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:

Induction of functional dopamine neurons from human astrocytes in vitro and mouse astrocytes in a Parkinson's disease model

This article has been updated

Abstract

Cell replacement therapies for neurodegenerative disease have focused on transplantation of the cell types affected by the pathological process. Here we describe an alternative strategy for Parkinson's disease in which dopamine neurons are generated by direct conversion of astrocytes. Using three transcription factors, NEUROD1, ASCL1 and LMX1A, and the microRNA miR218, collectively designated NeAL218, we reprogram human astrocytes in vitro, and mouse astrocytes in vivo, into induced dopamine neurons (iDANs). Reprogramming efficiency in vitro is improved by small molecules that promote chromatin remodeling and activate the TGFβ, Shh and Wnt signaling pathways. The reprogramming efficiency of human astrocytes reaches up to 16%, resulting in iDANs with appropriate midbrain markers and excitability. In a mouse model of Parkinson's disease, NeAL218 alone reprograms adult striatal astrocytes into iDANs that are excitable and correct some aspects of motor behavior in vivo, including gait impairments. With further optimization, this approach may enable clinical therapies for Parkinson's disease by delivery of genes rather than cells.

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

Access options

Rent or buy this article

Prices vary by article type

from$1.95

to$39.95

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

Figure 1: Human astrocytes can be reprogrammed with NeAL218.
Figure 2: Small molecules increase maturation of human-astrocyte-derived iDANs.
Figure 3: NeAL218 reprograms astrocytes into iDANs in vivo.
Figure 4: Reprogramming of adult striatal astrocytes into functional iDANs by NeAL218 in vivo.
Figure 5: iDANsNeAL218 induce behavioral recovery in a mouse model of Parkinson's disease.

Similar content being viewed by others

Accession codes

Primary accessions

Gene Expression Omnibus

Change history

  • 19 April 2017

    In the version of this article initially published, in the paragraph before the Discussion, the sentence that read in part “…the very high levels of DA that are required in the synapses to induce circling behavior…” should have read “…the very high levels of DA that are required in the synapses to reduce circling behavior…” The error has been corrected for the print, PDF and HTML versions of this article.

References

  1. DeMaagd, G. & Philip, A. Parkinson's disease and its management. Part 2: Introduction to the pharmacotherapy of parkinson's disease, with a focus on the use of dopaminergic agents. P&T 40, 591–600 (2015).

    Google Scholar 

  2. Brundin, P. et al. Intracerebral grafting of dopamine neurons. Experimental basis for clinical trials in patients with Parkinson's disease. Ann. NY Acad. Sci. 495, 473–496 (1987).

    CAS  PubMed  Google Scholar 

  3. Lindvall, O. et al. Fetal dopamine-rich mesencephalic grafts in Parkinson's disease. Lancet 2, 1483–1484 (1988).

    CAS  PubMed  Google Scholar 

  4. Arenas, E. Towards stem cell replacement therapies for Parkinson's disease. Biochem. Biophys. Res. Commun. 396, 152–156 (2010).

    CAS  PubMed  Google Scholar 

  5. Arenas, E., Denham, M. & Villaescusa, J.C. How to make a midbrain dopaminergic neuron. Development 142, 1918–1936 (2015).

    CAS  PubMed  Google Scholar 

  6. Takahashi, K. & Yamanaka, S. Induction of pluripotent stem cells from mouse embryonic and adult fibroblast cultures by defined factors. Cell 126, 663–676 (2006).

    CAS  PubMed  Google Scholar 

  7. Vierbuchen, T. et al. Direct conversion of fibroblasts to functional neurons by defined factors. Nature 463, 1035–1041 (2010).

    CAS  PubMed  PubMed Central  Google Scholar 

  8. Ladewig, J., Koch, P. & Brustle, O. Leveling Waddington: the emergence of direct reprogramming and the loss of cell fate hierarchies. Nat. Rev. Mol. Cell Biol. 14, 225–236 (2013).

    CAS  PubMed  Google Scholar 

  9. Chen, G. et al. In vivo reprogramming for brain and spinal cord repair(1,2,3). eNeuro. 2, e0106–15.2015 (2015).

    Google Scholar 

  10. Caiazzo, M. et al. Direct generation of functional dopaminergic neurons from mouse and human fibroblasts. Nature 476, 224–227 (2011).

    CAS  PubMed  Google Scholar 

  11. Kim, J. et al. Functional integration of dopaminergic neurons directly converted from mouse fibroblasts. Cell Stem Cell 9, 413–419 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  12. Liu, X. et al. Direct reprogramming of human fibroblasts into dopaminergic neuron-like cells. Cell Res. 22, 321–332 (2012).

    CAS  PubMed  Google Scholar 

  13. Pfisterer, U. et al. Direct conversion of human fibroblasts to dopaminergic neurons. Proc. Natl. Acad. Sci. USA 108, 10343–10348 (2011).

    CAS  PubMed  PubMed Central  Google Scholar 

  14. Addis, R.C. et al. Efficient conversion of astrocytes to functional midbrain dopaminergic neurons using a single polycistronic vector. PLoS One 6, e28719 (2011).

    CAS  PubMed  PubMed Central  Google Scholar 

  15. Heins, N. et al. Glial cells generate neurons: the role of the transcription factor Pax6. Nat. Neurosci. 5, 308–315 (2002).

    CAS  PubMed  Google Scholar 

  16. Heinrich, C. et al. Directing astroglia from the cerebral cortex into subtype specific functional neurons. PLoS Biol. 8, e1000373 (2010).

    PubMed  PubMed Central  Google Scholar 

  17. Heinrich, C. et al. Generation of subtype-specific neurons from postnatal astroglia of the mouse cerebral cortex. Nat. Protoc. 6, 214–228 (2011).

    CAS  PubMed  Google Scholar 

  18. Torper, O. et al. Generation of induced neurons via direct conversion in vivo. Proc. Natl. Acad. Sci. USA 110, 7038–7043 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  19. Su, Z., Niu, W., Liu, M.L., Zou, Y. & Zhang, C.L. In vivo conversion of astrocytes to neurons in the injured adult spinal cord. Nat. Commun. 5, 3338 (2014).

    PubMed  Google Scholar 

  20. Guo, Z. et al. In vivo direct reprogramming of reactive glial cells into functional neurons after brain injury and in an Alzheimer's disease model. Cell Stem Cell 14, 188–202 (2014).

    CAS  PubMed  Google Scholar 

  21. Heinrich, C. et al. Sox2-mediated conversion of NG2 glia into induced neurons in the injured adult cerebral cortex. Stem Cell Reports 3, 1000–1014 (2014).

    CAS  PubMed  PubMed Central  Google Scholar 

  22. Torper, O. et al. In vivo reprogramming of striatal NG2 glia into functional neurons that integrate into local host circuitry. Cell Rep. 12, 474–481 (2015).

    CAS  PubMed  PubMed Central  Google Scholar 

  23. Gascón, S. et al. Identification and successful negotiation of a metabolic checkpoint in direct neuronal reprogramming. Cell Stem Cell 18, 396–409 (2016).

    PubMed  Google Scholar 

  24. Brambrink, T. et al. Sequential expression of pluripotency markers during direct reprogramming of mouse somatic cells. Cell Stem Cell 2, 151–159 (2008).

    CAS  PubMed  PubMed Central  Google Scholar 

  25. Esteban, M.A. et al. Vitamin C enhances the generation of mouse and human induced pluripotent stem cells. Cell Stem Cell 6, 71–79 (2010).

    CAS  PubMed  Google Scholar 

  26. Chambers, S.M. et al. Highly efficient neural conversion of human ES and iPS cells by dual inhibition of SMAD signaling. Nat. Biotechnol. 27, 275–280 (2009).

    CAS  PubMed  PubMed Central  Google Scholar 

  27. Wernig, M. et al. Tau EGFP embryonic stem cells: an efficient tool for neuronal lineage selection and transplantation. J. Neurosci. Res. 69, 918–924 (2002).

    CAS  PubMed  Google Scholar 

  28. Kriks, S. et al. Dopamine neurons derived from human ES cells efficiently engraft in animal models of Parkinson's disease. Nature 480, 547–551 (2011).

    CAS  PubMed  PubMed Central  Google Scholar 

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

    CAS  PubMed  PubMed Central  Google Scholar 

  30. Huang, T., Liu, Y., Huang, M., Zhao, X. & Cheng, L. Wnt1-cre-mediated conditional loss of Dicer results in malformation of the midbrain and cerebellum and failure of neural crest and dopaminergic differentiation in mice. J. Mol. Cell Biol. 2, 152–163 (2010).

    CAS  PubMed  Google Scholar 

  31. Huangfu, D. et al. Induction of pluripotent stem cells by defined factors is greatly improved by small-molecule compounds. Nat. Biotechnol. 26, 795–797 (2008).

    CAS  PubMed  PubMed Central  Google Scholar 

  32. Pennarossa, G. et al. Brief demethylation step allows the conversion of adult human skin fibroblasts into insulin-secreting cells. Proc. Natl. Acad. Sci. USA 110, 8948–8953 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  33. Chung, T.L. et al. Vitamin C promotes widespread yet specific DNA demethylation of the epigenome in human embryonic stem cells. Stem Cells 28, 1848–1855 (2010).

    CAS  PubMed  Google Scholar 

  34. Liu, X. et al. Sequential introduction of reprogramming factors reveals a time-sensitive requirement for individual factors and a sequential EMT-MET mechanism for optimal reprogramming. Nat. Cell Biol. 15, 829–838 (2013).

    CAS  PubMed  Google Scholar 

  35. La Manno, G. et al. Molecular diversity of midbrain development in mouse, human, and stem cells. Cell 167, 566–580.e19 (2016).

    CAS  PubMed  PubMed Central  Google Scholar 

  36. Koyano-Nakagawa, N., Kim, J., Anderson, D. & Kintner, C. Hes6 acts in a positive feedback loop with the neurogenins to promote neuronal differentiation. Development 127, 4203–4216 (2000).

    CAS  PubMed  Google Scholar 

  37. Jhas, S. et al. Hes6 inhibits astrocyte differentiation and promotes neurogenesis through different mechanisms. J. Neurosci. 26, 11061–11071 (2006).

    CAS  PubMed  PubMed Central  Google Scholar 

  38. Dillon-Carter, O., Conejero, C., Tornatore, C., Poltorak, M. & Freed, W.J. N18-RE-105 cells: differentiation and activation of p53 in response to glutamate and adriamycin is blocked by SV40 large T antigen tsA58. Cell Tissue Res. 291, 191–205 (1998).

    CAS  PubMed  Google Scholar 

  39. Lundblad, M., Picconi, B., Lindgren, H. & Cenci, M.A. A model of L-DOPA-induced dyskinesia in 6-hydroxydopamine lesioned mice: relation to motor and cellular parameters of nigrostriatal function. Neurobiol. Dis. 16, 110–123 (2004).

    CAS  PubMed  Google Scholar 

  40. Darmopil, S., Muñetón-Gómez, V.C., de Ceballos, M.L., Bernson, M. & Moratalla, R. Tyrosine hydroxylase cells appearing in the mouse striatum after dopamine denervation are likely to be projection neurones regulated by L-DOPA. Eur. J. Neurosci. 27, 580–592 (2008).

    PubMed  Google Scholar 

  41. Masuda, M. et al. Postnatal development of tyrosine hydroxylase mRNA-expressing neurons in mouse neostriatum. Eur. J. Neurosci. 34, 1355–1367 (2011).

    PubMed  Google Scholar 

  42. Hu, G. et al. New fluorescent substrate enables quantitative and high-throughput examination of vesicular monoamine transporter 2 (VMAT2). ACS Chem. Biol. 8, 1947–1954 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  43. Freyberg, Z. et al. Mechanisms of amphetamine action illuminated through optical monitoring of dopamine synaptic vesicles in Drosophila brain. Nat. Commun. 7, 10652 (2016).

    CAS  PubMed  PubMed Central  Google Scholar 

  44. Brooks, S.P. & Dunnett, S.B. Tests to assess motor phenotype in mice: a user's guide. Nat. Rev. Neurosci. 10, 519–529 (2009).

    CAS  PubMed  Google Scholar 

  45. Bagga, V., Dunnett, S.B. & Fricker, R.A. The 6-OHDA mouse model of Parkinson's disease - Terminal striatal lesions provide a superior measure of neuronal loss and replacement than median forebrain bundle lesions. Behav. Brain Res. 288, 107–117 (2015).

    CAS  PubMed  Google Scholar 

  46. Fasano, A., Aquino, C.C., Krauss, J.K., Honey, C.R. & Bloem, B.R. Axial disability and deep brain stimulation in patients with Parkinson disease. Nat. Rev. Neurol. 11, 98–110 (2015).

    PubMed  Google Scholar 

  47. Kurz, M.J. et al. A chronic mouse model of Parkinson's disease has a reduced gait pattern certainty. Neurosci. Lett. 429, 39–42 (2007).

    CAS  PubMed  PubMed Central  Google Scholar 

  48. Bonito-Oliva, A., Masini, D. & Fisone, G. A mouse model of non-motor symptoms in Parkinson's disease: focus on pharmacological interventions targeting affective dysfunctions. Front. Behav. Neurosci. 8, 290 (2014).

    PubMed  PubMed Central  Google Scholar 

  49. Amende, I. et al. Gait dynamics in mouse models of Parkinson's disease and Huntington's disease. J. Neuroeng. Rehabil. 2, 20 (2005).

    PubMed  PubMed Central  Google Scholar 

  50. Wang, H., Li, X., Gao, S., Sun, X. & Fang, H. Transdifferentiation via transcription factors or microRNAs: current status and perspective. Differentiation 90, 69–76 (2015).

    CAS  PubMed  Google Scholar 

  51. Dickson, D.W. Parkinson's disease and parkinsonism: neuropathology. Cold Spring Harb. Perspect. Med. 2, a009258 (2012).

    PubMed  PubMed Central  Google Scholar 

  52. Zeisel, A. et al. Brain structure. Cell types in the mouse cortex and hippocampus revealed by single-cell RNA-seq. Science 347, 1138–1142 (2015).

    CAS  PubMed  Google Scholar 

  53. Love, M.I., Huber, W. & Anders, S. Moderated estimation of fold change and dispersion for RNA-seq data with DESeq2. Genome Biol. 15, 550 (2014).

    PubMed  PubMed Central  Google Scholar 

  54. Wang, J., Lin, W., Popko, B. & Campbell, I.L. Inducible production of interferon-gamma in the developing brain causes cerebellar dysplasia with activation of the Sonic hedgehog pathway. Mol. Cell. Neurosci. 27, 489–496 (2004).

    CAS  PubMed  Google Scholar 

  55. Ekstrand, M.I. et al. Progressive parkinsonism in mice with respiratory-chain-deficient dopamine neurons. Proc. Natl. Acad. Sci. USA 104, 1325–1330 (2007).

    CAS  PubMed  PubMed Central  Google Scholar 

Download references

Acknowledgements

We thank the members of the Arenas laboratory for help and suggestions; J. Söderlund and A. Nanni for technical and secretarial assistance; and the SciLife National Genomic Infrastructure (Stockholm) for RNA sequencing. This work was supported by grants from the Swedish Research Council (VR: DBRM, 2011-3116/3318 and 2016-01526), Swedish Foundation for Strategic Research (SRL), EU (NeuroStemcellRepair and DDPDgenes), Karolinska Institutet, Strat Regen, Hjärnfonden (FO2013:0108, FO2015:0202) and Cancerfonden (CAN 2016/572) to E.A.; VR (2012-13482 and 2015-02886), StratNeuro, Parkinsonfonden, Hjärnfonden and KI/NIH to G.F.; VR (2013-3080), EU (PAINCAGE), Hjärnfonden, NovoNordisk Foundation and the European Research Council (“Secret Cells”) to T.H.; and New York Stem Cell Foundation, NIH and CIRM to M.W. Support to P.R.d.V.C. was provided by VR (524-2011-962) and EMBO (ALTF583-2011); to R.A.R. by EMBO (ALTF596-2014) and Marie Curie (EMBOCOFUND2012, GA-2012-600394); to D.M. by KI and by the Brasilian Ministry of Education (CAPES) and to E.M.-M. by the Spanish Ministry of Education (José Castillejo). The authors acknowledge support from Science for Life Laboratory, the Knut and Alice Wallenberg Foundation, the National Genomics Infrastructure funded by the Swedish Research Council, and Uppsala Multidisciplinary Center for Advanced Computational Science for assistance with massively parallel sequencing and access to the UPPMAX computational infrastructure.

Author information

Authors and Affiliations

Authors

Contributions

E.A. and P.R.d.V.C. conceived the experiments and wrote the manuscript; P.R.d.V.C., R.A.R., G.S., D.M., E.M.-M., E.M.T., G.L.M., M.F., C.P., Y.-H.N. and S.P.S. performed the experiments; S. L., M.W., T.H., G.F. and E.A. provided expertise and funding.

Corresponding author

Correspondence to Ernest Arenas.

Ethics declarations

Competing interests

The authors declare no competing financial interests.

Integrated supplementary information

Supplementary Figure 1 Characterization of hIA cultures and midbrain markers after iDAN reprogramming.

a. Uninfected hIAs do not spontaneously become TH+ in MP media at day 17. b. hIAs infected with GFP show a high infection efficiency in MP media at day 17 c. Gene expression analysis by real-time q-PCR revealed an increase in multiple mDA markers in hIA cultures treated with AL-MP, AL218-MP or NeAL218-MP at day 10. Data are shown as fold increase over control cells. * p<0.05 and ** p<0.01, n≥3, n.d.: not detectable

Supplementary Figure 2 Properties of hIA-derived iDANsNeAL218-RTMP cells.

a. Percentage of hIAs that become TH+ after infection with NeAL218 in RTMP conditions at day 4(n=3) and day 10(n=3). b. hIAs infected with NeAL218 in RTMP conditions at day 4 (scale bar 50μm) c. Illustration of the experiment for doxycycline withdrawal. d. hIAs infected with NeAL218 and cultured in RTMP for 13 days, but with only 7 or 10 days of doxycycline supplementation (scale bars 50μm). e. Bright field image of a representative hIA infected with NeAL218 and treated with RTMP media, probed by patch-clamp electrophysiology with its current profile (f) depicted in voltage-clamp mode. g. Current clamp recordings show the presence of hIA-derived iDANsNeAL218-RTMP capable of generating single action potential upon current injection (5, 20 and 30 pA). h. Vertical box plots, reflecting amplitudes of outward and inward currents in control (n=16) and NeAL218-infected hIAs cells with RTMP protocol (n=12). Whitney Rank Sum Test for outward currents p = 0,043; for inward currents p < 0,001. i. Pie chart depicting the fraction of hIA-derived iDANsNeAL218-RTMP generating single action potentials among the 12 cells probed * p<0.05 and ** p<0.01.

Supplementary Figure 3 Properties of hPA-derived control and iDANsNeAL218-RTMP cells

a. Efficient expression of GFP by hPAs in RTMP media at day 17. b. Uninfected hPAs do not become TH+ in RTMP media at day 17 (scale bar 50μm). c. Representative voltage-clamp (top) and current-clamp (bottom) recordings from control-RTMP hPAs cells. Control cells did not generate any action potential (AP). d. Comparison of gene expression by real-time q-PCR of hPAs treated with control-RTMP or NeAL218-RTMP at day 10. Data are mean ± s.d. FOXA2 was compared to untreated cells at day 0, as Foxa2 mRNA was undetectable in control-RTMP cells at day 10.

Supplementary Figure 4 Characterization of the 6-OHDA lesion model, transduction of the GFAP-tTA mice and reporter activity of the GFAP-tTa;SLC6A3cre/+;Rosa26RTomato mice.

a. Ventral midbrain cryosection of GFAP-tTA adult mice infected in the striatum with NeAL218 for 13 weeks (scale bar 50μm). Comparison of the right and left substantia nigra showing the effect of the 6-OHDA injection in the ipsilateral medial forebrain bundle. b. Striatal cryosections of GFAP-tTA adult mice 2 weeks after unilateral 6-OHDA (scale bar 100μm) showing loss of TH and increase in GFAP in the ipsilateral striatum. c. Striatal cryosections of GFAP-tTA adult mice infected with GFP for 2 weeks (scale bar 25μm). d. Striatal cryosections of GFAP-tTA adult mice infected with GFP for 13 weeks (scale bar 10μm). e. Striatal cryosections of GFAP- tTa;SLC6A3cre/+;Rosa26RTomato adult mice infected with GFP for 5 weeks (scale bar 10μm). f. Cryosection of adult GFAP-tTa;SLC6A3cre/+;Rosa26RTomato mice showing the functionality of the reporter mice in the intact contralateral side of 6-OHDA lesioned mice infected with GFP or NeAL218 for 5 weeks (scale bar 10μm).

Supplementary Figure 5 Characterization of in vivo iDAN reprogramming by NeAL218.

a. Diagram depicting am alternative strategy used to visualize in vivo reprogrammed mature iDANs: GFAP-tTA; SLC6A3Cre/+ mice and flex-tdTomato virus. b. Time-lapse microphotographs showing release of a fluorescent dopamine derivative, FFN 206, upon KCl-induced depolarization in acute brain slices of GFAP-tTA adult mice, 5 weeks after NeAL218 or GFP injection (scale bar 25μm). c-e. Electrophysiological recordings at 5 weeks showing NeAL218-induced SLC6A3Tomato+ cells unable to generate action potentials in voltage clamp mode (c: voltage clamp profile, n=19 out of 28) or able to generate single action potentials upon somatic current injection (n=4 out of 28 in d and e, showing voltage-clamp and current clamp profiles, respectively). f. Control GFP+ cells recorded in voltage-clamp configuration (polarization for 100 ms from -100 to + 50 mV, n=9; 2 examples are presented) showed outward voltage-gated currents but no inward voltage gated (sodium/calcium) currents. g. Net rotations after systemic amphetamine administration to control and 6-OHDA lesioned GFAP-tTA mice ± intrastriatatal GFP or NeAL218 for 5, 9 or 13 weeks. GFP n = 6; NeAL218, n = 7. Two Way RM ANOVA with time effect F(4,44) = 34,75; p < 0,0001 and subject matching F(11, 44) = 7,908; p < 0,0001 followed by Fisher’s comparison (for GFP: naïve vs lesion p = 0,0003; W5 vs lesion p = 0,0125; W9 vs lesion p = 0,0039 and W13 vs lesion p = 0,0122 and for NeAL218: naïve vs lesion p < 0,0001; W5 vs lesion p = 0,0813; W9 vs lesion p = 0,0011 and W13 vs lesion p = 0,0649).

Supplementary information

Supplementary Text and Figures

Supplementary Figures 1–5 (PDF 1199 kb)

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Rivetti di Val Cervo, P., Romanov, R., Spigolon, G. et al. Induction of functional dopamine neurons from human astrocytes in vitro and mouse astrocytes in a Parkinson's disease model. Nat Biotechnol 35, 444–452 (2017). https://doi.org/10.1038/nbt.3835

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/nbt.3835

This article is cited by

Search

Quick links

Nature Briefing: Translational Research

Sign up for the Nature Briefing: Translational Research newsletter — top stories in biotechnology, drug discovery and pharma.

Get what matters in translational research, free to your inbox weekly. Sign up for Nature Briefing: Translational Research