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:

Age-related Huntington’s disease progression modeled in directly reprogrammed patient-derived striatal neurons highlights impaired autophagy

Abstract

Huntington’s disease (HD) is an inherited neurodegenerative disorder with adult-onset clinical symptoms, but the mechanism by which aging drives the onset of neurodegeneration in patients with HD remains unclear. In this study we examined striatal medium spiny neurons (MSNs) directly reprogrammed from fibroblasts of patients with HD to model the age-dependent onset of pathology. We found that pronounced neuronal death occurred selectively in reprogrammed MSNs from symptomatic patients with HD (HD-MSNs) compared to MSNs derived from younger, pre-symptomatic patients (pre-HD-MSNs) and control MSNs from age-matched healthy individuals. We observed age-associated alterations in chromatin accessibility between HD-MSNs and pre-HD-MSNs and identified miR-29b-3p, whose age-associated upregulation promotes HD-MSN degeneration by impairing autophagic function through human-specific targeting of the STAT3 3′ untranslated region. Reducing miR-29b-3p or chemically promoting autophagy increased the resilience of HD-MSNs against neurodegeneration. Our results demonstrate miRNA upregulation with aging in HD as a detrimental process driving MSN degeneration and potential approaches for enhancing autophagy and resilience of HD-MSNs.

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: Differential manifestation of neurodegeneration between MSNs derived from healthy control individuals and patients with HD.
Fig. 2: Identification of gene module associated with autophagy dysfunction in HD-MSNs by WGCNA.
Fig. 3: Autophagy inhibitor induces neurodegeneration of pre-HD-MSNs.
Fig. 4: Autophagy activator rescues HD-MSNs from degeneration.
Fig. 5: Comparative analysis of chromatin accessibility between pre-HD-MSNs and HD-MSNs.
Fig. 6: Inhibition of miR-29b-3p enhances autophagy and rescues HD-MSNs from degeneration.
Fig. 7: Inhibition of miR-29b-3p–STAT3 axis enhances autophagy and rescues HD-MSNs from degeneration.

Similar content being viewed by others

Data availability

All the deep-sequencing data (RNA-seq and ATAC-seq) have been uploaded to the Gene Expression Omnibus repository: GSE194243. Source data that support all findings of the study are available as supplementary data. The data that support this study are available from the corresponding author upon reasonable request. Source data are provided with this paper.

Code availability

No new code was used in this study.

References

  1. Vonsattel, J. P. & DiFiglia, M. Huntington disease. J. Neuropathol. Exp. Neurol. 57, 369–384 (1998).

    Article  CAS  PubMed  Google Scholar 

  2. Ross, C. A. et al. Huntington disease: natural history, biomarkers and prospects for therapeutics. Nat. Rev. Neurol. 10, 204–216 (2014).

    Article  CAS  PubMed  Google Scholar 

  3. Kremer, B. et al. A worldwide study of the Huntington’s disease mutation. The sensitivity and specificity of measuring CAG repeats. N. Engl. J. Med. 330, 1401–1406 (1994).

    Article  CAS  PubMed  Google Scholar 

  4. Brinkman, R. R., Mezei, M. M., Theilmann, J., Almqvist, E. & Hayden, M. R. The likelihood of being affected with Huntington disease by a particular age, for a specific CAG size. Am. J. Hum. Genet. 60, 1202–1210 (1997).

    CAS  PubMed  PubMed Central  Google Scholar 

  5. Arrasate, M., Mitra, S., Schweitzer, E. S., Segal, M. R. & Finkbeiner, S. Inclusion body formation reduces levels of mutant huntingtin and the risk of neuronal death. Nature 431, 805–810 (2004).

    Article  CAS  PubMed  Google Scholar 

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

    Article  CAS  PubMed  Google Scholar 

  7. 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 

  8. 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 

  9. Pang, Z. P. et al. Induction of human neuronal cells by defined transcription factors. Nature 476, 220–223 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  10. 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 

  11. Yoo, A. S., Staahl, B. T., Chen, L. & Crabtree, G. R. MicroRNA-mediated switching of chromatin-remodelling complexes in neural development. Nature 460, 642–646 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  12. 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 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

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

    Article  CAS  PubMed  Google Scholar 

  14. 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 

  15. 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 

  16. 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 

  17. Lee, S. W., Oh, Y. M., Lu, Y. L., Kim, W. K. & Yoo, A. S. MicroRNAs overcome cell fate barrier by reducing EZH2-controlled REST stability during neuronal conversion of human adult fibroblasts. Dev. Cell 46, 73–84 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

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

    Article  PubMed  PubMed Central  Google Scholar 

  19. Huh, C. J. et al. Maintenance of age in human neurons generated by microRNA-based neuronal conversion of fibroblasts. eLife 5, e18648 (2016).

    Article  PubMed  PubMed Central  Google Scholar 

  20. Mertens, J. et al. Directly reprogrammed human neurons retain aging-associated transcriptomic signatures and reveal age-related nucleocytoplasmic defects. Cell Stem Cell 17, 705–718 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  21. Patterson, M. et al. Defining the nature of human pluripotent stem cell progeny. Cell Res. 22, 178–193 (2012).

    Article  CAS  PubMed  Google Scholar 

  22. Miller, J. D. et al. Human iPSC-based modeling of late-onset disease via progerin-induced aging. Cell Stem Cell 13, 691–705 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  23. McCoy, M. J. & Fire, A. Z. Intron and gene size expansion during nervous system evolution. BMC Genomics 21, 360 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  24. 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 

  25. Gabel, H. W. et al. Disruption of DNA-methylation-dependent long gene repression in Rett syndrome. Nature 522, 89–93 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  26. Lu, Y. L., Liu, Y., McCoy, M. J., & Yoo, A. S. MiR-124 synergism with ELAVL3 enhances target gene expression to promote neuronal maturity. Natl Acad. Sci. USA 118, e2015454118 (2021).

    Article  CAS  Google Scholar 

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

    Article  PubMed  PubMed Central  Google Scholar 

  28. 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 

  29. You, L. et al. The role of STAT3 in autophagy. Autophagy 11, 729–739 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  30. Pensa, S. et al. Signal transducer and activator of transcription 3 and the phosphatidylinositol 3-kinase regulatory subunits p55α and p50α regulate autophagy in vivo. FEBS J. 281, 4557–4567 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  31. Jung, J. E. et al. STAT3 is a potential modulator of HIF-1-mediated VEGF expression in human renal carcinoma cells. FASEB J. 19, 1296–1298 (2005).

    Article  CAS  PubMed  Google Scholar 

  32. Chen, P. H. et al. The inhibition of microRNA-128 on IGF-1-activating mTOR signaling involves in temozolomide-induced glioma cell apoptotic death. PLoS ONE 11, e0167096 (2016).

    Article  PubMed  PubMed Central  Google Scholar 

  33. Byun, J. Y. et al. The Rac1/MKK7/JNK pathway signals upregulation of Atg5 and subsequent autophagic cell death in response to oncogenic Ras. Carcinogenesis 30, 1880–1888 (2009).

    Article  CAS  PubMed  Google Scholar 

  34. Ao, X., Zou, L. & Wu, Y. Regulation of autophagy by the Rab GTPase network. Cell Death Differ. 21, 348–358 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  35. Wang, F., Chen, X., Yu, X. & Lin, Q. Degradation of CCNB1 mediated by APC11 through UBA52 ubiquitination promotes cell cycle progression and proliferation of non-small cell lung cancer cells. Am. J. Transl. Res. 11, 7166–7185 (2019).

    CAS  PubMed  PubMed Central  Google Scholar 

  36. Ocker, M. & Schneider-Stock, R. Histone deacetylase inhibitors: signalling towards p21cip1/waf1. Int. J. Biochem. Cell Biol. 39, 1367–1374 (2007).

    Article  CAS  PubMed  Google Scholar 

  37. 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 

  38. Yoshii, S. R. & Mizushima, N. Monitoring and measuring autophagy. Int. J. Mol. Sci. 18, 1865 (2017).

    Article  PubMed Central  Google Scholar 

  39. 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 

  40. 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 

  41. 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 

  42. 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. Pharm. 12, 797541 (2021).

    Article  CAS  Google Scholar 

  43. 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 

  44. 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 

  45. Ryabaya, O. O. et al. Autophagy inhibitors chloroquine and LY294002 enhance temozolomide cytotoxicity on cutaneous melanoma cell lines in vitro. Anticancer Drugs 28, 307–315 (2017).

    Article  CAS  PubMed  Google Scholar 

  46. 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 

  47. 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 

  48. Martinez-Vicente, M. et al. Cargo recognition failure is responsible for inefficient autophagy in Huntington’s disease. Nat. Neurosci. 13, 567–576 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  49. 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 

  50. Marti, E. et al. A myriad of miRNA variants in control and Huntington’s disease brain regions detected by massively parallel sequencing. Nucleic Acids Res. 38, 7219–7235 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

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

    Article  CAS  PubMed  Google Scholar 

  52. Xue, Y., Fan, X., Yang, R., Jiao, Y. & Li, Y. miR-29b-3p inhibits post-infarct cardiac fibrosis by targeting FOS. Biosci. Rep. 40, BSR20201227 (2020).

  53. Zeng, Y., Cui, Z., Liu, J., Chen, J. & Tang, S. MicroRNA-29b-3p promotes human retinal microvascular endothelial cell apoptosis via blocking SIRT1 in diabetic retinopathy. Front. Physiol. 10, 1621 (2019).

    Article  PubMed  Google Scholar 

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

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  55. Hanna, J., Hossain, G. S. & Kocerha, J. The potential for microRNA therapeutics and clinical research. Front. Genet. 10, 478 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  56. Baumann, V. & Winkler, J. miRNA-based therapies: strategies and delivery platforms for oligonucleotide and non-oligonucleotide agents. Future Med. Chem. 6, 1967–1984 (2014).

    Article  CAS  PubMed  Google Scholar 

  57. Liu, S. et al. AIAP: a quality control and integrative analysis package to improve ATAC-seq data analysis. Genomics Proteomics Bioinformatics 19, 641–651 (2021).

    Article  PubMed  PubMed Central  Google Scholar 

  58. Chen, E. Y. et al. Enrichr: interactive and collaborative HTML5 gene list enrichment analysis tool. BMC Bioinformatics 14, 128 (2013).

    Article  PubMed  PubMed Central  Google Scholar 

  59. Kuleshov, M. V. et al. Enrichr: a comprehensive gene set enrichment analysis web server 2016 update. Nucleic Acids Res. 44, W90–W97 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

Download references

Acknowledgements

We thank the Genome Technology Access Center at Washington University for deep-sequencing experiments. HD and control brain tissues were received through the NIH NeuroBioBank. Images were created with BioRender. We thank D. H. Geschwind and R. Kawaguchi (UCLA) for their suggestions on statistics of miRNAs on gene networks. This study was supported by the following programs, grants and fellowships: Cellular and Molecular Biology Training Program (T32 GM007067) (K.C.), Cure Alzheimer’s Fund, CHDI Fund, Hereditary Disease Foundation Grant, RF1AG056296 (National Institute on Aging), R01NS107488 (National Institute of Neurological Disorders and Stroke), Farrell Foundation Fund and Mallinckrodt Scholar Award (A.S.Y.).

Author information

Authors and Affiliations

Authors

Contributions

Y.M.O. and A.S.Y. conceived and developed the idea, designed the experiments and analyzed data. Y.M.O. performed all experiments and associated assays and analyses. Y.M.O. and S.W.L. performed neuronal reprogramming throughout figures shown in the study. S.W.L. performed western blot and immunostaining for p62. Y.M.O. and W.K.K. performed ATAC-seq analysis. Y.M.O. and S.C. performed SYTOX assay of G2 analog. T.L. and B.Z. performed WGCNA. V.A.C and S.D. provided the RNAs of human brain samples. K.C. performed LGE analysis. R.E.D., S.C.P., G.A.S. and D.H.P. developed the G2 analog. Y.M.O. 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 Neuroscience thanks Johan Jakobsson and Neelroop Parikshak 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 Healthy control and HD patient fibroblasts can be directly reprogrammed into MSNs.

a. Reprogrammed cells by transduction of miR-9/9*-124+CDM immunostained for neuronal markers, TUBB3, and MSN marker, DARPP32 at PID30. Scale bars 20 μm. b. Expression of each CDM factor in reprogramming cells by immunostaining for CTIP2, DLX1, DLX2, MYT1L, and TUBB3. Non-transduced fibroblasts were used as a negative control for immunostaining. Scale bars 20 μm. c. The neuronal morphology across all samples we used in the current study stained positive for TUBB3 successfully undergo direct conversion by miR-9/9*-124+CDM. Scale bars 20 μm. d. Left, representative images of healthy control Young, Old, Pre-HD, and HD MSNs marked by TUBB3. Images processed by CellProfiler to identify neurites and associated cell soma. Scale bars, 100 μm. Right, the measurement of mean neurite length and mean number of neurite branches in reprogrammed Healthy control Young, Old, Pre-HD, and HD-MSNs at PID21 and PID35 (n = 24 individual’s reprogrammed MSNs). Statistical significance was determined using one-way ANOVA, ****p < 0.0001, ***p < 0.001 (Old vs. HD in neurite length p = 0.0007, Old vs. HD in neurite branches p = 0.0001, Pre-HD vs. HD in neurite branches p = 0.0002), **p < 0.01 (Young vs. HD in neurite length p = 0.0012, Pre-HD vs. HD in neurite length p = 0.0026), ns, not significant. Mean±s.e.m.

Source data

Extended Data Fig. 2 Genetic networks altered in HD-MSNs by WGCNA.

a, The signed association of protein-coding genes with Age, Symptomatic onset, and Sex condition of Huntington’s disease. Modules with positive values indicate increased expression in HD-MSNs; modules with negative values indicate decreased expression in HD-MSNs. The red dotted lines indicate correlation values of 0.7 or -0.7 with p = 10-7 for age, p = 10-6 for symptomatic onset, and p = 10-6 for sex. b, An expression heatmap of our selected modules of HD samples (blue, lightcyan1, brown, and greenyellow) from WGCNA. c, Pathways enriched in downregulated (log2FC < -1) or upregulated (log2FC > 1) genes in the greenyellow, blue and lightcyan1 module of Huntington’s disease by BioPlanet analysis. d, The signed association of protein-coding genes with age and sex of healthy control. Modules with positive values indicate increased expression in Old-MSNs; modules with negative values indicate decreased expression in Old-MSNs. The red dotted lines indicate correlation values of 0.7 or -0.7 with p = 10-7 for age and p = 10-6 for sex condition. e, Pathways enriched in upregulated (log2FC > 1) genes in the honeydew1 module of healthy control by BioPlanet analysis. f, Summary preservation statistics as a function of the module size. Left: the composite preservation statistic (Zsummary), Middle: the connectivity statistics (Zconnectivity), Right: the density statistics (Zdensity). Each point represents a module, labeled by color. The dashed blue and green lines indicate the thresholds Z = 2 and Z = 10, respectively. g. Representative images of MSNs expressing the tandem monomeric mCherry-GFP-LC3 reporter immunostained for TUBB3. Autophagosome (that is, mCherry + , GFP + puncta) and autolysosome (that is, mCherry + , GFP − puncta) compartments in MSNs from HD patients and control individual. Scale bar 20 μm. h. Reprogrammed cells immunostained for p62 and TUBB3 from independent HD and healthy control lines. Scale bar 20 μm. i. Immunoblot analysis for p62 and GAPDH of three independent pre-HD- and three independent HD-MSN lines at PID26. P62 signal intensities were normalized by GAPDH signals and relative fold changes in HD-MSNs were calculated over pre-HD-MSNs (**p = 0.0019 by two-tailed unpaired t-test; Mean±s.e.m.).

Source data

Extended Data Fig. 3 The treatment of autophagy inhibitor or inducer in reprogrammed MSNs.

a, Immunoblot of p62 and GAPDH in pre-HD-MSNs treated with DMSO or 50 μM LY294002 and HD-MSNs treated with DMSO or 0.5uM G2-115 at PID26 (left). p62 Intensity values were normalized by GAPDH intensities and the relative fold change over DMSO condition was calculated from immunoblot images of three independent Pre-HD-MSNs (middle, *p = 0.0277) and HD-MSNs (*p = 0.0460). Each dot represents one individual’s reprogrammed MSN. b, left: representative images of pre-HD-MSNs treated with DMSO or 50 μM LY294002 and HD-MSNs treated with DMSO or 0.5 uM G2-115 (PID 30). MSNs express the tandem monomeric RFP-GFP-LC3 reporter. Scale bars 20 μm. right: quantification of autophagosome (that is, mCherry + , GFP + puncta) and autolysosome (that is, mCherry + , GFP − puncta) compartments from three independent Pre-HD-MSNs (***p = 0.0004) and HD-MSNs (***p = 0.0005). Measurements were performed in cells having at least 3 puncta per cell (from more than 50 cells per MSN line). Each dot represents one individual’s reprogrammed MSNs. c. Synthetic route for the preparation of G2-115. d, Measurement of neuronal cell death of HD.40-MSNs with the treatment of DMSO or three different concentrations of G2 compound at PID 30 by Caspase-3/7 reagents (left, *p = 0.0199, ****p < 0.0001) or Annexin V reagents (right, *p = 0359, *p = 0.0326) (n = 4-8 cultures). For all figures shown, statistical significance was determined using two-tailed unpaired t-test (a,b) and one-way ANOVA (d); ****p < 0.0001, ***p < 0.001, *p < 0.05. Mean±s.e.m.

Source data

Extended Data Fig. 4 Pre-HD-MSNs and HD-MSNs display differential chromatin accessibilities.

a, A heatmaps showing gene expression levels for DEGs that positively correlated with signal intensities of ATAC-seq in their promoter regions. Signal intensity is based on normalized CPM values. Data are shown as Z-score normalized log2CPM (adjusted p < 0.05, │log2FC│>1). b, c, GO terms associated with opened and upregulated genes (b) and closed and downregulated genes (c). d, Transcription regulators predicted as upstream regulators of the brown module (Huntington’s disease) and the lavenderblush3 (healthy control) by Ingenuity Pathway Analysis (IPA). e, Mature microRNAs predicted as an upstream regulator of the brown module (Huntington’s disease) and the lavenderblush3 (healthy control) by Ingenuity Pathway Analysis (IPA). f, MicroRNAs predicted as an upstream regulator of the brown module (Huntington’s disease) by miRTarBase and TargetScan. g, Pathways enriched in target genes of miR-29b-3p in the brown module (Huntington’s disease) by BioPlanet analysis.

Extended Data Fig. 5 miR-29b-3p expression changes by miR-29-3p inhibitor or miR-29b overexpression lentivirus transduction.

a, RT-qPCR analysis of mature miR-29-3p expression levels in four HD-MSNs with control or miR-29b-3p inhibitor at PID26 (n = 4 individual’s reprogrammed HD-MSNs, ***p = 0.0002 by two-tailed unpaired t-test; Mean±s.e.m.). b, Pre-HD-MSNs expressing RFP only or RFP-miR-29b immunostained with RFP and DAPI. (Scale bars, 20 μm). c, RT-qPCR analysis of mature miR-29-3p expression levels in three independent pre-HD-MSNs expressing control or miR-29b. Each dot represents one individual’s reprogrammed MSNs. (**p = 0.0019 by two-tailed unpaired t-test; Mean±s.e.m).

Source data

Extended Data Fig. 6 The autophagy-related genes regulated by STAT3 in HD-MSNs.

a, The target genes of miR-29b-3p functionally related to autophagy in the brown module. Visualized by NetworkAnalyst. b, A heatmap of representation of ATAC-seq signal intensities for autophagy-related genes that contained STAT3 binding site in the closed DARs in HD-MSNs (n = 8-9 per group). c, Percentage of decrease of STAT3 levels in old versus young Ctrl-MSNs, and HD-MSNs versus pre-HD-MSNs, replotted from Fig. 7(f). Each dot represents one individual’s reprogrammed MSNs. (****p < 0.0001 by two-tailed unpaired t-test; Mean±s.e.m.). d, RT-qPCR analysis of STAT3 mRNA levels in three independent pre-HD-MSN lines with shControl or shSTAT3 at PID26. Each dot represents one individual’s reprogrammed Pre-HD-MSNs. (****p < 0.0001 by two-tailed unpaired t-test; Mean±s.e.m.). e, Western bot for STAT3 expression in human adult fibroblasts with shControl or shSTAT3. f, RT-qPCR analysis of STAT3 mRNA levels in three independent HD-MSN lines with Control or STAT3 overexpression at PID26. Each dot represents one individual’s reprogrammed HD-MSNs. (*p = 0.0143 by two-tailed unpaired t-test; Mean±s.e.m.).

Source data

Extended Data Fig. 7 Prediction of DAR proximal to miR-29b and HD-specific phenotype in fibroblasts.

a, Predicted transcription binding sites for the DAR proximal to miR29B1 (chr7:130,878,800-130,879,437). Image from UCSC Genome Browser on Human (GRCh38/hg38). JASPAR CORE 2022, Minimum Score: 500. b, Quantification of Sytox-positive cells, CYTO-ID signal, STAT3 expression, and miR-29b-3p expression in fibroblasts from healthy control Young/Old, Pre-HD, and HD (n = 14 or 19 individual fibroblasts, statistical significance was determined by one-way ANOVA. ns, not significant; Mean±s.e.m.).

Supplementary information

Supplementary Information

Supplementary Table 1. The list of six pre-symptomatic patients with HD, six HD patient fibroblasts and 12 healthy control fibroblasts

Reporting Summary

Supplementary Table 2

Detailed information on pathway enrichment analysis

Source data

Source Data Fig. 1

Statistical Source Data

Source Data Fig. 2

Statistical Source Data

Source Data Fig. 3

Statistical Source Data

Source Data Fig. 4

Statistical Source Data

Source Data Fig. 4

Unprocessed image of western blot

Source Data Fig. 5

Statistical Source Data

Source Data Fig. 6

Statistical Source Data

Source Data Fig. 7

Statistical Source Data

Source Data Extended Data Fig. 1

Statistical Source Data

Source Data Extended Data Fig. 2

Unprocessed image of western blot

Source Data Extended Data Fig. 3

Unprocessed image of western blot

Source Data Extended Data Fig. 5

Statistical Source Data

Source Data Extended Data Fig. 6

Statistical Source Data

Source Data Extended Data Fig. 6

Unprocessed image of western blot

Rights and permissions

Springer Nature or its licensor 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

Oh, Y.M., Lee, S.W., Kim, W.K. et al. Age-related Huntington’s disease progression modeled in directly reprogrammed patient-derived striatal neurons highlights impaired autophagy. Nat Neurosci 25, 1420–1433 (2022). https://doi.org/10.1038/s41593-022-01185-4

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41593-022-01185-4

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