Skip to main content

Thank you for visiting nature.com. You are using a browser version with limited support for CSS. To obtain the best experience, we recommend you use a more up to date browser (or turn off compatibility mode in Internet Explorer). In the meantime, to ensure continued support, we are displaying the site without styles and JavaScript.

  • Article
  • Published:

Long-term maturation of human cortical organoids matches key early postnatal transitions

Abstract

Human stem-cell-derived models provide the promise of accelerating our understanding of brain disorders, but not knowing whether they possess the ability to mature beyond mid- to late-fetal stages potentially limits their utility. We leveraged a directed differentiation protocol to comprehensively assess maturation in vitro. Based on genome-wide analysis of the epigenetic clock and transcriptomics, as well as RNA editing, we observe that three-dimensional human cortical organoids reach postnatal stages between 250 and 300 days, a timeline paralleling in vivo development. We demonstrate the presence of several known developmental milestones, including switches in the histone deacetylase complex and NMDA receptor subunits, which we confirm at the protein and physiological levels. These results suggest that important components of an intrinsic in vivo developmental program persist in vitro. We further map neurodevelopmental and neurodegenerative disease risk genes onto in vitro gene expression trajectories to provide a resource and webtool (Gene Expression in Cortical Organoids, GECO) to guide disease modeling.

This is a preview of subscription content, access via your institution

Access options

Buy this article

Prices may be subject to local taxes which are calculated during checkout

Fig. 1: Methylation and transcriptional maturation in long-term hCS.
Fig. 2: Biological processes and cell-type marker changes in long-term hCS.
Fig. 3: RNA editing in hCS.
Fig. 4: Developmental isoform switches in hCS.
Fig. 5: Mapping neurodevelopmental and neuropsychiatric disorder genes onto hCS differentiation.
Fig. 6: Mapping neurodegenerative disorder genes onto hCS differentiation.

Similar content being viewed by others

Data availability

Gene expression data and methylation data are available in the Gene Expression Omnibus (GEO) under accession numbers GSE150122 and GSE150123. The accompanying GECO webtool can be accessed at https://labs.dgsom.ucla.edu/geschwind/files/view/html/GECO.html. The BrainSpan data are available in the database of Genotypes and Phenotypes (dbGaP) under Study Accession phs000755.v2.p1. Single-cell data from human fetal cerebral cortex can be found at http://geschwindlab.dgsom.ucla.edu/pages/codexviewer and at dbGaP under Study accession phs001836. eCLIP data for FXR1 and FMR1are available in GEO with accession number GSE107895. Human cortical organoid single-cell sequencing data are available in GEO with accession number GSE107771. Source data are provided with this paper.

Code availability

The code used in this manuscript can be found at https://github.com/dhglab/human_cortical_organoid_maturation.

References

  1. Pasca, S. P. The rise of three-dimensional human brain cultures. Nature 553, 437–445 (2018).

    Article  CAS  PubMed  Google Scholar 

  2. Mertens, J., Marchetto, M. C., Bardy, C. & Gage, F. H. Evaluating cell reprogramming, differentiation and conversion technologies in neuroscience. Nat. Rev. Neurosci. 17, 424–437 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  3. Lancaster, M. A. et al. Cerebral organoids model human brain development and microcephaly. Nature 501, 373–379 (2013).

    Article  CAS  PubMed  Google Scholar 

  4. Qian, X. et al. Brain-region-specific organoids using mini-bioreactors for modeling ZIKV exposure. Cell 165, 1238–1254 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  5. Pasca, A. M. et al. Functional cortical neurons and astrocytes from human pluripotent stem cells in 3D culture. Nat. Methods 12, 671–678 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  6. Sloan, S. A. et al. Human astrocyte maturation captured in 3D cerebral cortical spheroids derived from pluripotent stem cells. Neuron 95, 779–790 e776 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  7. Camp, J. G. et al. Human cerebral organoids recapitulate gene expression programs of fetal neocortex development. Proc. Natl Acad. Sci. USA 112, 15672–15677 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  8. Amiri, A., et al. Transcriptome and epigenome landscape of human cortical development modeled in organoids. Science 362, eaat6720 (2018).

  9. Quadrato, G. et al. Cell diversity and network dynamics in photosensitive human brain organoids. Nature 545, 48–53 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  10. Trevino, A. E., et al. Chromatin accessibility dynamics in a model of human forebrain development. Science 367, eaay1645 (2020).

  11. Yoon, S. J. et al. Reliability of human cortical organoid generation. Nat. Methods 16, 75–78 (2019).

    Article  CAS  PubMed  Google Scholar 

  12. Horvath, S. DNA methylation age of human tissues and cell types. Genome Biol. 14, R115 (2013).

    Article  PubMed  PubMed Central  Google Scholar 

  13. McEwen, L. M. et al. Systematic evaluation of DNA methylation age estimation with common preprocessing methods and the Infinium MethylationEPIC BeadChip array. Clin. Epigenetics 10, 123 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  14. Horvath, S. et al. Epigenetic clock for skin and blood cells applied to Hutchinson Gilford progeria syndrome and ex vivo studies. Aging (Albany NY) 10, 1758–1775 (2018).

    CAS  Google Scholar 

  15. Stein, J. L. et al. A quantitative framework to evaluate modeling of cortical development by neural stem cells. Neuron 83, 69–86 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  16. Li, M., et al. Integrative functional genomic analysis of human brain development and neuropsychiatric risks. Science 362, eaat7615 (2018).

  17. Kang, H. J. et al. Spatio-temporal transcriptome of the human brain. Nature 478, 483–489 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  18. Khan, T. A., et al. Neuronal defects in a human cellular model of 22q11.2 deletion syndrome. Nat. Med. 26, 1888–1898 (2020).

  19. Pollen, A. A. et al. Establishing cerebral organoids as models of human-specific brain evolution. Cell 176, 743–756.e717 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  20. Bhaduri, A., et al. Cell stress in cortical organoids impairs molecular subtype specification. Nature 578,142–148 (2020).

  21. Polioudakis, D. et al. A single-cell transcriptomic atlas of human neocortical development during mid-gestation. Neuron 103, 785–801.e788 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  22. Birey, F. et al. Assembly of functionally integrated human forebrain spheroids. Nature 545, 54–59 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  23. Marton, R. M., et al. Differentiation and maturation of oligodendrocytes in human three-dimensional neural cultures. Nat. Neurosci. 22, 484–491 (2019).

  24. Hwang, T. et al. Dynamic regulation of RNA editing in human brain development and disease. Nat. Neurosci. 19, 1093–1099 (2016).

    Article  CAS  PubMed  Google Scholar 

  25. Sanjana, N. E., Levanon, E. Y., Hueske, E. A., Ambrose, J. M. & Li, J. B. Activity-dependent A-to-I RNA editing in rat cortical neurons. Genetics 192, 281–287 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  26. Tran, S. S. et al. Widespread RNA editing dysregulation in brains from autistic individuals. Nat. Neurosci. 22, 25–36 (2019).

    Article  CAS  PubMed  Google Scholar 

  27. Morris, M. J., Karra, A. S. & Monteggia, L. M. Histone deacetylases govern cellular mechanisms underlying behavioral and synaptic plasticity in the developing and adult brain. Behav. Pharmacol. 21, 409–419 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  28. Sheng, M., Cummings, J., Roldan, L. A., Jan, Y. N. & Jan, L. Y. Changing subunit composition of heteromeric NMDA receptors during development of rat cortex. Nature 368, 144–147 (1994).

    Article  CAS  PubMed  Google Scholar 

  29. Watanabe, M., Inoue, Y., Sakimura, K. & Mishina, M. Developmental changes in distribution of NMDA receptor channel subunit mRNAs. Neuroreport 3, 1138–1140 (1992).

    Article  CAS  PubMed  Google Scholar 

  30. Liu, H., Hu, Q., Kaufman, A., D’Ercole, A. J. & Ye, P. Developmental expression of histone deacetylase 11 in the murine brain. J. Neurosci. Res. 86, 537–543 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  31. Le Magueresse, C. & Monyer, H. GABAergic interneurons shape the functional maturation of the cortex. Neuron 77, 388–405 (2013).

    Article  PubMed  Google Scholar 

  32. Wyllie, D. J., Livesey, M. R. & Hardingham, G. E. Influence of GluN2 subunit identity on NMDA receptor function. Neuropharmacology 74, 4–17 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  33. Galluzzi, L., Yamazaki, T. & Kroemer, G. Linking cellular stress responses to systemic homeostasis. Nat. Rev. Mol. Cell Biol. 19, 731–745 (2018).

    Article  CAS  PubMed  Google Scholar 

  34. Nowakowski, T. J. et al. Spatiotemporal gene expression trajectories reveal developmental hierarchies of the human cortex. Science 358, 1318–1323 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  35. Sunwoldt, J., Bosche, B., Meisel, A. & Mergenthaler, P. Neuronal culture microenvironments determine preferences in bioenergetic pathway use. Front. Mol. Neurosci. 10, 305 (2017).

    Article  PubMed  PubMed Central  Google Scholar 

  36. Ma, E. H. et al. Metabolic profiling using stable isotope tracing reveals distinct patterns of glucose utilization by physiologically activated CD8(+) T cells. Immunity 51, 856–870 e855 (2019).

    Article  CAS  PubMed  Google Scholar 

  37. Gaspard, N. et al. An intrinsic mechanism of corticogenesis from embryonic stem cells. Nature 455, 351–357 (2008).

    Article  CAS  PubMed  Google Scholar 

  38. Shen, Q. et al. The timing of cortical neurogenesis is encoded within lineages of individual progenitor cells. Nat. Neurosci. 9, 743–751 (2006).

    Article  CAS  PubMed  Google Scholar 

  39. Zhang, W., Qu, J., Liu, G. H. & Belmonte, J. C. I. The ageing epigenome and its rejuvenation. Nat. Rev. Mol. Cell Biol. 21, 137–150 (2020).

    Article  CAS  PubMed  Google Scholar 

  40. Raj, B. & Blencowe, B. J. Alternative splicing in the mammalian nervous system: recent insights into mechanisms and functional roles. Neuron 87, 14–27 (2015).

    Article  CAS  PubMed  Google Scholar 

  41. Sohal, V. S. & Rubenstein, J. L. R. Excitation–inhibition balance as a framework for investigating mechanisms in neuropsychiatric disorders. Mol. Psychiatry 24, 1248–1257 (2019).

    Article  PubMed  PubMed Central  Google Scholar 

  42. Mosser, C. A., Baptista, S., Arnoux, I. & Audinat, E. Microglia in CNS development: shaping the brain for the future. Prog. Neurobiol. 149–150, 1–20 (2017).

    Article  PubMed  Google Scholar 

  43. Lin, Y. T. et al. APOE4 causes widespread molecular and cellular alterations associated with Alzheimer’s disease phenotypes in human iPSC-derived brain cell types. Neuron 98, 1294 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  44. Cakir, B. et al. Engineering of human brain organoids with a functional vascular-like system. Nat. Methods 16, 1169–1175 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  45. Balik, A., Penn, A. C., Nemoda, Z. & Greger, I. H. Activity-regulated RNA editing in select neuronal subfields in hippocampus. Nucleic Acids Res. 41, 1124–1134 (2013).

    Article  CAS  PubMed  Google Scholar 

  46. Paşca, S. P. et al. Using iPSC-derived neurons to uncover cellular phenotypes associated with Timothy syndrome. Nat. Med. 17, 1657–1662 (2011).

    Article  PubMed  PubMed Central  Google Scholar 

  47. Dobin, A. et al. STAR: ultrafast universal RNA-seq aligner. Bioinformatics 29, 15–21 (2013).

    Article  CAS  PubMed  Google Scholar 

  48. Li, B. & Dewey, C. N. RSEM: accurate transcript quantification from RNA-Seq data with or without a reference genome. BMC Bioinf. 12, 323 (2011).

    Article  CAS  Google Scholar 

  49. Oldham, M. C., Langfelder, P. & Horvath, S. Network methods for describing sample relationships in genomic datasets: application to Huntington’s disease. BMC Syst. Biol. 6, 63 (2012).

    Article  PubMed  PubMed Central  Google Scholar 

  50. McKenna, A. et al. The genome analysis Toolkit: a MapReduce framework for analyzing next-generation DNA sequencing data. Genome Res. 20, 1297–1303 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  51. Purcell, S. et al. PLINK: a tool set for whole-genome association and population-based linkage analyses. Am. J. Hum. Genet. 81, 559–575 (2007).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  52. Leek, J. T., Johnson, W. E., Parker, H. S., Jaffe, A. E. & Storey, J. D. The sva package for removing batch effects and other unwanted variation in high-throughput experiments. Bioinformatics 28, 882–883 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  53. Cleveland, W. S. & Devlin, S. J. Locally weighted regression: an approach to regression analysis by local fitting. J. Am. Stat. Assoc. 83, 596–610 (1988).

    Article  Google Scholar 

  54. Wickham, H. ggplot2: Elegant Graphics for Data Analysis (Springer, 2016).

  55. Liao, Y., Smyth, G. K. & Shi, W. featureCounts: an efficient general purpose program for assigning sequence reads to genomic features. Bioinformatics 30, 923–930 (2014).

    Article  CAS  PubMed  Google Scholar 

  56. Plaisier, S. B., Taschereau, R., Wong, J. A. & Graeber, T. G. Rank-rank hypergeometric overlap: identification of statistically significant overlap between gene-expression signatures. Nucleic Acids Res. 38, e169 (2010).

    Article  PubMed  PubMed Central  Google Scholar 

  57. Robinson, M. D., McCarthy, D. J. & Smyth, G. K. edgeR: a Bioconductor package for differential expression analysis of digital gene expression data. Bioinformatics 26, 139–140 (2010).

    Article  CAS  PubMed  Google Scholar 

  58. Ritchie, M. E. et al. limma powers differential expression analyses for RNA-sequencing and microarray studies. Nucleic Acids Res. 43, e47 (2015).

    Article  PubMed  PubMed Central  Google Scholar 

  59. Triche, T. J. Jr., Weisenberger, D. J., Van Den Berg, D., Laird, P. W. & Siegmund, K. D. Low-level processing of Illumina Infinium DNA Methylation BeadArrays. Nucleic Acids Res. 41, e90 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  60. Korotkevich, G., Sukhov, V. & Sergushichev, A. Fast gene set enrichment analysis. Preprint at bioRxiv https://doi.org/10.1101/060012 (2019).

  61. Langfelder, P., Luo, R., Oldham, M. C. & Horvath, S. Is my network module preserved and reproducible? PLoS Comput. Biol. 7, e1001057 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  62. Langfelder, P. & Horvath, S. WGCNA: an R package for weighted correlation network analysis. BMC Bioinformatics 9, 559 (2008).

    Article  PubMed  PubMed Central  Google Scholar 

  63. Sloan, S. A., Andersen, J., Pasca, A. M., Birey, F. & Pasca, S. P. Generation and assembly of human brain region-specific three-dimensional cultures. Nat. Protoc. 13, 2062–2085 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  64. Cribari-Neto, F. & Zeileis, A. Beta regression in R. J. Stat. Softw. 34, 1–24 (2010).

  65. Kim, D., Langmead, B. & Salzberg, S. L. HISAT: a fast spliced aligner with low memory requirements. Nat. Methods 12, 357–360 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  66. Bahn, J. H. et al. Accurate identification of A-to-I RNA editing in human by transcriptome sequencing. Genome Res. 22, 142–150 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  67. Lee, J. H., Ang, J. K. & Xiao, X. Analysis and design of RNA sequencing experiments for identifying RNA editing and other single-nucleotide variants. RNA 19, 725–732 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  68. Porath, H. T., Carmi, S. & Levanon, E. Y. A genome-wide map of hyper-edited RNA reveals numerous new sites. Nat. Commun. 5, 4726 (2014).

    Article  CAS  PubMed  Google Scholar 

  69. Parikshak, N. N., et al. Genome-wide changes in lncRNA, splicing, and regional gene expression patterns in autism. Nature 540, 423–427 (2016).

  70. Wu, Y. E., Parikshak, N. N., Belgard, T. G. & Geschwind, D. H. Genome-wide, integrative analysis implicates microRNA dysregulation in autism spectrum disorder. Nat. Neurosci. 19, 1463–1476 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  71. Nishikura, K. A-to-I editing of coding and non-coding RNAs by ADARs. Nat. Rev. Mol. Cell Biol. 17, 83–96 (2016).

    Article  CAS  PubMed  Google Scholar 

  72. Parikshak, N. N. et al. Integrative functional genomic analyses implicate specific molecular pathways and circuits in autism. Cell 155, 1008–1021 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  73. Wang, D., et al. Comprehensive functional genomic resource and integrative model for the human brain. Science 362, eaat8464 (2018).

  74. Jansen, I. E. et al. Genome-wide meta-analysis identifies new loci and functional pathways influencing Alzheimer’s disease risk. Nat. Genet. 51, 404–413 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  75. Cacace, R., Sleegers, K. & Van Broeckhoven, C. Molecular genetics of early-onset Alzheimer’s disease revisited. Alzheimers Dement. 12, 733–748 (2016).

    Article  PubMed  Google Scholar 

  76. Chang, D. et al. A meta-analysis of genome-wide association studies identifies 17 new Parkinson’s disease risk loci. Nat. Genet. 49, 1511–1516 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  77. Farrer, M. J. Genetics of Parkinson disease: paradigm shifts and future prospects. Nat. Rev. Genet. 7, 306–318 (2006).

    Article  CAS  PubMed  Google Scholar 

  78. Chen, J. A. et al. Joint genome-wide association study of progressive supranuclear palsy identifies novel susceptibility loci and genetic correlation to neurodegenerative diseases. Mol. Neurodegener. 13, 41 (2018).

    Article  PubMed  PubMed Central  Google Scholar 

  79. Greaves, C. V. & Rohrer, J. D. An update on genetic frontotemporal dementia. J. Neurol. 266, 2075–2086 (2019).

    Article  PubMed  PubMed Central  Google Scholar 

  80. Yu, G., Wang, L. G., Han, Y. & He, Q. Y. clusterProfiler: an R package for comparing biological themes among gene clusters. OMICS 16, 284–287 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  81. Skene, N. G. & Grant, S. G. Identification of vulnerable cell types in major brain disorders using single cell transcriptomes and expression weighted cell type enrichment. Front. Neurosci. 10, 16 (2016).

    Article  PubMed  PubMed Central  Google Scholar 

Download references

Acknowledgements

We acknowledge experimental support from members of the Pasca Lab at Stanford University: A. M. Pasca, N. Huber, T. Khan, F. Birey, A. Puno, L. Li and T. Li., and members of the Pasca lab and Geschwind lab for helpful discussions and support, including A. Elkins for setting up the GECO web browser. This work was supported by a Distinguished Investigator Award from the Paul G. Allen Frontiers Group (to D.H.G.); by grants from California Institute of Regenerative Medicine (CIRM) and the National Institute of Mental Health Convergent Neuroscience Consortium (U01 MH115745) (to D.H.G. and S.P.P.), the Stanford Human Brain Organogenesis Program in the Wu Tsai Neuroscience Institute (to S.P.P.), Stanford Bio-X (to S.P.P.), the Stanford Wu Tsai Neuroscience Institute Big Idea Grant (to S.P.P.), the Kwan Funds (to S.P.P.), the Senkut Research Fund (to S.P.P.), the Autism Science Foundation (ASF) and the Brain and Behavior Research Foundation Young Investigator award (Brain & Behavior Research Foundation) (to A.G.). S.P.P. is a New York Stem Cell Foundation (NYSCF) Robertson Stem Cell Investigator and a Chan Zuckerberg Initiative (CZI) Ben Barres Investigator.

Author information

Authors and Affiliations

Authors

Contributions

A.G., S.P.P. and D.H.G. planned and directed experiments, guided analyses, and wrote the manuscript with assistance from all authors. A.G. performed RNA-seq analysis and methylation analysis. S.-J.Y. performed cell culture, DNA and RNA extraction. S.S.T. performed RNA editing analysis. C.D.M. performed electrophysiology recordings. J.A. performed immunohistochemistry. J.Y.P. and A.M.V. performed western blots. S.H. analyzed the methylation data and interpreted the findings. X.X. supervised RNA editing analysis and interpretation. J.R.H. supervised electrophysiology experiments and interpretation.

Corresponding authors

Correspondence to Sergiu P. Pașca or Daniel H. Geschwind.

Ethics declarations

Competing interests

S.P.P. is listed on a patent held by Stanford University that covers the generation of region-specific brain organoids (US patent 62/477,858). All other authors declare no competing interests.

Additional information

Peer review information Nature Neuroscience thanks Kristen Brennand and the other, anonymous, reviewer(s) for their contribution to the peer review of this work.

Publisher’s note Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Extended data

Extended Data Fig. 1 Data description and quality.

a, Timepoints and hiPSC line information for the 62 samples used for RNA sequencing (left). Samples were differentiated from 5 cell lines derived from 4 individuals. Timepoints and hiPSC information for the 50 samples used for DNA methylation (right). Samples were differentiated from 6 cell lines derived from 5 individuals (see Supplementary Tables 1 and 2). Two samples (blue) were hybridized in replicate for quality control purposes and their values were averaged. Each point represents one sample from a specific cell line (y-axis) and differentiation day (x-axis). Full circles represent sample coming from males and rings represent samples coming from females. Gray and white background shading show aggregation of differentiation days into stages. b, Principal component analysis (PCA) of the expression data. The values represent the adjusted r squared of the PC with the covariates indicated. The numbers in brackets on axis titles are the percent of variance explained by the PC. The first 5 PCs, which explain 57.1% of the total variance, show high association with differentiation day. c, Dendrogram of hierarchical clustering of samples demonstrating that differentiation day but no other covariates (individual, Sex, batch) is driving the clustering of samples. d, Violin plots of the variance explained by each of the covariates for each gene. Outlines represent the density of the percent of variance explained. The numbers are the median value of percent of explained variance for each variable. Boxplots in d show: center – median, lower hinge – 25% quantile, upper hinge – 75% quantile, lower whisker – smallest observation greater than or equal to lower hinge –1.5× interquartile range, upper whisker – largest observation less than or equal to upper hinge +1.5× interquartile range. n = 62 samples from 5 hiPSC lines derived from 4 individuals.

Extended Data Fig. 2 Cell stress in hCS.

a, Trajectories of metabolic cell stress genes20 hCS (top) and in vivo (bottom). b, In vitro (left) and in vivo (right) module eigen genes of glycolysis (organoid.Sloan.human.ME.paleturquoise) and ER stress (organoid.human.ME.darkred) previously suggested to be upregulated in vitro20. Gray areas denote time of shift from prenatal to postnatal gene expression. In (a) and (b) shaded gray area around the trajectory represents the 95% confidence interval, vertical gray lines represent birth and vertical gray bars denote the shift from prenatal to postnatal gene expression based on matching to in vivo patterns. For in vitro data n = 62 samples from 5 hiPSC lines derived from 4 individuals and for in vivo data n = 196 from 24 individuals. c, Scatterplot visualization of cells in in developing fetal cortex colored by major cell types22. vRG, ventral radial glia; oRG, outer radial glia; CGE, caudal ganglionic eminence; MGE, medial ganglionic eminence; OPC, oligodendrocyte precursor cell; IP, intermediate progenitors.

Extended Data Fig. 3 Changes in biological processes between early and later stages of differentiation.

a, Number of differentially expressed genes when comparing differentiation day 200 to differentiation day 25 (left) and differentiation day 400 to differentiation day 200 (right). Magenta bar represents upregulated genes and the green bar represents down-regulated genes. b, Top 3 up- and downregulated GO terms enriched in genes ranked by logFC using gene set enrichment analysis, (GSEA; FDR < 0.05). c, Normalized expression of marker genes in vivo for neurons, intermediate progenitors, astrocytes, and radial glia as well as superficial and deep layer cortical neurons. d, Scaled expression of fetal and mature astroglial genes7 during differentiation. A shift between fetal and mature gene sets occurs at ~250 days of hCS differentiation. e, Normalized expression of marker genes for inhibitory neurons and oligodendrocyte precursor cells (OPCs) that are not preserved in hCS. f, Normalized expression of activity-dependent genes that are not preserved in hCS. In (c), (e) and (f) shaded gray area around the trajectory represents the 95% confidence interval, vertical gray lines represent birth and vertical gray bars denote the shift from prenatal to postnatal gene expression based on matching to in vivo patterns. For in vitro data n = 62 samples from 5 hiPSC lines derived from 4 individuals and for in vivo data n = 196 from 24 individuals.

Extended Data Fig. 4 Overlap between hCS and in vivo WGCNA modules.

Overlap of genes in hCS and the BrainSpan in vivo modules. Significant ORs are presented. Modules were clustered using complete-linkage hierarchal clustering. Color represents the OR of each overlap. In vivo neuronal modules (green) and glial modules (purple) are marked.

Extended Data Fig. 5 Overlap between hCS and in vivo editing modules.

a, Overlap of editing sites in hCS and BrainSpan in vivo modules. Significant ORs are presented. b, Distributions showing the closest distances between editing sites from BrainSpan editing modules and FMRP or FXR1P eCLIP peaks (blue). The median of 10,000 sets of control sites (black) is depicted for comparison. See Methods for details of P-value calculation. N, number of editing sites shown. c, Overlap of editing sites within 1000bp of a CLIP site in hCS and BrainSpan in vivo modules. Significant ORs are presented. *** FDR < 0.005.

Extended Data Fig. 6 Expression of select genes in the in vivo fetal cerebral cortex.

a, Immunohistochemistry of HDAC2 and the deep layer marker CTIP2 (BCL11B) at post conception week 21 (PCW21). CP, cortical plate. Scale bars, 100 μm. The Immunohistochemistry experiment was performed once. b, Scatterplot visualization of cells in developing fetal human cerebral cortex colored by major cell types 22. vRG, ventral radial glia; oRG, outer radial glia; CGE, caudal ganglionic eminence; MGE, medial ganglionic eminence; OPC, oligodendrocyte precursor cell, IP, Intermediate progenitors.

Extended Data Fig. 7 Mapping neurodegenerative and epilepsy disorder genes onto hCS differentiation.

Mapping of genes associated with progressive supranuclear palsy (PSP) and frontotemporal dementia (FTD) (a), and epilepsy (b) onto hCS differentiation trajectories. The first column shows clustering of scaled normalized expression of genes associated with a disorder. Genes (in rows) are clustered using hierarchical clustering on the Euclidean distance between genes. Samples (columns) are ordered by differentiation day (represented by gray bars) with the earliest days on the left and latest timepoints on the right. The 5 most representative genes (highest correlation with the cluster eigengene) are shown. The second column shows the cluster eigengenes (first principal component) for the identified gene clusters. Shaded gray area around the trajectory line represents the 95% confidence interval. The third column shows the top GO terms enriched in the identified clusters. The fourth column shows cell types over expressed in either all the genes associated with a disorder (above line) or in the genes from the identified clusters. Number and color represent the fold change. Significance was tested using a one-sided permutation test with 100,000 permutations. P values were corrected for multiple testing using the Benjamini-Hochberg method. * FDR < 0.05, ** FDR < 0.01, *** FDR < 0.005. n = 62 samples from 5 hiPSC lines derived from 4 individuals. IP, intermediate progenitors; GlutN, glutamatergic neurons; IN, interneurons; OPC, oligodendrocyte progenitor cells.

Supplementary information

Reporting Summary

Supplementary Tables

Supplementary Tables 1–5.

Source data

Source Data Fig. 4

Unprocessed western blots.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Gordon, A., Yoon, SJ., Tran, S.S. et al. Long-term maturation of human cortical organoids matches key early postnatal transitions. Nat Neurosci 24, 331–342 (2021). https://doi.org/10.1038/s41593-021-00802-y

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41593-021-00802-y

This article is cited by

Search

Quick links

Nature Briefing

Sign up for the Nature Briefing newsletter — what matters in science, free to your inbox daily.

Get the most important science stories of the day, free in your inbox. Sign up for Nature Briefing