitotic Tuj1+ neurons derived from control iPSCs. (Scale bars: 50 μm.) (C) Representative pictures showing reduced progenitor rosette formation and Pax6 expression in THRA mutation-containing cells compared with control cultures at day 15. At least 3 independent differentiations were performed for 1 clone from each patient. (Scale bars: 50 μm.) (D) Quantification of Pax6 expression within and outside of neural progenitor cell rosettes in THRA mutant and control cell cultures at day 15 of neural induction. Rosettes were manually delineated based on morphology in 8 or 9 images from 2 or 3 independent inductions for each cell line, and Pax6 expression was measured relative to DAPI (Methods). Error bars indicate SEM. *P < 0.001, 2-sided Student’s t test comparing a total of nine control and 24 THRA mutant images.

Krieger et al. PNAS Latest Articles | 3of10 A

BC

Fig. 3. RNA-seq at day 12 of in vitro cortical induction reveals down-regulation of transcription, neurogenesis, and cell–cell adhesion in THRA mutation- containing cells. (A) 2D hierarchical clustering of log2(RPKM) values for 3 inductions each from control iPSCs and the 3 patient lines. Colors correspond to relative expression in each row, from lowest (blue) to highest (red). Inductions cluster by cell line of origin (vertical axis), while genes separate into 4 dif- ferentially expressed groups (horizontal axis). (B)FDRq values of GO terms enriched among group C genes. (C) GO analysis reveals that genes implicated in nervous system development that are down-regulated in THRA mutation-containing cells partially overlap with cell–cell adhesion genes (8/42; 19%) and DNA- templated transcription (10/42; 24%).

CD271−CD44− Cortical Progenitor Cells Generate Functional Neurons stimulation (SI Appendix,Fig.S3A). Spontaneous miniature excit- In Vitro. To compare neural development from THRA mutation- atory postsynaptic currents were observed in a similar proportion of containing versus control cortical progenitors while controlling THRA mutation-containing and control neurons (SI Appendix,Fig. for differences in induction efficiency, cultures were enriched for S3A). The THRA mutation-containing neurons appeared more neural stem cells by FACS at day 16 after neural induction. mature than controls, with a greater proportion of cells firing action Based on an established neural stem cell signature (28), the potentials (10/11 in FS397 vs. 8/14 in H9; P < 0.1, 2-proportion Z- − − CD271 CD44 population was isolated and propagated (Fig. test), more action potentials fired per burst (mean, 7.3 ± 2.5 in 4A). CD271 is expressed in neural crest precursors (29, 30), FS397 vs. 1.6 ± 1.0 in H9; P < 0.05, Student’s t test), larger mean which are occasionally observed as a contaminant in control peak sodium currents (mean, −398 ± 66 in FS397 vs. −225 ± 48 in cortical inductions in vitro and identified based on their char- H9; P < 0.05, Student’s t test), and a lower average membrane acteristic morphology as more prominent in THRA mutant in- potential (−54 ± 2mVinFS397vs.−48 ± 3mVinH9;P = 0.12, ductions. Negative selection for CD44, a pluripotent stem cell Student’s t test). These differences might represent variations in marker (28), eliminated cells that had failed induction to neural individual neuronal properties but could also be due to earlier pro- stem cell fate. Depending on induction efficiency, approximately duction of neurons in THRA mutation-containing cultures. 50% to 80% of induced THRA mutation-containing cells were By day 50 postinduction, both control and THRA mutation- FACS-selected for propagation (Fig. 4A). FACS-enriched cul- containing cells developed spontaneous neuronal network ac- tures expressed PAX6 and the intermediate filament protein tivity, visualized by calcium imaging (Fig. 4D and SI Appendix, vimentin, both of which are neural progenitor cell markers (Fig. Fig. S3B). Spontaneous activity was blocked by the sodium 4A and SI Appendix, Fig. S2F). channel blocker tetrodotoxin and substantially reduced by the Over time, FACS-sorted progenitors gave rise to neurons AMPA receptor antagonist CNQX, indicating that a significant expressing the transcription factors TBR1 and CTIP2, expressed element of the observed spontaneous neuronal activity is medi- by deep cortical layer neurons (Fig. 4B). Most neurons also ated by excitatory synaptic transmission (SI Appendix, Fig. S3C). showed punctate vesicular glutamate transporter (VGLUT1) ex- pression, indicating their glutamatergic identity (Fig. 4B). Overall, Compromised Cortical Progenitor Cell Dynamics in THRA Mutation- FACS successfully enriched for cortical neural progenitors derived Containing Cells. The microencephaly and reduced white matter from THRA mutation-containing iPSCs, which proceeded to tract density seen in patients with RTHα suggestthatstruc- generate cortical excitatory neurons over time in vitro. tural differences in the cerebral cortex might contribute to their To assess neuronal function, 60-d-old THRA mutation-containing neurocognitive phenotype. With human cortex comprising col- and control cultures were analyzed by single-cell electrophysiology. umns of clonally related excitatory neurons preferentially synaps- The THRA mutation-containing neurons fired action potentials ing with one another (31), the clonal output of a cortical progenitor normally in response to stepwise current injections (Fig. 4C)and cell is a major contributor to cortical size, architecture, and function. displayed sodium and potassium currents in response to voltage To determine whether defects in early cortical development could

4of10 | www.pnas.org/cgi/doi/10.1073/pnas.1908762116 Krieger et al. AB Control TRα1-403X Control TRα1-403X TRα1-FS397 TRα1-FS382 4 4 10 10

0.5% 0.1% 15.2% 0.2% 3 3

10 10 DAPI

2 2 10 98.6% 10 79.6% MAP2

1 1 10 10 0.9% 5.0% CD271-PE Tbr1 0 0 10 10 0 1 2 3 4 0 1 2 3 4 10 10 10 10 10 10 10 10 10 10

CD44-FITC Ctip2 DAPI DAPI Vim. Vim. vGlut1 Pax6 Tuj1 Tuj1

0.07 C D Control TRα1-FS397 * 8/14 0.06

Ca * A 0.05

Control 0.04

0.03 40 mV 0.1 s 10/11 0.02 MEDICAL SCIENCES Calcium activity 0.01

0 TRα1-FS397

30pA 1-403X Control-H9

-10pA TR TR 1-FS397 TR 1-FS382 Control-NDC Control-NAS6

Fig. 4. FACS-enriched THRA mutation-containing cortical progenitor cells generate functional neuronal networks. (A) Representative FACS results for WT and THRA mutation-containing cells at day 18. Immunocytochemistry confirmed the cortical identity of FACS- selected progenitor cells. (Scale bars: 50 μm.) (B) Neurons pro- duced by FACS-purified progenitor cells expressed the cortical layer-specific transcription factors TBR1 (layer VI) and CTIP2 (layer V), as well as the vesicular glutamate transporter vGlut1, at day 40. (Scale bars: 50 μm.) (C) Representative traces showing action potential firing in response to stepwise current injection in THRA mutation-containing and control neurons. Numbers indicate the proportion of cells that fired action potentials. (D) Calcium indicator Oregon Green BAPTA was used as a proxy for action potential firing to measure spontaneous neuronal activity. Fluorescence images show control and THRA mutant cultures after loading. Heat maps highlight different levels of activity across the cultures. Total calcium activity across the field of view, ACa, was quantified based on 3 to 6 videos from each cell line (403X and FS382, n = 6 samples from 2 independent inductions; all others, n = 3 samples from 1 induction; *P < 0.05, Student’s t test). Error bars indicate SEM. contribute to the neurologic phenotypes of patients, we undertook 40, whereas the size of THRA mutation-containing clones la- single-cell clonal lineage analysis of cortical progenitors, comparing beled at day 40 was markedly reduced. cell dynamics and clonal output of THRA mutation-containing cells Analyzing clone composition, the average number of progenitor vs. control cells. A small proportion (≤1%) of THRA mutation- cells per clone (as assessed by Ki67 expression) was increased in C THRA containing or control cortical progenitor cells, infected with a control cultures at day 30 and 40 (Fig. 5 ). In mutation- GFP-expressing replication-incompetent lentivirus, were mixed containing cultures, progenitor cell numbers increased only slightly with unlabeled control cultures at day 30 and day 40 after at day 30 and decreased at day 40, suggesting a differing pattern of THRA cortical induction. Their progeny was then traced over time, en- cell division with earlier cell cycle exit of mutation-containing abling the delineation of cell-intrinsic differences between THRA progenitors. Thus, depletion of the THRA mutation-containing progenitor pool leads to premature neuron production and a mutation-containing and control progenitor cells (Fig. 5A). The size and composition of clones (2 or more cells) derived decrease in clonal output over time. By live imaging of cells over a 48-h period, we confirmed that these differences were not due from labeled progenitors were analyzed at 2, 6, and 10 d after to altered cell cycle kinetics or increased cell death (SI Appendix, mixing. The majority of labeled cells expressed either Ki67 or the − + Fig. S4). Furthermore, the majority of KI67 cells were TUJ1 in neuron-specific marker βIII-tubulin, indicating that clones con- both control and THRA mutation-containing cell cultures (Fig. sisted of cycling progenitors and neurons (Fig. 5B). Across the 3 B THRA 5 ); thus, any difference in the proportion of progenitors en- mutation-containing and 3 control cell lines (human iPSC tering quiescence, if present, is likely negligible. lines NDC and NAS6 and human embryonic stem cell line H9), the average size of clones increased progressively over the 10-d Premature Cell Cycle Exit of THRA Mutation-Containing Cortical Progenitor period after labeling (Fig. 5C). Average clone sizes were similar Cells. To further quantitate the division pattern of progenitor cells, in control cortical progenitor cells labeled at either day 30 or day fully differentiated clones were distinguished from persisting clones

Krieger et al. PNAS Latest Articles | 5of10 GFP-Lentivirus AB30+2 days 30+6 days 30+10 days

FACS TRα1

Control Control /

Mix Day 30 Time

Fix Fix Fix TRα1 / Control 30+2 30+6 30+10 GFP Ki67 Tuj1 DAPI

C Control TRα1 Day 30 Day 40 Day 30 Day 40 16 16 Control-NDC TRα1-403X 14 Control-NAS6 14 TRα1-FS397 Control-H9 TRα1-FS382 12 12 10 10 8 8 6 6

4 clone size Average 4 Average clone size Average

2 (number of cells per clone) 2 (number of cells per clone) 0 0 + + 6 6 4 4 2 2

0 cells per clone cells per clone 0 Number of Ki67 Number of Ki67 +2d +6d +10d +2d +6d +10d +2d +6d +10d +2d +6d +10d

Fig. 5. Clonal lineage analysis reveals marked differences between THRA mutation-containing and control progenitor cell dynamics. (A) Schematic of the experimental protocol. GFP-labeled cortical progenitor cells were enriched for by FACS and mixed with WT cultures at day 30 or 40. Cultures were fixed for analysis after 2, 6, or 10 d of incubation. (B) Representative images of THRA mutant and control clones derived from a single progenitor cell after 2, 6, and + 10 d, immunostained for Ki67 (progenitors) and βIII-tubulin (neurons). (Scale bars: 50 μm.) (C) Average clone size and average number of Ki67 cells per clone over time; each data point represents the average of n = 50 to 72 clones. Error bars indicate SEM. containing at least 1 progenitor cell (Fig. 6A). The proportion of to a neurogenic phase, in which they divide asymmetrically to give persisting clones decreased more rapidly in THRA mutation- rise to IPCs with variable but limited neurogenic potential (32). We containing than control cultures (Fig. 6A). At the same time, the found that this model could largely account for differences in the average size of persisting clones increased near-exponentially in con- clonal behavior of THRA mutation-containing vs. control cells, in- trols but more linearly in THRA mutant cells at day 40 (Fig. 6B), cluding the distribution of clone size and composition, under the suggesting that a higher proportion of control progenitor cells are single assumption that THRA mutation-containing RGs differenti- self-renewing symmetrically, whereas THRA mutation-containing ated into IPCs at a higher rate compared with control RGs (Fig. 6 B progenitor cells have transitioned to a more asymmetric, neurogenic and C). Therefore, the clonal lineage data and the computational division pattern. model together demonstrate that THRA mutation-containing pro- To test this, we used a computational model of human cortical genitor cells switch to neurogenesis earlier than control cells, which development in which self-renewing radial glia (RGs) produce in- continue proliferating. termediate progenitor cells (IPCs), which generate neurons through During human neurogenesis, the deepest neuronal layers of asymmetric or terminal divisions (SI Appendix, Supplementary the cerebral cortex emerge first, whereas neurons of the upper Methods and Fig. S5). Based on the findings of a recent in vivo layers appear subsequently and migrate away from the pro- genetic labeling study of cortical neurogenesis in mouse, we pre- genitor cell zone to their final positions. Therefore, expression of sumed that RGs transit from a phase of symmetrical proliferation TBR1, which is high in layer VI neurons in the developing cortex,

6of10 | www.pnas.org/cgi/doi/10.1073/pnas.1908762116 Krieger et al. AB GFP Ki67 Tuj1 DAPI Day 30 Day 40 18 16 Mean number 14 of cells Mean number 12 of Ki67+ cells 10 Ki67+ clone Mixed clone Ki67- clone 8

1.0 1.0 100 100 Control 6 Ki67- 4 0.8 0.8

80 80 Number of cells 2 0.6 0.6 60 60 0 +2d +6d +10d +2d +6d +10d 400.4 Mixed 400.4 Ki67+ Day 30 200.2 200.2 18 0.00 00.0 2dpi 6dpi 10dpi 2dpi 6dpi 10dpi 16 14 1001.0 1001.0 12 800.8 800.8 10 600.6 600.6 8 TRα1 400.4 400.4 6

Day 40 0.2 0.2 4

20 20 Number of cells 2 0.0 0.0 0 0 0 2dpi 6dpi 10dpi 2dpi 6dpi 10dpi Frequency (%) Frequency (%) +2d +6d +10d +2d +6d +10d Control TRα1 C 30

20

10 Control

0 MEDICAL SCIENCES Day 30

1 20 α Frequency (%)

TR 10

0 0102030400 10 20 30 40 0 10 20 30 40 50 30

20

10 Control

0 Day 40

1 20 α Frequency (%)

TR 10

0 0102030400 10 20 30 40 0 10 20 30 40 50 2 dpi 6 dpi 10 dpi Clone size (number of cells per clone)

Fig. 6. Clonal dynamics are consistent with a simple model of reduced progenitor output in THRA mutants. (A) The frequency of clones consisting of only + − Ki67 cells (rose), only Ki67 cells (gray), or both (green) at the different time points (days postmixing, [dpm]). (Scale bar: 50 μm.) (B) Average clone sizes of + persisting clones, with model fit (green line), and average number of Ki67 cells with the model prediction using the best-fitting parameters (red line). (C) Total clone size distributions and model predictions (lines); each histogram represents n = 152 to 176 clones. In A and B, data points represent the average of 3 control and 3 THRA mutant lines; error bars indicate SEM. is thought to precede expression of CTIP2, which is low in layer pression data. The functional units of corticogenesis in the in VI neurons but high in layer V neurons (33). Quantification of vitro system are rosette-like arrangements of progenitor cells, CTIP2 and TBR1 expression in THRA mutation-containing and which self-assemble and recreate aspects of the in vivo niche + control cells at day 40 showed significantly more TBR1 and (22). We observed that following dissociation, THRA mutation- + CTIP2 cells in mutant compared with control cell cultures, containing progenitors self-organized into rosettes less readily providing independent evidence for premature neuronal differ- than controls. To quantify the efficiency of rosette formation, entiation of THRA mutation-containing progenitor cells (SI progenitor cells were plated at a defined density onto micro- Appendix, Fig. S6). patterned chips that only allow cell adhesion in circular areas of different diameters (Fig. 7A). Cell Polarity and Rosette Self-Assembly Defects in THRA Mutation- In control lines, rosettes formed efficiently within 48 h after Containing Cortical Progenitor Cells. Previous studies have shown plating, with areas of 140 μm diameter predominantly hosting a that cortical progenitor cell polarity and attachment at the ven- single rosette (Fig. 7B). The organization of cell types within rosettes tricular surface maintain progenitor identity and proliferation in showed clear apical-basal polarity (Fig. 7C). In contrast, rosette the mouse cortex (34), consistent with foregoing our gene ex- formation was substantially impaired in THRA mutation-containing

Krieger et al. PNAS Latest Articles | 7of10 cell cultures (Fig. 7D). Treatment of control cultures with the control progenitor cells but not in THRA mutation-containing competitive TR antagonist 1-850 (35) decreased rosette forma- cells (SI Appendix,Fig.S7D). Thus, defects in progenitor cell po- tion efficiency and proliferation, indicating that aberrant TH larity and spatial organization may contribute to their premature signaling likely contributes to impaired rosette formation (SI Ap- cell cycle exit. pendix,Fig.S7A). The mechanisms underlying neural rosette formation in vitro Discussion are largely enigmatic (36). Live imaging of control cultures for up Consistent with the known critical role of TH in brain develop- to 72 h after plating found no evidence for directional movement ment, we have documented neurocognitive deficits (e.g., reduced of cells, suggesting that cells do not respond to chemotactic signals nonverbal IQ, motor incoordination, impaired visual motor in- (SI Appendix,Fig.S7B). To test which factors are required for tegration) in patients with RTHα, with some deficits (e.g., motor rosette formation, we developed a computational model in which incoordination, poor finger/hand dexterity) resembling conse- cells move randomly and form attachments at their apical ends quences of untreated congenital . Structural ab- when in close proximity. Both the number of rosettes and the dis- normalities include reduction in cerebellar volume, cortical tribution of rosette centers observed in vitro were statistically microencephaly in adult cases, and decreased white matter tract reproduced in silico, indicating that migration and apical attach- density. The neurocognitive, structural, and functional deficits ment are sufficient for rosette formation (SI Appendix,Fig.S7C). As observed in patients with RTHα suggest abnormalities in several migration defects were not apparent from the live imaging data (SI brain regions, including the axis from the cerebellum to the Appendix,Fig.S7B), impaired apical attachments may be limiting neostriatum and the cerebral cortex. Here we focused on the rosette formation efficiency in THRA mutation-containing cultures. effects of mutant TRα1 on cerebral cortex development and In support of this notion, N-cadherin was highly localized to the function, using cortical neural progenitor cells derived from apical end of progenitor cells in control rosettes but expressed THRA mutation-containing iPSCs from these patients. While our along the cell membrane in THRA mutation-containing cells, experimental model effectively recapitulates defined aspects suggesting impaired adherens junction formation (Fig. 7 D and E), of corticogenesis in vitro, it does not capture other brain de- consistent with the down-regulation of protocadherins observed velopmental processes, such as myelination and cerebellar de- on RNA-seq. Apical expression of the polarity complex protein velopment, that may also be impaired in patients and should be PKCζ and the centrosome-associated γ-tubulin was observed in elucidated in future studies.

ADissociated NPCs Micropattern chip Cortical rosettes C DAPI

I

P in in

DA

ment

67 Vi

Ki

n n i

2

MAP

adher

c

N-

B 24h 48h

DAPI

DAPI

n

n n

i

ubuli

t

Nest

γ-

aPKC aPKC Pax6 Pax6

D Control TRα1-403X TRα1-FS397 TRα1-FS382 E * 100

80

60

acetylated α-tubulin N-cadherin DAPI 40

% rosette formation 20 F Control TRα1-403X TRα1-FS397 TRα1-FS382

0

DAPI 1-403X Control-H9 TR TR 1-FS397 TR 1-FS382 Control-NDC Control-NAS6

N-cadherin

Fig. 7. Cortical rosette self-organized assembly is impaired in THRA mutant cultures. (A) Schematic of the experimental protocol. Dissociated NPCs were plated onto micropatterned chips containing fields of diameter of 140 μm and cultured for several days. (B) Phase-contrast images of individual micro- patterned fields at 24 h and 48 h after plating (control-H9 cells). (Scale bar: 100 μm.) (C) Immunostaining of rosettes fixed at 48 h after plating (control cells). Since RGs in vitro are connected by adherens junctions (AJs) at their apical ends (22), expression of the AJ component N-cadherin (Top Left) was used to visualize rosette centers. Rosettes contained Ki67+ cycling cells, and MAP2+ neurons were frequently observed at the periphery (Top Right). The centrosome protein γ-tubulin and the apically expressed atypical protein kinase C were localized to rosette centers (Bottom), demonstrating the apical-basal polarity of + + Pax6 and Nestin progenitor cells. (Scale bars: 50 μm.) (D) Representative images of THRA mutant and control progenitors plated onto micropatterned chips. N-cadherin (red) was used to visualize rosette centers. (Scale bars: 50 μm.) (E) Bar chart summarizing the proportion of fields of diameter of 140 μm that were occupied by a single rosette (n = 405 for TD, 267 for CK, 187 for KB, 72 for NDC, 279 for NAS6, and 165 for H9, from 2 independent inductions per cell line; *P < 0.001, Student’s t test comparing the 6 THRA mutant vs. 6 control inductions). (F) Representative images of THRA mutant and control progenitors plated onto micropatterned chips. N-cadherin expression (green) is localized to rosette centers in control but not in THRA mutant cultures. (Scale bars: 50 μm.)

8of10 | www.pnas.org/cgi/doi/10.1073/pnas.1908762116 Krieger et al. Our cortical differentiation protocol revealed that THRA mutation- (43–46). A number of genes differentially expressed in THRA containing cortical progenitor cells differentiate prematurely in mutation-containing cells, including the cell surface dystroglycan vitro, with attendant depletion of the progenitor cell pool reducing receptor NRXN1 (neurexin) and the transmembrane protein clonal output, which in vivo would result in microencephaly. NPTN (neuroplastin), have also been implicated in other Quantitative analysis of lineage tracing data confirmed that the neurodevelopmental disorders (26, 27). Since normal cortical fate choices of THRA mutation-containing cortical progenitor progenitor self-renewal requires the orchestration of a range of cells are biased toward neuronal differentiation as opposed to the cell-intrinsic and environmental factors, including transcription continued self-renewal observed in control cells. As single-labeled factors determining cortical stem cell identity and genes regulating THRA mutation-containing progenitor cells were studied in WT cell attachment, proliferation, and movement, as well as cytoskeletal background cultures, such alterations in clonal dynamics likely organization and polarity (34), a range of gene defects may converge reflect distinct cell-autonomous fate behaviors. onto similar abnormal neurodevelopmental phenotypes. Specifically, RNA-seq of early cortical progenitor cells indicated down- our observations suggest that impaired cell adhesion, leading to regulated expression of genes involved in cell polarity or adher- defective cortical size and architecture, may ultimately result in ence in THRA mutation-containing cells. These included several neurocognitive deficits across a range of genetic disorders. protocadherins whose cellular functions are not fully understood In addition to mouse models, novel human iPSC-derived in vitro but may act as adhesion molecules on their own or in association systems are being increasingly used to study neurodevelopmental with other cadherins (37). THRA mutation-containing cortical conditions, hitherto focusing on neuronal maturation and network progenitor cells also showed disordered expression of N-cadherin, function (47–51). Extending these approaches, this study describes a well-described classical cadherin, and defective self-organization a method of assessing the proliferative potential and fate of neural into rosettes on micropatterned chips, which may be explained by stem cells in vitro. Through quantitative analysis of lineage tracing reduced apical cell–cell adhesion based on our computational data, even minor differences in clonal dynamics can be resolved. In model. Taken together, these results indicate that apical adherens addition, we have developed a quantitative assay for the self- junction formation is impaired in THRA mutation-containing organizing capacity of neural stem cells. Together, these ap- cortical progenitor cells. In rodent studies, ventricular attach- proaches provide an experimental platform to investigate the ment through adherens junctions is required for apical-basal po- function of cortical progenitor cells with differing genetic back- larity, spatial organization, and self-renewal of cortical RGs (34, grounds, study neural progenitor cell dynamics during normal 38, 39), and there is emerging evidence indicating that protocadherins and pathological development, and ultimately screen for agents play a role in these processes (40, 41). to ameliorate defective neurogenesis. In the developing cerebral cortex, neural progenitor cell in- MEDICAL SCIENCES teractions through adherens junctions have been shown to consti- Materials and Methods tute a self-supportive niche maintaining progenitor cell proliferation The methodology used in this study is described in more detail in SI Appendix. and inhibiting differentiation (42). Thus, defects in cell polarity and attachment to neighboring stem cells may mediate the premature Neuroimaging. Whole-brain MRI, DTI, and proton MRS were performed at cell cycle exit observed in THRA mutation-containing cortical Great Ormond Street Hospital, London and compared with data obtained progenitor cells. Our findings also suggest that some cell-intrinsic from healthy age-matched controls. All clinical investigations, including consequences of mutant TRα1 dysfunction may become appar- neuroimaging and derivation of iPSCs, were undertaken as part of a protocol ent only through interactions with the local niche. We note, approved by our Research Ethics Committee (Cambridgeshire; LREC 98/154) or however, that our in vitro cortical differentiation protocol gen- were clinically indicated and were performed with previous written informed erates a reduced model of the actual human cortex, lacking, for consent of patients and/or parents. example, vasculature and surrounding supportive tissues, as well Neural Differentiation of iPSCs. Neural induction of human control PSCs and as nonneuronal cells, such as microglia. Neural progenitor cells THRA mutant iPSCs was performed as described previously (21). For FACS in vivo reside in a distinct 3D architecture with axial patterning sorting, 106 cells per sample were collected and stained with fluorochrome- signals and neurophysiological activity not replicated in vitro. conjugated antibodies. Total RNA from cortical cultures was isolated using Thus, our in vitro results on the structure and organization of TRIzol (Sigma-Aldrich), and semiquantitative RT-PCR was performed using neural progenitor cell rosettes might not directly translate to the primers against FOXG1, PAX6,andGAPDH. For Western blot analysis, protein in vivo situation. Future studies may establish the functional was extracted from overnight-frozen cell pellets, and the insoluble fraction hierarchy of determinants of cortical progenitor cell identity in was removed by centrifugation. Band intensity was measured using Image vivo, to further delineate the mechanisms through which mutant Studio software (Li-COR) and normalized to β-actin. Immunofluorescence TRα1 affects neurogenesis. It will also be important in future staining was performed following fixation with 4% paraformaldehyde. studies to validate the modeling presented here that predicts a Electrophysiology and Calcium Imaging. change in intermediate progenitor cell dynamics. For electrophysiological recordings, cortical neurons were incubated with artificial cerebral spinal fluid and Overall, we suggest that abnormal proliferation and adhesion THRA subjected to step depolarizations or stepwise current injections. Recordings of mutation-containing cortical neural progenitors forms were made using a Multiclamp 700 A amplifier (Molecular Devices). For the basis, at least in part, of structural and functional changes calcium imaging, cells were incubated with calcium indicator Oregon Green culminating in neurocognitive deficits seen in patients with RTHα. 488 BAPTA, and calcium activity was recorded on a Deltavision fluorescence We note that our lineage-tracing experiments and RNA-seq data microscope with an EMCCD camera (Applied Precision). cannot resolve whether the observed differences in gene expres- sion and cell fate behavior are due to transcriptional changes Clonal Lineage Analysis. For clonal lineage analysis, third-generation replication- within a comparable population of neural progenitor cells, changes incompetent lentivirus was produced by calcium phosphate transfection of in the cellular composition of the progenitor cell populations, or a HEK293T cells, using pBOP-GFP plasmids combined with packaging plasmids pRSV-Rev, pMDLg/pRRE, and pMD2.G. Cortical progenitor cells were infected at combination of these 2 possibilities. As our understanding of hu- − − man corticogenesis is still evolving, with the cell states and cell high titer before isolation of the CD271 CD44 population by FACS. At day 30 and day 40, sorted progenitors were mixed with unlabeled control cultures. types involved not yet definitively identified, this distinction is at These mixed cultures were fixed and immunostained at 2, 6, and 10 d after plating. present semantic and does not affect the conclusions of our study. Reduced cortical progenitor output and impaired cortical ar- Micropattern Chip Cultures. For micropattern chip cultures, 106 cells were chitecture have been observed in human patients and mouse plated onto laminin-coated CYTOOchips and cultured for 1 to 7 d. For live models of several other neurodevelopmental disorders, including imaging, tissue culture dishes containing cells in N2B27 medium were im-

Down syndrome, schizophrenia, and autism spectrum disorders aged in a BioStation CT (Nikon) at 37 °C with 7% CO2.

Krieger et al. PNAS Latest Articles | 9of10 RNA-Seq. For RNA-seq, total RNA was extracted from 3 control lines and 3 ACKNOWLEDGMENTS. We thank Charles Bradshaw and George Allen for THRA mutant lines at day 12 of in vitro cortical induction. Libraries were help with RNA-seq data analysis. T.G.K. was supported by the Engineering prepared using the Illumina TruSeq Stranded Total RNA sample preparation and Physical Sciences Research Council and the Raymond and Beverly Sackler kit and single-end sequenced on an Illumina HiSeq 1500 system. Gene ex- Foundation. F.J.L., K.C. (210755/Z/18/Z), and B.D.S. are Wellcome Trust pression profiles were clustered using GeneE software, based on Pearson’s Investigators. C.M.M. and K.C. are supported by the National Institute of Health correlation. GO analysis was performed using the PANTHER database Research, Cambridge Biomedical Centre. This research was supported by core (www.panther.org). funding to the Gurdon Institute by the Wellcome Trust and Cancer Research UK.

1. S. C. Bath, C. D. Steer, J. Golding, P. Emmett, M. P. Rayman, Effect of inadequate 26. H.-G. Kim et al., Disruption of neurexin 1 associated with autism spectrum disorder. status in UK pregnant women on cognitive outcomes in their children: Results Am. J. Hum. Genet. 82, 199–207 (2008). from the Avon Longitudinal Study of Parents and Children (ALSPAC). Lancet 382, 331– 27. S. Desrivières et al.; IMAGEN Consortium, Single nucleotide polymorphism in the 337 (2013). neuroplastin locus associates with cortical thickness and intellectual ability in ado- 2. V. Mohan et al., Maternal thyroid hormone deficiency affects the fetal neocortico- lescents. Mol. Psychiatry 20, 263–274 (2015). genesis by reducing the proliferating pool, rate of neurogenesis and indirect neuro- 28. S. H. Yuan et al., Cell-surface marker signatures for the isolation of neural stem cells, Exp. Neurol. – genesis. 237, 477 488 (2012). glia and neurons derived from human pluripotent stem cells. PLoS One 6, e17540 3. A. Pathak, R. A. Sinha, V. Mohan, K. Mitra, M. M. Godbole, Maternal thyroid hormone (2011). before the onset of fetal thyroid function regulates reelin and downstream signaling 29. G. Lee et al., Isolation and directed differentiation of neural crest stem cells derived Cereb. Cortex – cascade affecting neocortical neuronal migration. 21,11 21 (2011). from human embryonic stem cells. Nat. Biotechnol. 25, 1468–1475 (2007). et al 4. R. Lavado-Autric ., Early maternal hypothyroxinemia alters histogenesis and 30. S. J. Morrison, P. M. White, C. Zock, D. J. Anderson, Prospective identification, iso- cerebral cortex cytoarchitecture of the progeny. J. Clin. Invest. 111, 1073–1082 (2003). lation by flow cytometry, and in vivo self-renewal of multipotent mammalian neural 5. E. Ausó et al., A moderate and transient deficiency of maternal thyroid function at crest stem cells. Cell 96, 737–749 (1999). the beginning of fetal neocorticogenesis alters neuronal migration. Endocrinology 31. Y.-C. Yu, R. S. Bultje, X. Wang, S.-H. Shi, Specific synapses develop preferentially 145, 4037–4047 (2004). among sister excitatory neurons in the neocortex. Nature 458, 501–504 (2009). 6. N. K. Moog et al., Influence of maternal during gestation on fetal 32. P. Gao et al., Deterministic progenitor behavior and unitary production of neurons in brain development. Neuroscience 342,68–100 (2017). the neocortex. Cell 159, 775–788 (2014). 7. S. Horn, H. Heuer, Thyroid hormone action during brain development: More ques- 33. B. J. Molyneaux, P. Arlotta, J. R. L. Menezes, J. D. Macklis, Neuronal subtype specifi- tions than answers. Mol. Cell. Endocrinol. 315,19–26 (2010). cation in the cerebral cortex. Nat. Rev. Neurosci. 8, 427–437 (2007). 8. J. Bernal, Thyroid hormone receptors in brain development and function. Nat. Clin. 34. M. Götz, W. B. Huttner, The cell biology of neurogenesis. Nat. Rev. Mol. Cell Biol. 6, Pract. Endocrinol. Metab. 3, 249–259 (2007). – 9. J. M. Vaquerizas, S. K. Kummerfeld, S. A. Teichmann, N. M. Luscombe, A census of 777 788 (2005). et al human transcription factors: Function, expression and evolution. Nat. Rev. Genet. 10, 35. M. Schapira ., Discovery of diverse thyroid hormone receptor antagonists by high- Proc. Natl. Acad. Sci. U.S.A. – 252–263 (2009). throughput docking. 100, 7354 7359 (2003). 10. D. J. Bradley, H. C. Towle, W. S. Young, 3rd, Spatial and temporal expression of alpha- 36. M. Karus, S. Blaess, O. Brüstle, Self-organization of neural tissue architectures from J. Comp. Neurol. – and beta-thyroid hormone receptor mRNAs, including the beta 2-subtype, in the pluripotent stem cells. 522, 2831 2844 (2014). developing mammalian nervous system. J. Neurosci. 12, 2288–2302 (1992). 37. J. A. Weiner, J. D. Jontes, Protocadherins, not prototypical: A complex tale of their 11. I. Astapova et al., The nuclear corepressor, NCoR, regulates thyroid hormone action in interactions, expression, and functions. Front. Mol. Neurosci. 6, 4 (2013). vivo. Proc. Natl. Acad. Sci. U.S.A. 105, 19544–19549 (2008). 38. Y. Kosodo et al., Asymmetric distribution of the apical plasma membrane during 12. M. Ito, C. X. Yuan, H. J. Okano, R. B. Darnell, R. G. Roeder, Involvement of the neurogenic divisions of mammalian neuroepithelial cells. EMBO J. 23, 2314–2324 TRAP220 component of the TRAP/SMCC coactivator complex in embryonic develop- (2004). ment and thyroid hormone action. Mol. Cell 5, 683–693 (2000). 39. M. Kadowaki et al., N-cadherin mediates cortical organization in the mouse brain. 13. E. Bochukova et al., A mutation in the thyroid hormone receptor alpha gene. N. Engl. Dev. Biol. 304,22–33 (2007). J. Med. 366, 243–249 (2012). 40. C. Compagnucci et al., Characterizing PCDH19 in human induced pluripotent stem 14. A. van Mullem et al., Clinical phenotype and mutant TRα1. N. Engl. J. Med. 366, 1451– cells (iPSCs) and iPSC-derived developing neurons: Emerging role of a protein involved 1453 (2012). in controlling polarity during neurogenesis. Oncotarget 6, 26804–26813 (2015). 15. C. Moran et al., Resistance to thyroid hormone caused by a mutation in thyroid 41. P. Zhang et al., Protocadherin 11 x regulates differentiation and proliferation of hormone receptor (TR)α1 and TRα2: Clinical, biochemical, and genetic analyses of neural stem cell in vitro and in vivo. J. Mol. Neurosci. 54, 199–210 (2014). three related patients. Lancet Diabetes Endocrinol. 2, 619–626 (2014). 42. J. Zhang et al., Cortical neural precursors inhibit their own differentiation via et al 16. C. Moran ., An adult female with resistance to thyroid hormone mediated by N-cadherin maintenance of β-catenin signaling. Dev. Cell 18, 472–479 (2010). α J. Clin. Endocrinol. Metab. – defective thyroid hormone receptor . 98, 4254 4261 43. L. Chakrabarti, Z. Galdzicki, T. F. Haydar, Defects in embryonic neurogenesis and (2013). initial synapse formation in the forebrain of the Ts65Dn mouse model of Down et al 17. S. Espiard ., A novel mutation in THRA gene associated with an atypical phe- syndrome. J. Neurosci. 27, 11483–11495 (2007). J. Clin. Endocrinol. Metab. – notype of resistance to thyroid hormone. 100, 2841 2848 44. D. W. Meechan, E. S. Tucker, T. M. Maynard, A.-S. LaMantia, Diminished dosage of (2015). 22q11 genes disrupts neurogenesis and cortical development in a mouse model of 18. A. Tylki-Szymanska et al., Thyroid hormone resistance syndrome due to mutations in 22q11 deletion/DiGeorge syndrome. Proc. Natl. Acad. Sci. U.S.A. 106, 16434–16445 the thyroid hormone receptor α gene (THRA). J. Med. Genet. 52, 312–316 (2015). (2009). 19. T. Fauquier et al., Severe impairment of cerebellum development in mice expressing a 45. D. Gallagher et al., Ankrd11 is a chromatin regulator involved in autism that is es- dominant-negative mutation inactivating thyroid hormone receptor alpha1 isoform. sential for neural development. Dev. Cell 32,31–42 (2015). Dev. Biol. 356, 350–358 (2011). 46. B. E. LaMonica, J. H. Lui, X. Wang, A. R. Kriegstein, OSVZ progenitors in the human 20. C. Moran, K. Chatterjee, Resistance to thyroid hormone α-emerging definition of a cortex: An updated perspective on neurodevelopmental disease. Curr. Opin. Neurobiol. disorder of thyroid hormone action. J. Clin. Endocrinol. Metab. 101, 2636–2639 (2016). – 21. Y. Shi, P. Kirwan, F. J. Livesey, Directed differentiation of human pluripotent stem cells 22, 747 753 (2012). et al to cerebral cortex neurons and neural networks. Nat. Protoc. 7, 1836–1846 (2012). 47. K. J. Brennand ., Modelling schizophrenia using human induced pluripotent stem Nature – 22. Y. Shi, P. Kirwan, J. Smith, H. P. C. Robinson, F. J. Livesey, Human cerebral cortex cells. 473, 221 225 (2011). et al development from pluripotent stem cells to functional excitatory synapses. Nat. 48. S. P. Pas¸ca ., Using iPSC-derived neurons to uncover cellular phenotypes associ- Nat. Med. – Neurosci. 15, 477–486 (2012). ated with Timothy syndrome. 17, 1657 1662 (2011). et al 23. I.-H. Park et al., Disease-specific induced pluripotent stem cells. Cell 134, 877–886 49. S. J. Chamberlain ., Induced pluripotent stem cell models of the genomic im- (2008). printing disorders Angelman and Prader-Willi syndromes. Proc. Natl. Acad. Sci. U.S.A. 24. M. B. Johnson et al., Single-cell analysis reveals transcriptional heterogeneity of 107, 17668–17673 (2010). neural progenitors in human cortex. Nat. Neurosci. 18, 637–646 (2015). 50. M. C. N. Marchetto et al., A model for neural development and treatment of Rett 25. S. J. Chou, C. G. Perez-Garcia, T. T. Kroll, D. D. M. O’Leary, Lhx2 specifies regional fate syndrome using human induced pluripotent stem cells. Cell 143, 527–539 (2010). in Emx1 lineage of telencephalic progenitors generating cerebral cortex. Nat. Neurosci. 51. A. R. Muotri et al., L1 retrotransposition in neurons is modulated by MeCP2. Nature 12,1381–1389 (2009). 468, 443–446 (2010).

10 of 10 | www.pnas.org/cgi/doi/10.1073/pnas.1908762116 Krieger et al.