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:

Longitudinal modeling of human neuronal aging reveals the contribution of the RCAN1–TFEB pathway to Huntington’s disease neurodegeneration

Abstract

Aging is a common risk factor in neurodegenerative disorders. Investigating neuronal aging in an isogenic background stands to facilitate analysis of the interplay between neuronal aging and neurodegeneration. Here we perform direct neuronal reprogramming of longitudinally collected human fibroblasts to reveal genetic pathways altered at different ages. Comparative transcriptome analysis of longitudinally aged striatal medium spiny neurons (MSNs) in Huntington’s disease identified pathways involving RCAN1, a negative regulator of calcineurin. Notably, RCAN1 protein increased with age in reprogrammed MSNs as well as in human postmortem striatum and RCAN1 knockdown rescued patient-derived MSNs of Huntington’s disease from degeneration. RCAN1 knockdown enhanced chromatin accessibility of genes involved in longevity and autophagy, mediated through enhanced calcineurin activity, leading to TFEB’s nuclear localization by dephosphorylation. Furthermore, G2-115, an analog of glibenclamide with autophagy-enhancing activities, reduced the RCAN1–calcineurin interaction, phenocopying the effect of RCAN1 knockdown. Our results demonstrate that targeting RCAN1 genetically or pharmacologically can increase neuronal resilience in Huntington’s disease.

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: Identification of RCAN1 as an age-associated factor in reprogrammed MSNs from longitudinally collected fibroblasts.
Fig. 2: RCAN1 KD protects HD-MSNs from degeneration and induces chromatin accessibility changes.
Fig. 3: RCAN1 KD- and CaN KD-induced chromatin changes.
Fig. 4: Enhancing TFEB function by RCAN1 KD via its nuclear localization.
Fig. 5: RCAN1 KD promotes neuronal resilience through enhancing TFEB nuclear localization.
Fig. 6: G2-115 promotes TFEB function by reducing RCAN1–CaN interaction and promoting TFEB nuclear localization.

Similar content being viewed by others

Data availability

RNA-seq data and ATAC-seq data presented in the current study will be available through the Gene Expression Omnibus (GEO) at NCBI with accession IDs GSE241430 and GSE210996. Raw data that support all findings of the study are available as Source data and supplementary tables provided with this paper. All other data that support the findings of this study are available from the corresponding author upon reasonable request.

References

  1. Wyss-Coray, T. Ageing, neurodegeneration and brain rejuvenation. Nature 539, 180–186 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  2. Rose, M. R. Adaptation, aging, and genomic information. Aging (Albany NY) 1, 444–450 (2009).

    Article  PubMed  Google Scholar 

  3. Machiela, E. & Southwell, A. L. Biological aging and the cellular pathogenesis of Huntington’s disease. J. Huntingtons Dis. 9, 115–128 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  4. Niccoli, T. & Partridge, L. Ageing as a risk factor for disease. Curr. Biol. 22, R741–752 (2012).

    Article  CAS  PubMed  Google Scholar 

  5. Hou, Y. et al. Ageing as a risk factor for neurodegenerative disease. Nat. Rev. Neurol. 15, 565–581 (2019).

    Article  PubMed  Google Scholar 

  6. Huh, C. J. et al. Maintenance of age in human neurons generated by microRNA-based neuronal conversion of fibroblasts. eLife https://doi.org/10.7554/eLife.18648 (2016).

  7. Abernathy, D. G. et al. MicroRNAs induce a permissive chromatin environment that enables neuronal subtype-specific reprogramming of adult human fibroblasts. Cell Stem Cell 21, 332–348.e9 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  8. Cates, K. et al. Deconstructing stepwise fate conversion of human fibroblasts to neurons by microRNAs. Cell Stem Cell 28, 127–140.e9 (2021).

    Article  CAS  PubMed  Google Scholar 

  9. Victor, M. B. et al. Striatal neurons directly converted from Huntington’s disease patient fibroblasts recapitulate age-associated disease phenotypes. Nat. Neurosci. 21, 341–352 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  10. Oh, Y. M. et al. Age-related Huntington’s disease progression modeled in directly reprogrammed patient-derived striatal neurons highlights impaired autophagy. Nat. Neurosci. https://doi.org/10.1038/s41593-022-01185-4 (2022).

  11. Oh, Y. M., Lee, S. W. & Yoo, A. S. Modeling Huntington disease through microRNA-mediated neural reprogramming identifies age-associated autophagy dysfunction driving the onset of neurodegeneration. Autophagy https://doi.org/10.1080/15548627.2023.2175572 (2023).

  12. Klee, C. B., Crouch, T. H. & Krinks, M. H. Calcineurin: a calcium- and calmodulin-binding protein of the nervous system. Proc. Natl Acad. Sci. USA 76, 6270–6273 (1979).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  13. Rusnak, F. & Mertz, P. Calcineurin: form and function. Physiol. Rev. 80, 1483–1521 (2000).

    Article  CAS  PubMed  Google Scholar 

  14. Hoeffer, C. A. et al. The Down syndrome critical region protein RCAN1 regulates long-term potentiation and memory via inhibition of phosphatase signaling. J. Neurosci. 27, 13161–13172 (2007).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  15. Hogan, P. G., Chen, L., Nardone, J. & Rao, A. Transcriptional regulation by calcium, calcineurin, and NFAT. Genes Dev. 17, 2205–2232 (2003).

    Article  CAS  PubMed  Google Scholar 

  16. Li, Y. et al. The structure of the RCAN1:CN complex explains the inhibition of and substrate recruitment by calcineurin. Sci. Adv. https://doi.org/10.1126/sciadv.aba3681 (2020).

  17. Mitchell, A. N. et al. Brain expression of the calcineurin inhibitor RCAN1 (Adapt78). Arch. Biochem. Biophys. 467, 185–192 (2007).

    Article  CAS  PubMed  Google Scholar 

  18. Porta, S., Marti, E., de la Luna, S. & Arbones, M. L. Differential expression of members of the RCAN family of calcineurin regulators suggests selective functions for these proteins in the brain. Eur. J. Neurosci. 26, 1213–1226 (2007).

    Article  PubMed  Google Scholar 

  19. Cook, C. N., Hejna, M. J., Magnuson, D. J. & Lee, J. M. Expression of calcipressin1, an inhibitor of the phosphatase calcineurin, is altered with aging and Alzheimer’s disease. J. Alzheimers Dis. 8, 63–73 (2005).

    Article  CAS  PubMed  Google Scholar 

  20. Fuentes, J. J. et al. DSCR1, overexpressed in Down syndrome, is an inhibitor of calcineurin-mediated signaling pathways. Hum. Mol. Genet. 9, 1681–1690 (2000).

    Article  CAS  PubMed  Google Scholar 

  21. Fuentes, J. J. et al. A new human gene from the Down syndrome critical region encodes a proline-rich protein highly expressed in fetal brain and heart. Hum. Mol. Genet. 4, 1935–1944 (1995).

    Article  CAS  PubMed  Google Scholar 

  22. Rothermel, B. et al. A protein encoded within the Down syndrome critical region is enriched in striated muscles and inhibits calcineurin signaling. J. Biol. Chem. 275, 8719–8725 (2000).

    Article  CAS  PubMed  Google Scholar 

  23. Ermak, G., Morgan, T. E. & Davies, K. J. Chronic overexpression of the calcineurin inhibitory gene DSCR1 (Adapt78) is associated with Alzheimer’s disease. J. Biol. Chem. 276, 38787–38794 (2001).

    Article  CAS  PubMed  Google Scholar 

  24. Harris, C. D., Ermak, G. & Davies, K. J. RCAN1-1L is overexpressed in neurons of Alzheimer’s disease patients. FEBS J. 274, 1715–1724 (2007).

    Article  CAS  PubMed  Google Scholar 

  25. Sardiello, M. et al. A gene network regulating lysosomal biogenesis and function. Science 325, 473–477 (2009).

    Article  CAS  PubMed  Google Scholar 

  26. Settembre, C. et al. TFEB links autophagy to lysosomal biogenesis. Science 332, 1429–1433 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  27. Settembre, C. et al. TFEB controls cellular lipid metabolism through a starvation-induced autoregulatory loop. Nat. Cell Biol. 15, 647–658 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  28. Wang, Y. et al. An analog of glibenclamide selectively enhances autophagic degradation of misfolded α1-antitrypsin Z. PLoS ONE 14, e0209748 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  29. Victor, M. B. et al. Generation of human striatal neurons by microRNA-dependent direct conversion of fibroblasts. Neuron 84, 311–323 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  30. Yoo, A. S. et al. MicroRNA-mediated conversion of human fibroblasts to neurons. Nature 476, 228–231 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  31. Han, K. A. et al. Histone deacetylase 3 promotes RCAN1 stability and nuclear translocation. PLoS ONE 9, e105416 (2014).

    Article  PubMed  PubMed Central  Google Scholar 

  32. Noh, E. H. et al. Covalent NEDD8 conjugation increases RCAN1 protein stability and potentiates its inhibitory action on calcineurin. PLoS ONE 7, e48315 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  33. Dudilot, A., Trillaud-Doppia, E. & Boehm, J. RCAN1 regulates bidirectional synaptic plasticity. Curr. Biol. 30, 1167–1176 e1162 (2020).

    Article  CAS  PubMed  Google Scholar 

  34. Genetic Modifiers of Huntington’s Disease, C. Identification of genetic factors that modify clinical onset of Huntington’s disease. Cell 162, 516–526 (2015).

    Article  Google Scholar 

  35. Hickey, M. A. & Chesselet, M. F. Apoptosis in Huntington’s disease. Prog. Neuropsychopharmacol. Biol. Psychiatry 27, 255–265 (2003).

    Article  CAS  PubMed  Google Scholar 

  36. Khan, S. et al. Implication of caspase-3 as a common therapeutic target for multineurodegenerative disorders and its inhibition using nonpeptidyl natural compounds. BioMed Res. Int. 2015, 379817 (2015).

    Article  PubMed  PubMed Central  Google Scholar 

  37. Ona, V. O. et al. Inhibition of caspase-1 slows disease progression in a mouse model of Huntington’s disease. Nature 399, 263–267 (1999).

    Article  CAS  PubMed  Google Scholar 

  38. Portera-Cailliau, C., Hedreen, J. C., Price, D. L. & Koliatsos, V. E. Evidence for apoptotic cell death in Huntington disease and excitotoxic animal models. J. Neurosci. 15, 3775–3787 (1995).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  39. Soles-Tarres, I. et al. Pituitary adenylate cyclase-activating polypeptide (PACAP) protects striatal cells and improves motor function in Huntington’s disease models: role of PAC1 receptor. Front. Pharmacol. 12, 797541 (2021).

    Article  CAS  PubMed  Google Scholar 

  40. Ganz, J. et al. A novel specific PERK activator reduces toxicity and extends survival in Huntington’s disease models. Sci. Rep. 10, 6875 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  41. Leitman, J. et al. ER stress-induced eIF2-alpha phosphorylation underlies sensitivity of striatal neurons to pathogenic huntingtin. PLoS ONE 9, e90803 (2014).

    Article  PubMed  PubMed Central  Google Scholar 

  42. Corces, M. R. et al. An improved ATAC-seq protocol reduces background and enables interrogation of frozen tissues. Nat. Methods 14, 959–962 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  43. Fox, L. M. et al. Huntington’s disease pathogenesis is modified in vivo by Alfy/Wdfy3 and selective macroautophagy. Neuron 105, 813–821.e6 (2020).

    Article  CAS  PubMed  Google Scholar 

  44. Ashkenazi, A. et al. Polyglutamine tracts regulate beclin 1-dependent autophagy. Nature 545, 108–111 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  45. Azzi, J. R., Sayegh, M. H. & Mallat, S. G. Calcineurin inhibitors: 40 years later, can’t live without. J. Immunol. 191, 5785–5791 (2013).

    Article  CAS  PubMed  Google Scholar 

  46. Medina, D. L. et al. Lysosomal calcium signalling regulates autophagy through calcineurin and TFEB. Nat. Cell Biol. 17, 288–299 (2015).

    Article  PubMed  PubMed Central  Google Scholar 

  47. Brun, M., Glubrecht, D. D., Baksh, S. & Godbout, R. Calcineurin regulates nuclear factor I dephosphorylation and activity in malignant glioma cell lines. J. Biol. Chem. 288, 24104–24115 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  48. Jain, J. et al. The T-cell transcription factor NFATp is a substrate for calcineurin and interacts with Fos and Jun. Nature 365, 352–355 (1993).

    Article  CAS  PubMed  Google Scholar 

  49. Shalizi, A. et al. A calcium-regulated MEF2 sumoylation switch controls postsynaptic differentiation. Science 311, 1012–1017 (2006).

    Article  CAS  PubMed  Google Scholar 

  50. Li, H., Rao, A. & Hogan, P. G. Interaction of calcineurin with substrates and targeting proteins. Trends Cell Biol. 21, 91–103 (2011).

    Article  CAS  PubMed  Google Scholar 

  51. Huang, C. C. et al. Calcineurin-mediated dephosphorylation of c-Jun Ser-243 is required for c-Jun protein stability and cell transformation. Oncogene 27, 2422–2429 (2008).

    Article  CAS  PubMed  Google Scholar 

  52. Fornes, O. et al. JASPAR 2020: update of the open-access database of transcription factor binding profiles. Nucleic Acids Res. 48, D87–D92 (2020).

    CAS  PubMed  Google Scholar 

  53. Li, L. et al. RB1CC1-enhanced autophagy facilitates PSCs activation and pancreatic fibrogenesis in chronic pancreatitis. Cell Death Dis. 9, 952 (2018).

    Article  PubMed  PubMed Central  Google Scholar 

  54. Yao, J. et al. Deletion of autophagy inducer RB1CC1 results in degeneration of the retinal pigment epithelium. Autophagy 11, 939–953 (2015).

    Article  PubMed  PubMed Central  Google Scholar 

  55. Zhen, X., Uryu, K., Cai, G., Johnson, G. P. & Friedman, E. Age-associated impairment in brain MAPK signal pathways and the effect of caloric restriction in Fischer 344 rats. J. Gerontol. A Biol. Sci. Med. Sci. 54, B539–548 (1999).

    Article  CAS  PubMed  Google Scholar 

  56. Liu, Y. et al. Age-related decline in mitogen-activated protein kinase activity in epidermal growth factor-stimulated rat hepatocytes. J. Biol. Chem. 271, 3604–3607 (1996).

    Article  CAS  PubMed  Google Scholar 

  57. Pyo, J. O. et al. Overexpression of Atg5 in mice activates autophagy and extends lifespan. Nat. Commun. 4, 2300 (2013).

    Article  PubMed  Google Scholar 

  58. Melendez, A. et al. Autophagy genes are essential for dauer development and life-span extension in C. elegans. Science 301, 1387–1391 (2003).

    Article  CAS  PubMed  Google Scholar 

  59. Jia, K. & Levine, B. Autophagy is required for dietary restriction-mediated life span extension in C. elegans. Autophagy 3, 597–599 (2007).

    Article  PubMed  Google Scholar 

  60. Lapierre, L. R. et al. The TFEB orthologue HLH-30 regulates autophagy and modulates longevity in Caenorhabditis elegans. Nat. Commun. 4, 2267 (2013).

    Article  PubMed  Google Scholar 

  61. Hansen, M. et al. A role for autophagy in the extension of lifespan by dietary restriction in C. elegans. PLoS Genet. 4, e24 (2008).

    Article  PubMed  PubMed Central  Google Scholar 

  62. Nakamura, S. et al. Mondo complexes regulate TFEB via TOR inhibition to promote longevity in response to gonadal signals. Nat. Commun. 7, 10944 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  63. Settembre, C. et al. A lysosome-to-nucleus signalling mechanism senses and regulates the lysosome via mTOR and TFEB. EMBO J. 31, 1095–1108 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  64. Roczniak-Ferguson, A. et al. The transcription factor TFEB links mTORC1 signaling to transcriptional control of lysosome homeostasis. Sci. Signal. 5, ra42 (2012).

    Article  PubMed  PubMed Central  Google Scholar 

  65. Martina, J. A., Chen, Y., Gucek, M. & Puertollano, R. MTORC1 functions as a transcriptional regulator of autophagy by preventing nuclear transport of TFEB. Autophagy 8, 903–914 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  66. Yoshii, S. R. & Mizushima, N. Monitoring and measuring autophagy. Int. J. Mol. Sci. https://doi.org/10.3390/ijms18091865 (2017).

  67. Leeman, D. S. et al. Lysosome activation clears aggregates and enhances quiescent neural stem cell activation during aging. Science 359, 1277–1283 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  68. Pankiv, S. et al. p62/SQSTM1 binds directly to Atg8/LC3 to facilitate degradation of ubiquitinated protein aggregates by autophagy. J. Biol. Chem. 282, 24131–24145 (2007).

    Article  CAS  PubMed  Google Scholar 

  69. Hidvegi, T. et al. An autophagy-enhancing drug promotes degradation of mutant α1-antitrypsin Z and reduces hepatic fibrosis. Science 329, 229–232 (2010).

    Article  CAS  PubMed  Google Scholar 

  70. Thellung, S., Corsaro, A., Nizzari, M., Barbieri, F. & Florio, T. Autophagy activator drugs: a new opportunity in neuroprotection from misfolded protein toxicity. Int. J. Mol. Sci. https://doi.org/10.3390/ijms20040901 (2019).

  71. Hudry, E. et al. Inhibition of the NFAT pathway alleviates amyloid beta neurotoxicity in a mouse model of Alzheimer’s disease. J. Neurosci. 32, 3176–3192 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  72. Bechstein, W. O. Neurotoxicity of calcineurin inhibitors: impact and clinical management. Transpl. Int. 13, 313–326 (2000).

    Article  CAS  PubMed  Google Scholar 

  73. Lee, S. et al. The calcineurin inhibitor Sarah (Nebula) exacerbates Aβ42 phenotypes in a Drosophila model of Alzheimer’s disease. Dis. Model. Mech. 9, 295–306 (2016).

    CAS  PubMed  PubMed Central  Google Scholar 

  74. Richner, M., Victor, M. B., Liu, Y., Abernathy, D. & Yoo, A. S. MicroRNA-based conversion of human fibroblasts into striatal medium spiny neurons. Nat. Protoc. 10, 1543–1555 (2015).

    Article  CAS  PubMed  Google Scholar 

  75. Lu, Y. L. & Yoo, A. S. Mechanistic insights into microRNA-induced neuronal reprogramming of human adult fibroblasts. Front. Neurosci. 12, 522 (2018).

    Article  PubMed  PubMed Central  Google Scholar 

  76. McCoy, M. J. et al. LONGO: an R package for interactive gene length dependent analysis for neuronal identity. Bioinformatics 34, i422–i428 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  77. Church, V. A. et al. Generation of human neurons by microRNA-mediated direct conversion of dermal fibroblasts. Methods Mol. Biol. 2239, 77–100 (2021).

    Article  CAS  PubMed  Google Scholar 

Download references

Acknowledgements

We thank the Genome Technology Access Center at Washington University for deep-sequencing experiments. An image of the experimental scheme was created with BioRender.com. This study was supported by a Hereditary Disease Foundation (HDF) Grant, the Cure Alzheimer’s Fund (CAF), the CHDI Foundation Research Agreement, National Institute on Aging (NIA) Grant no. RF1AG056296, NIA grant no. R01AG078964, National Institute of Neurological Disorders and Stroke (NINDS) Grant no. R01NS107488, the Farrell Foundation Fund and the Mallinckrodt Scholar Award (A.S.Y.).

Author information

Authors and Affiliations

Authors

Contributions

S.W.L. and A.S.Y. conceived and developed the idea, designed the experiments and analyzed data. S.W.L. and Y.M.O. performed all experiments and analyses unless specified. M.B.V., S.C. and I.S. performed reduction-of-function testing and analysis of candidate HD modifiers. Y.Y. performed whole-cell recordings. S.D. provided human striatal section samples. R.E.D., S.C.P., G.A.S. and D.H.P. developed the G2 analog. S.W.L. and A.S.Y. wrote the manuscript. A.S.Y. supervised the overall project.

Corresponding author

Correspondence to Andrew S. Yoo.

Ethics declarations

Competing interests

The authors declare no competing interests.

Peer review

Peer review information

Nature Aging thanks Scott Zeitlin and the other, anonymous, reviewer(s) for their contribution to the peer review of this work.

Additional information

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

Extended data

Extended Data Fig. 1 Gene expression profiling in longitudinally collected fibroblasts and corresponding reprogrammed MSNs.

a-c, Whole-cell recording from reprogrammed MSNs from three independent longitudinal groups (individual I (a); II (b); III (c)) co-cultured with human astrocytes showing the inward/outward currents and multiple action potentials (APs). d, RT-qPCR analysis of DARPP-32 expression in longitudinally aged MSNs (n = 6, ****p < 0.0001, The sample size (n) corresponds to the number of biologically independent samples). Statistical significance was determined using two-tailed unpaired t-test and mean±s.e.m. e and f, Heatmap of Differentially Expressed Genes (DEGs) in fibroblasts (e) and MSNs (f) (FDR < 0.05, │FC│ ≥ 1.5). g, Venn diagram of the genes enriched in calcium signaling pathway from old HD-MSNs.

Source data

Extended Data Fig. 2 Age-associated RCAN1 in longitudinally aged MSN.

a, Upstream regulator analysis of up- or down-regulated genes in old fibroblasts and MSNs. b, Gene network of upstream regulators and DEGs. c, Representative immunoblotting (top) and quantification (bottom) of RCAN1 expression in six MSNs from 22, 29, 24 (young) and 53, 50, 60 (old) years old-individuals (young n = 3, old n = 3, **p = 0.0038). d, Quantification of RCAN1 mRNA from six independent fibroblasts and MSNs from three longitudinal individuals (I, II, and III) (Fibroblasts n = 12, MSNs n = 12). e, Representative immunoblotting (top, left) and quantification of relative RCAN1 expression normalized to GAPDH (bottom. left) in Young / Old-MSNs from three longitudinal individuals treated with cyclohexamide (CHX). Comparison of RCAN1 expression in CHX-treated MSNs (Young and Old) from three longitudinal individuals in the presence of DMSO, MG132 or, CQ (right) (n = 6). f. Representative immunoblotting (top) and quantification (bottom) of RCAN1 expression in age-matched control-Old-MSNs and HD-MSNs (n = 6). g. Representative Immunoblotting (top) and quantification (bottom) of HDAC3 expression in MSNs from three longitudinal individuals (n = 6, *p = 0.0467). h. Representative immunoblotting (top) and quantification (bottom) of RCAN1 expression in individual III’s MSNs (n = 2). Statistical significance was determined using two-tailed unpaired t-test (c,f,g) and one-way ANOVA with Tukey’s post hoc test (d,h). *p < 0.05, **p < 0.01, ns, not significant, and mean±s.e.m (c,d,f,g). The sample size (n) corresponds to the number of biologically independent samples (c-h).

Source data

Extended Data Fig. 3 Identification of modifier genes whose reduction protects HD-MSNs from degeneration.

a, Experimental scheme of genetic modifiers testing in HD-MSNs. b, Representative images (left) and quantification (right) of MAP2-, NCAM-, NEUN-, ACTL6B-, DARPP-32-, and GABA-positive cells from four independent HD-MSNs (HD.43, HD.40, HD.47, HD.45). An average of 300 cells per each were counted from three or more randomly chosen fields (n = 4). Scale bars represent 20 μm. c, High-content imaging of Sytox green dye accumulation in HD-MSNs (HD.46) in a 96-well format. Representative images of HD-MSNs in each well of a 96-well plate, immunostained with anti-GABA, TUBB3, and DARPP-32 antibodies (left). Example pictures for high content image analysis to measure cell death levels (right): Hoechst for whole cell population and Sytox-green for dead cells. d, Quantification of Sytox-positive cells from HD-MSNs (HD.46) and healthy control (Ctrl.17) at post-induction day 35 (n = 2). e, Quantification of Sytox-positive cells in HD-MSNs (HD.46) transduced with shRNAs of modifier genes. The genes whose reduction significantly lowered cell death levels were marked (red) within the pink area (± 10 % of cell death level from healthy control) compared to control shRNA. Statistical significance was determined using unpaired t-test and mean±s.e.m (n = 2, RCAN1: p = 0.0143); RTCA: p = 0.0198); UBE2D4: p = 0.0073). f, Representative image (left) and quantification (right) of Sytox-positive cells from three independent HD-MSNs (HD.46, HD.44, HD.43) transduced with shRNAs of each gene (n = 12). Scale bars represent 100 μm. Box-and-whiskers plot: The center line denotes the median value while the box contains the 25th to 75th percentiles of dataset. The whiskers mark minimal value to maximal value. ****p < 0.0001. g, Representative image (left) and quantification (right) of cells with HTT inclusion bodies (IBs) in HD-MSNs (HD.40) transduced with shRNAs of each gene. Cells were immunostained with anti-HTT and TUBB3 antibodies. An average of 100 cells per each were counted from four to six randomly chosen fields (n = 6, 4, 6, 4). Scale bars represent 10μm. Statistical significance was determined using unpaired t-test (e) and one-way ANOVA with Tukey’s post-hoc test (f,g); ****p < 0.0001, *p < 0.05, ns, not significant, and mean±s.e.m. The sample size (n) corresponds to the number of biologically independent samples (b,d,e,f,g).

Source data

Extended Data Fig. 4 Validation of reprogrammed neurons of rescuing or non-rescuing condition for ATAC-sequencing.

a, RCAN1 expression in fibroblasts transduced with shRCAN1 (top) or RCAN1 (middle) in a dose-dependent manner. RCAN1 expression in HD-MSNs (HD.43) transduced with shCtrl, shRCAN1, or RCAN1 (bottom). b, Representative image (top) and quantification (bottom) of DARPP-32-positive cells from four independent HD-MSNs transduced with shCtrl, shRCAN1, or shCaN (HD.43, HD.40, HD.47, HD.45). Cells were immunostained with anti-DARPP-32 and TUBB3 antibodies. An average of 183 cells of each were counted from three or more randomly chosen fields (n = 4). Scale bars represent 10 μM. c, RT-qPCR analysis of the expression of RCAN1 and CaN in (b) (n = 12, 12, 8, 8). Statistical significance was determined using one-way ANOVA with Tukey’s post-hoc test (b) and two-tailed unpaired t-test (c); ****p < 0.0001, ns, not significant, and mean±s.e.m (b,c). The sample size (n) corresponds to the number of biologically independent samples (b,c).

Source data

Extended Data Fig. 5 RCAN1 promotes nuclear localization of TFEB for HD survival.

a, Expression of phosphor-TFEB in fibroblasts transduced with Control, TFEB wildtype, or phosphor-mutant (S142/211 A). b. Representative image (left) and quantification of nuclear TFEB from three independent HD-MSNs (HD.45, HD.45b, HD.47) transduced with shCtrl or shRCAN1. Cells were treated with DMSO or Cyclosporin A (CaN inhibitor) (n = 3). shCtrl versus shRCAN1 ***p = 0.0002, shRCAN1 versus shRCAN1+Cyclosporin A ***p = 0.0001. c, Representative image (left) and quantification of Sytox-positive cells (middle) from three independent HD-MSNs (HD.45, HD.45b, HD.47) transduced with shCtrl, shRCAN1, or shTFEB. Expression of RCAN1 and TFEB in HD-MSNs transduced with shCtrl, shRCAN1, or shTFEB (right) (n = 3). Statistical significance was determined using one-way ANOVA with Tukey’s post hoc test (b,c); ***p < 0.001, *p < 0.05, ns, not significant, and mean±s.e.m (b,c). The sample size (n) corresponds to the number of biologically independent samples (b,c).

Source data

Extended Data Fig. 6 Neuroprotective role of G2-115 through reducing RCAN1-CaN interaction.

a, Immunoprecipitation analysis of RCAN1-transduced fibroblasts with anti-CaN antibody followed by immunoblotting with anti-RCAN1 antibody. Cells are treated with 0.5 μM of G2-115 and 60 μM of chloroquine (lysosome inhibitor). b, Immunoprecipitation analysis of RCAN1-transduced fibroblasts with anti-CaN followed by immunoblotting with anti-RCAN1 antibody. Cells were treated with DMSO or 0.5 μM of G2-115, 8 mM of metformin, and 500 nM of rapamycin. c, Experimental scheme of NanoBit binding assay (top). Binding assay of HEK293 cells transfected with RCAN1 fused to large Bit and CaN fused to small Bit. Cells were treated with 0.5, 1.0, 1.5, and 2.0 μM of G2-115 in a dose-dependent manner (bottom). (n = 3, The sample size (n) corresponds to the number of independent experiments). DMSO versus G2-115 1.0 μM 0.5 hr *p = 0.0379, 1.0 hr **p = 0.0091, 2.0 hr **p = 0.0082, DMSO versus G2-115 2.0 μM 0.5 hr **p = 0.0044, 1.0 hr *p = 0.0230, 2.0 hr **p = 0.0084. d, Quantification of CYTO-ID-positive cells from three independent HD-MSNs (HD.45, HD.45b, HD.47) treated with DMSO or G2-115. Cells were transduced with RCAN1 (n = 3). DMSO versus G2-115 **p = 0.0025, G2-115 versus G2-115 + RCAN1 cDNA **p = 0.0060. e, Representative image (left) and quantification (right) of nuclear TFEB from three-independent HD-MSNs (HD.45, HD.45b, HD.47) treated with DMSO or G2-115. Cells were immunostained with anti-TFEB and TUBB3 antibodies. An average of 107 cells per each were counted from three or more randomly chosen fields (n = 3, ****p < 0.0001). Scale bars represent 20 μm. f, Graphical work model to illustrate the function of RCAN1-CaN-TFEB cascade in Young/Old-MSNs (left) and the neuroprotective role of RCAN1 KD for HD survival. Statistical significance was determined using one-way ANOVA with Tukey’s post hoc test in (d) and two-tailed unpaired t-test (c,e); ****p < 0.0001, **p < 0.01, *p < 0.05, ns, not significant, and mean±s.e.m (c-e). The sample size (n) corresponds to the number of biologically independent samples (d,e).

Source data

Supplementary information

Reporting Summary

Supplementary Table 1

Information on fibroblast and brain samples used in the study.

Supplementary Table 2

List of GeM-HD modifiers used in the study.

Source data

Source Data Fig. 1

Unprocessed western blots and/or gels.

Source Data Fig. 2

Statistical source data.

Source Data Fig. 3

Statistical source data.

Source Data Fig. 4

Unprocessed western blots and/or gels.

Source Data Fig. 5

Unprocessed western blots and/or gels.

Source Data Fig. 6

Unprocessed western blots and/or gels.

Source Data Extended Data Fig. 1

Statistical source data.

Source Data Extended Data Fig. 2

Unprocessed western blots and/or gels.

Source Data Extended Data Fig. 3

Statistical source data.

Source Data Extended Data Fig. 4

Unprocessed western blots and/or gels.

Source Data Extended Data Fig. 5

Unprocessed western blots and/or gels.

Source Data Extended Data Fig. 6

Unprocessed western blots and/or gels.

Rights and permissions

Springer Nature or its licensor (e.g. a society or other partner) holds exclusive rights to this article under a publishing agreement with the author(s) or other rightsholder(s); author self-archiving of the accepted manuscript version of this article is solely governed by the terms of such publishing agreement and applicable law.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Lee, S.W., Oh, Y.M., Victor, M.B. et al. Longitudinal modeling of human neuronal aging reveals the contribution of the RCAN1–TFEB pathway to Huntington’s disease neurodegeneration. Nat Aging 4, 95–109 (2024). https://doi.org/10.1038/s43587-023-00538-3

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s43587-023-00538-3

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