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:

Cerebellar granule cell replenishment postinjury by adaptive reprogramming of Nestin+ progenitors

Abstract

Regeneration of several organs involves adaptive reprogramming of progenitors, but the intrinsic capacity of the developing brain to replenish lost cells remains largely unknown. Here we found that the developing cerebellum has unappreciated progenitor plasticity, since it undergoes near full growth and functional recovery following acute depletion of granule cells, the most plentiful neuron population in the brain. We demonstrate that following postnatal ablation of granule cell progenitors, Nestin-expressing progenitors, specified during mid-embryogenesis to produce astroglia and interneurons, switch their fate and generate granule neurons in mice. Moreover, Hedgehog signaling in two Nestin-expressing progenitor populations is crucial not only for the compensatory replenishment of granule neurons but also for scaling interneuron and astrocyte numbers. Thus, we provide insights into the mechanisms underlying robustness of circuit formation in the cerebellum and speculate that adaptive reprogramming of progenitors in other brain regions plays a greater role than appreciated in developmental regeneration.

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

Figure 1: Irradiation of cerebella at P1 results in a major loss of the EGL by P3, but growth largely recovers and motor behavior is intact at P30.
Figure 2: The EGL is progressively replenished after injury.
Figure 3: NEPs populate the EGL and generate granule cells and rescue growth.
Figure 4: PCL NEPs expand and migrate into the EGL after injury.
Figure 5: WM NEPs delay their expansion and production of interneurons and astroglia.
Figure 6: SHH signaling is required for NEPs to expand and populate the EGL and for substantial recovery of the cerebellum.

Similar content being viewed by others

Accession codes

Primary accessions

Gene Expression Omnibus

References

  1. Jessen, K.R., Mirsky, R. & Arthur-Farraj, P. The role of cell plasticity in tissue repair: adaptive cellular reprogramming. Dev. Cell 34, 613–620 (2015).

    CAS  PubMed  Google Scholar 

  2. Azevedo, F.A. et al. Equal numbers of neuronal and nonneuronal cells make the human brain an isometrically scaled-up primate brain. J. Comp. Neurol. 513, 532–541 (2009).

    PubMed  Google Scholar 

  3. Herculano-Houzel, S., Mota, B. & Lent, R. Cellular scaling rules for rodent brains. Proc. Natl. Acad. Sci. USA 103, 12138–12143 (2006).

    CAS  PubMed  PubMed Central  Google Scholar 

  4. Steinlin, M. The cerebellum in cognitive processes: supporting studies in children. Cerebellum 6, 237–241 (2007).

    PubMed  Google Scholar 

  5. Tavano, A. et al. Disorders of cognitive and affective development in cerebellar malformations. Brain 130, 2646–2660 (2007).

    PubMed  Google Scholar 

  6. Fatemi, S.H. et al. Consensus paper: pathological role of the cerebellum in autism. Cerebellum 11, 777–807 (2012).

    PubMed  PubMed Central  Google Scholar 

  7. Altman, J. & Bayer, S.A. Development of the Cerebellar System in Relation to Its Evolution, Structure, and Functions (CRC Press, 1997).

  8. Rakic, P. & Sidman, R.L. Histogenesis of cortical layers in human cerebellum, particularly the lamina dissecans. J. Comp. Neurol. 139, 473–500 (1970).

    CAS  PubMed  Google Scholar 

  9. Wang, S.S., Kloth, A.D. & Badura, A. The cerebellum, sensitive periods, and autism. Neuron 83, 518–532 (2014).

    CAS  PubMed  PubMed Central  Google Scholar 

  10. Hoshino, M. et al. Ptf1a, a bHLH transcriptional gene, defines GABAergic neuronal fates in cerebellum. Neuron 47, 201–213 (2005).

    CAS  PubMed  Google Scholar 

  11. Wingate, R.J. & Hatten, M.E. The role of the rhombic lip in avian cerebellum development. Development 126, 4395–4404 (1999).

    CAS  PubMed  Google Scholar 

  12. Machold, R. & Fishell, G. Math1 is expressed in temporally discrete pools of cerebellar rhombic-lip neural progenitors. Neuron 48, 17–24 (2005).

    CAS  PubMed  Google Scholar 

  13. Wang, V.Y., Rose, M.F. & Zoghbi, H.Y. Math1 expression redefines the rhombic lip derivatives and reveals novel lineages within the brainstem and cerebellum. Neuron 48, 31–43 (2005).

    CAS  PubMed  Google Scholar 

  14. Sillitoe, R.V. & Joyner, A.L. Morphology, molecular codes, and circuitry produce the three-dimensional complexity of the cerebellum. Annu. Rev. Cell Dev. Biol. 23, 549–577 (2007).

    CAS  PubMed  Google Scholar 

  15. Corrales, J.D., Blaess, S., Mahoney, E.M. & Joyner, A.L. The level of sonic hedgehog signaling regulates the complexity of cerebellar foliation. Development 133, 1811–1821 (2006).

    CAS  PubMed  Google Scholar 

  16. Lewis, P.M., Gritli-Linde, A., Smeyne, R., Kottmann, A. & McMahon, A.P. Sonic hedgehog signaling is required for expansion of granule neuron precursors and patterning of the mouse cerebellum. Dev. Biol. 270, 393–410 (2004).

    CAS  PubMed  Google Scholar 

  17. Altman, J., Anderson, W.J. & Wright, K.A. Early effects of x-irradiation of the cerebellum in infant rats: decimation and reconstitution of the external granular layer. Exp. Neurol. 24, 196–216 (1969).

    CAS  PubMed  Google Scholar 

  18. Fleming, J.T. et al. The Purkinje neuron acts as a central regulator of spatially and functionally distinct cerebellar precursors. Dev. Cell 27, 278–292 (2013).

    CAS  PubMed  Google Scholar 

  19. Joyner, A. & Sudarov, A. Genetic Neuroanatomy. in The Mouse Nervous System (eds. Watson, C., Paxinos G. & Puelles L.) 36–47 (Academic Press, 2011).

  20. Milosevic, A. & Goldman, J.E. Potential of progenitors from postnatal cerebellar neuroepithelium and white matter: lineage specified vs. multipotent fate. Mol. Cell. Neurosci. 26, 342–353 (2004).

    CAS  PubMed  Google Scholar 

  21. Alexander, T., Nolte, C. & Krumlauf, R. Hox genes and segmentation of the hindbrain and axial skeleton. Annu. Rev. Cell Dev. Biol. 25, 431–456 (2009).

    CAS  PubMed  Google Scholar 

  22. Buffo, A. & Rossi, F. Origin, lineage and function of cerebellar glia. Prog. Neurobiol. 109, 42–63 (2013).

    PubMed  Google Scholar 

  23. Li, P. et al. A population of Nestin-expressing progenitors in the cerebellum exhibits increased tumorigenicity. Nat. Neurosci. 16, 1737–1744 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  24. Parmigiani, E. et al. Heterogeneity and bipotency of astroglial-like cerebellar progenitors along the interneuron and glial lineages. J. Neurosci. 35, 7388–7402 (2015).

    CAS  PubMed  PubMed Central  Google Scholar 

  25. De Luca, A. et al. Exogenous sonic hedgehog modulates the pool of GABAergic interneurons during cerebellar development. Cerebellum 14, 72–85 (2015).

    CAS  PubMed  Google Scholar 

  26. Srivastava, D. & DeWitt, N. In vivo cellular reprogramming: the next generation. Cell 166, 1386–1396 (2016).

    CAS  PubMed  PubMed Central  Google Scholar 

  27. Pascual, M. et al. Cerebellar GABAergic progenitors adopt an external granule cell-like phenotype in the absence of Ptf1a transcription factor expression. Proc. Natl. Acad. Sci. USA 104, 5193–5198 (2007).

    CAS  PubMed  PubMed Central  Google Scholar 

  28. Millen, K.J., Steshina, E.Y., Iskusnykh, I.Y. & Chizhikov, V.V. Transformation of the cerebellum into more ventral brainstem fates causes cerebellar agenesis in the absence of Ptf1a function. Proc. Natl. Acad. Sci. USA 111, E1777–E1786 (2014).

    CAS  PubMed  PubMed Central  Google Scholar 

  29. Yamada, M. et al. Specification of spatial identities of cerebellar neuron progenitors by ptf1a and atoh1 for proper production of GABAergic and glutamatergic neurons. J. Neurosci. 34, 4786–4800 (2014).

    PubMed  PubMed Central  Google Scholar 

  30. Lee, A. et al. Isolation of neural stem cells from the postnatal cerebellum. Nat. Neurosci. 8, 723–729 (2005).

    CAS  PubMed  PubMed Central  Google Scholar 

  31. Lao, Z., Raju, G.P., Bai, C.B. & Joyner, A.L. MASTR: a technique for mosaic mutant analysis with spatial and temporal control of recombination using conditional floxed alleles in mice. Cell Reports 2, 386–396 (2012).

    CAS  PubMed  Google Scholar 

  32. Joyner, A.L. & Zervas, M. Genetic inducible fate mapping in mouse: establishing genetic lineages and defining genetic neuroanatomy in the nervous system. Dev. Dyn. 235, 2376–2385 (2006).

    PubMed  Google Scholar 

  33. Mignone, J.L., Kukekov, V., Chiang, A.S., Steindler, D. & Enikolopov, G. Neural stem and progenitor cells in nestin-GFP transgenic mice. J. Comp. Neurol. 469, 311–324 (2004).

    CAS  PubMed  Google Scholar 

  34. Legué, E., Riedel, E. & Joyner, A.L. Clonal analysis reveals granule cell behaviors and compartmentalization that determine the folded morphology of the cerebellum. Development 142, 1661–1671 (2015).

    PubMed  PubMed Central  Google Scholar 

  35. Huang, W.H. et al. Atoh1 governs the migration of postmitotic neurons that shape respiratory effectiveness at birth and chemoresponsiveness in adulthood. Neuron 75, 799–809 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  36. Ben-Arie, N. et al. Math1 is essential for genesis of cerebellar granule neurons. Nature 390, 169–172 (1997).

    CAS  PubMed  Google Scholar 

  37. Flora, A., Klisch, T.J., Schuster, G. & Zoghbi, H.Y. Deletion of Atoh1 disrupts sonic hedgehog signaling in the developing cerebellum and prevents medulloblastoma. Science 326, 1424–1427 (2009).

    CAS  PubMed  PubMed Central  Google Scholar 

  38. Klein, C., Butt, S.J., Machold, R.P., Johnson, J.E. & Fishell, G. Cerebellum- and forebrain-derived stem cells possess intrinsic regional character. Development 132, 4497–4508 (2005).

    CAS  PubMed  Google Scholar 

  39. Angot, E. et al. Chemoattractive activity of sonic hedgehog in the adult subventricular zone modulates the number of neural precursors reaching the olfactory bulb. Stem Cells 26, 2311–2320 (2008).

    CAS  PubMed  Google Scholar 

  40. Bijlsma, M.F., Borensztajn, K.S., Roelink, H., Peppelenbosch, M.P. & Spek, C.A. Sonic hedgehog induces transcription-independent cytoskeletal rearrangement and migration regulated by arachidonate metabolites. Cell. Signal. 19, 2596–2604 (2007).

    CAS  PubMed  Google Scholar 

  41. Petrova, R. & Joyner, A.L. Roles for Hedgehog signaling in adult organ homeostasis and repair. Development 141, 3445–3457 (2014).

    CAS  PubMed  PubMed Central  Google Scholar 

  42. Williams, J.P. Catch-up growth. J. Embryol. Exp. Morphol. 65 (Suppl), 89–101 (1981).

    PubMed  Google Scholar 

  43. Boersma, B. & Wit, J.M. Catch-up growth. Endocr. Rev. 18, 646–661 (1997).

    CAS  PubMed  Google Scholar 

  44. Chen, P., Johnson, J.E., Zoghbi, H.Y. & Segil, N. The role of Math1 in inner ear development: uncoupling the establishment of the sensory primordium from hair cell fate determination. Development 129, 2495–2505 (2002).

    CAS  PubMed  Google Scholar 

  45. Perl, A.K., Wert, S.E., Nagy, A., Lobe, C.G. & Whitsett, J.A. Early restriction of peripheral and proximal cell lineages during formation of the lung. Proc. Natl. Acad. Sci. USA 99, 10482–10487 (2002).

    CAS  PubMed  PubMed Central  Google Scholar 

  46. Buch, T. et al. A Cre-inducible diphtheria toxin receptor mediates cell lineage ablation after toxin administration. Nat. Methods 2, 419–426 (2005).

    CAS  PubMed  Google Scholar 

  47. Long, F., Zhang, X.M., Karp, S., Yang, Y. & McMahon, A.P. Genetic manipulation of hedgehog signaling in the endochondral skeleton reveals a direct role in the regulation of chondrocyte proliferation. Development 128, 5099–5108 (2001).

    CAS  PubMed  Google Scholar 

  48. Mao, J. et al. A novel somatic mouse model to survey tumorigenic potential applied to the Hedgehog pathway. Cancer Res. 66, 10171–10178 (2006).

    CAS  PubMed  PubMed Central  Google Scholar 

  49. Hu, Y. & Smyth, G.K. ELDA: extreme limiting dilution analysis for comparing depleted and enriched populations in stem cell and other assays. J. Immunol. Methods 347, 70–78 (2009).

    CAS  PubMed  Google Scholar 

  50. Famulski, J.K. et al. Siah regulation of Pard3A controls neuronal cell adhesion during germinal zone exit. Science 330, 1834–1838 (2010).

    CAS  PubMed  PubMed Central  Google Scholar 

  51. Girirajan, S. et al. How much is too much? Phenotypic consequences of Rai1 overexpression in mice. Eur. J. Hum. Genet. 16, 941–954 (2008).

    CAS  PubMed  Google Scholar 

Download references

Acknowledgements

We thank Z. Yang for providing the Nestin-CFP line and for discussions of unpublished data. We are grateful to B. Palikuqi (in S. Rafii's lab) for technical support with initial Nes-CFP FACS and R. Sillitoe, T. Stay and E. Lackey for their advice with performing the behavior tests. We thank K. Zaret for thoughtful comments on the manuscript, and past and present members of the Joyner laboratory for discussions. We also thank the Flow Cytometry, Bio-informatics, Center for Comparative Medicine and Pathology, and Integrated Genomics Operation (IGO) core facilities of MSKCC for outstanding technical support. We also gratefully acknowledge P. Zanzonico for his help with mouse irradiation and B. Nieman for advice; Q. Chen and the MSKCC Small-Animal Imaging Core Facility for technical services; and a Shared Resources Grant from the MSKCC Geoffrey Beene Cancer Research Center, which provided funding support for the purchase of the XRad 225Cx Microirradiator. This work was supported by grants from the Brain Tumor Center at MSKCC (to A.W.), National Brain Tumor Association (to A.L.J.) and NINDS (R01 NS092096 to A.L.J. and F32 NS086163 to A.K.L.), as well as a National Cancer Institute Cancer Center Support Grant (P30 CA008748-48).

Author information

Authors and Affiliations

Authors

Contributions

A.W. and A.L.J. conceived the project; A.W., A.K.L., N.S.B. and A.L.J. designed the research; A.W., A.K.L., N.S.B., Z.L. and D.N.S. performed the experiments; A.W., A.K.L., N.S.B. and A.L.J. analyzed the data and all authors discussed the data; A.W. and A.L.J. wrote the manuscript with contributions from all authors.

Corresponding authors

Correspondence to Alexandre Wojcinski or Alexandra L Joyner.

Ethics declarations

Competing interests

The authors declare no competing financial interests.

Integrated supplementary information

Supplementary Figure 1 Irradiation does not lead to major cell death in the SOX2+ layers.

( A- D) FIHC detection of the indicated proteins and dapi on midsagittal sections (lobule IV/V) of Non-IR (A-B) and IR (C-D) mice at the indicated ages (n=4). External granule layer (EGL), molecular layer (ML), Purkinje Cell Layer (PCL) and white matter (WM)+ internal granule layer (IGL) are delimited by yellow dotted lines. White arrowheads in C indicate TUNEL+ cells present in the PCL. Black scale bar indicates 100μm.

Supplementary Figure 2 NEPs generate interneurons (INs) and astrocytes but not granule neurons (GCs) in the WT cerebellum.

(A-H) FIHC detection of the indicated proteins and dapi on sagittal sections of the vermis (lobule III) of Nes-FlpoER/+; R26FSF-TDTom/+; Atoh1-GFP/+ (Nes-TDTom; Atoh1-GFP), Nes-FlpoER/+; R26MASTR/+ (Nes-EGFP-Cre) and Nes-CFP/+ mice at the indicated ages. (B) Nes-FlpoER given tamoxifen at P0 initially (P1) mainly marks SOX2+ progenitors, and not Atoh1+ (GFP+) GCPs (backwards arrow in Fig. 1A), and give rise at P8 to S100+ astrocytes (astro), Bergmann glia (Bg) and PAX2+ interneurons (INs) (arrows C, E, F), only rare PAX6+ GCPs or GCs (backward arrows in D)(See Fig. 3 for quantifications). Higher magnification images are shown in (’ and ’’) and white matter (WM), internal granule layer (IGL), PCL and molecular layer (ML) are delimited by white doted lines. (G-H) FIHC detection of the indicated proteins and dapi on sagittal sections of the vermis (lobule IV/V) of P4 Nes-CFP/+ mice. Scale bars indicate 50μm.

Supplementary Figure 3 Nes-CFP+ cells can self-renew and are multipotent in vitro.

(A) P4 Nes-CFP expression in SOX+ NSCs. (B) Flow cytometric analysis of P4 cerebellar Nes-CFP+ cells. Nes-CFP+ cells represent 6.19 % of the total population. (C-F) FACS-isolated CFP+ cells form neurospheres after 7 days, when cultured in neurosphere media containing bFGF and EGF. (C and D) Neurospheres expresses CFP (D) and Sox2 (D’). Bright-field view in (C). (E and F) Graphs showing the number of neurospheres formed when CF+ and CFP- cells are plated at the indicated numbers (60 cells/well: p<0.0001, t(4)=28; 600 cells/well: p=0.0031, t(4)=6.369; 6000 cells/well: p=0.0042, t(4)=5.884) (E) and indicated serial passages (secondary: p<0.0001, t(4)=18.30; tertiary: p=0.0408, t(4)=2.98; primary vs secondary: p=0.0155, t(4)=4.049; primary vs tertiary: p=0.2999, t(4)=1.19; secondary vs tertiary: p=09523, t(4)=0.06369) (F) (n=3 experiments). (G-J) Nes-CFP+ neurospheres were dissociated, plated as a monolayer and were grown in either neurosphere media containing bFGF and EGF or differentiation media with 10% serum for 7 days. The levels of CFP and SOX2 (G-G’), GFAP (H-H’), TUJ1(I-I’) and O4 (J-J’) are shown using immunofluorescent microscopy to assess multipotency. Nes-CFP+ cells showed morphological changes and were able to differentiate into neurons, astrocytes and oligodendrocytes when subjected to differentiation. Graphical data are present as means ± SEM and significance was determined with the two-tailed Student’s t test, ****P<0.0001, **P<0.01, *P<0.05. All scale bars indicate 100 μm.

Supplementary Figure 4 Tamoxifen administration delays the response of NEPs to injury.

(A-D) FIHC detection of the indicated proteins, EdU and dapi on sagittal sections at the indicated ages of the vermis (lobule IV/V) of Non-IR (A, B) and IR (C, D) Nes-CFP/+mice administrated with Tm at P0. (E-J) Graphs of the changes in density at the indicated stages of CFP+ cells per mm of PCL (E) (n=3; P4 no Tm: p=0.017, t(4)=3.917; P4 Tm: p=0.127, t(4)=1.924; P6 no Tm: p=0.027, t(4)=3.418; P6 Tm: p=0.068, t(4)=2.48; P4 Non-IR Tm vs no Tm: p=0.634, t(4)=0.514; P6 Non-IR Tm vs no Tm: p=0.033, t(4)=3.171; P4 IR Tm vs no Tm: p=0.02, t(4)=3.738; P6 IR Tm vs no Tm: p=0.044, t(4)=2.898) and per mm2 of IGL+WM (H) (n=3; P4 no Tm: p=0.737, t(4)=0.36; P4 Tm: p=0.219, t(4)=1.46; P6 no Tm: p=0.179, t(4)=1.63; P6 Tm: p=0.027, t(4)=3.42; P4 Non-IR Tm vs no Tm: p=0.596, t(4)=0.56; P6 Non-IR Tm vs no Tm: p=0.0044, t(4)=5.79; P4 IR Tm vs no Tm: p=0.428, t(4)=0.88; P6 IR Tm vs no Tm: p=0.01, t(4)=4.53), the proportion of proliferating (Ki67+) CFP+ cells in the PCL (F) (n=3; P4 no Tm: p=0.0011, t(4)=8.5; P4 Tm: p=0.764, t(4)=0.322; P6 no Tm: p=0.03, t(4)=3.25; P6 Tm: p=0.541, t(4)=0.67; P4 Non-IR Tm vs no Tm: p=0.589, t(4)=0.586; P6 Non-IR Tm vs no Tm: p=0.36, t(4)=1.03; P4 IR Tm vs no Tm: p=0.096, t(4)=2.16; P6 IR Tm vs no Tm: p=0.01, t(4)=4.57) and in the WM+IGL (I) (n=3; P4 no Tm: p=0.054, t(4)=2.7; P4 Tm: p=0.0042, t(4)=5.86; P6 no Tm: p=0.506, t(4)=0.73; P6 Tm: p=0.557, t(4)=0.64; P4 Non-IR Tm vs no Tm: p=0.097, t(4)=2.16; P6 Non-IR Tm vs no Tm: p=0.11, t(4)=2.05; P4 IR Tm vs no Tm: p=0.943, t(4)=0.075; P6 IR Tm vs no Tm: p=0.012, t(4)=4.4), the proliferation index in PCL (% [Ki67+ GFP+ EdU+] cells of all [GFP+Ki67+] cells) (G) (n=3; P4 no Tm: p=0.717, t(4)=0.389; P4 Tm: p=0.678, t(4)=0.678; P6 no Tm: p=0.457, t(4)=0.457; P6 Tm: p=0.36, t(4)=0.363; P4 Non-IR Tm vs no Tm: p=0.263, t(4)=1.3; P6 Non-IR Tm vs no Tm: p=0.108, t(4)=2.06; P4 IR Tm vs no Tm: p=0.759, t(4)=0.33; P6 IR Tm vs no Tm: p=0.348, t(4)=1.06) and in the WM+IGL (J) (n=3; P4 no Tm: p=0.127, t(4)=1.92; P4 Tm: p=0.0088, t(4)=4.78; P6 no Tm: p=0.18, t(4)=1.62; P6 Tm: p=0.109, t(4)=2.06; P4 Non-IR Tm vs no Tm: p=0.822, t(4)=0.24; P6 Non-IR Tm vs no Tm: p=0.535, t(4)=0.68; P4 IR Tm vs no Tm: p=0.0429, t(4)=2.93; P6 IR Tm vs no Tm: p=0.384, t(4)=0.98). All of the analyses were performed on 3 midline sections per brain and for lobule IV/V. All graphical data are present as means ± SEM and the significance was determined with using two-tailed Student’s t test, ***P<0.001, **P<0.01, *P<0.05. Scale bars indicate 100 μm.

Supplementary Figure 5 Completion of cerebellum development is delayed after injury.

(A-B) H&E staining of sagittal sections of the vermis of P16 Nes-FlpoER/+; R26FSF-TDtom/+ (Nes-TDTom) mice given Tm at P0 with or without irradiation. (‘ and “) are high magnifications of the yellow dotted boxed areas indicated in A and B. Yellow arrowheads indicate the thicker EGL in the IR condition than control. Scale bars indicate 500 μm.

Supplementary Figure 6 Diphtheria toxin-induced GCP cell death triggers the replenishment of the EGL by NEPs.

(A) Schematic representation of the experimental design. (B-C) FIHC detection of Diphtheria toxin receptor (DTR) and Ki67 proteins and dapi on sagittal sections of the vermis (lobule IV/V) (B) and the para-vermis (C) of P1 Atoh1-tTA/+;TRE-Cre/+;R26LSL-DTR/+ (Atoh-DTR) mice not injected with DT. Yellow dotted circles in C and C’ indicate the position of the cerebellar nuclei (CN). (D-E) FIHC detection of PAX6 protein, TUNEL and dapi on sagittal sections of the vermis (lobule IV/V) of P2 Atoh1-tTA/+;R26LSL-DTR/+ (Control) and Atoh-DTR mice (DT at P1). Note the increase of TUNEL staining in the EGL of Atoh-DTR mice compared to controls (no DT). (F-I) H&E and FIHC detection of the indicated proteins and dapi on sagittal sections of the vermis of P4 Control; Nes-CFP/+ (F-G) and Atoh-DTR; Nes-CFP/+ (H-I) mice (DT at P1). G and I are from lobule IV/V. Yellow dotted lines in G’ and I’ indicate the EGL. Red arrowhead in I’ indicates the presence of CFP+ cells in the EGL. (J-K) FIHC detection of the indicated proteins and dapi on sagittal sections of the vermis (fissure between lobule IV/V and VI) of P6 Control; Nes-FlpoER/+; R26FSF-TDTom/+ (Nes-TDTom) (J) and Atoh-DTR; Nes-FlpoER/+; R26FSF-TDTom/+ (K) mice (Tm at P0 and DT at P1). Yellow dotted lines in J’ and K’ indicate the EGL. Red arrowhead in K’ indicates the presence of TDTom+ cells in the EGL. Black and white scale bars indicate 500 and 100 μm, respectively.

Supplementary Figure 7 NEPs progressively extinguished Nestin expression in the EGL

(A-C) FIHC detection of CFP on sagittal sections of the vermis (lobule IV/V) of IR Nes-CFP/+ mice at the indicated ages. The EGL is delimitated by yellow doted lines. Red arrowheads indicate the progressive decrease of CFP staining in the EGL. Scale bar indicates 100μm.

Supplementary Figure 8 NEPs are highly motile after EGL depletion.

(A-H) Detection of native CFP fluorescence on sagittal slices of the vermis (lobule IV/V) of P6 Non-IR (A-D) and IR (E-H) Nes-CFP/+mice showing tracks of CFP+ cells in each layer during the 5h45min of imaging. The length of the tracks represents the distance travelled by the cells. The color code for the tracks is as indicated. Black scale bar indicates 100μm.

Supplementary Figure 9 Irradiation changes the gene expression profile of NEPs.

(A) Heat map showing differential gene expression between Non-IR and IR P5 NEPs after irradiation of the CB at P1. Color code is as indicated. (B) Gene Ontology categories of the genes significantly changed by ≥1.5 fold in P5 IR NEPs compared to Non-IR. Gene Ontology was generated using DAVID Bioinformatics Resources 6.7. (C-F) In situ hybridization of Gli1 on P5 (C, E) and P8 (D, F) midsagittal cerebellar sections (lobule IV/V). Yellow arrowheads indicate the PCL expression is higher after IR. Yellow asterisks indicate the PCL expression is similar in both Non-IR and IR CB at P8. (G) Q-rtPCR analysis of the indicated genes in Nes-CFP+ NEPs isolated from P4 (Gli1: p=0.0533, t(4)=2.7; Ptch1: p=0.0632, t(4)=2.551; Ptch2: p=0.0526, t(4)=2.727) and P6 (Gli1: p=0.1654, t(4)=1.695; Ptch1: p=0.1943, t(4)=1.6; Ptch2: p=0.1557, t(4)=1.746) Non-IR and IR cerebella (n=3 FACS experiments/genotype).

Supplementary Figure 10 SHH signaling regulates the expansion of NEPs and the fate of NEP-derived cells.

(A-F) FIHC detection of the indicated proteins and dapi on midsagittal sections (lobule VIII) of P8 control Nes-FlpoER/+; R26MASTR/+ (Nes-GFP, n=3, A and D) and experimental Nes-FlpoER/+; R26MASTR/+; Smolox/lox (Nes-Smo CKO, n=3, B and E) oe Nes-FlpoER/+; R26MASTR/SmoM2-YPF (Nes-SmoM2, n=6, C and F) mice (Tm at P0). Scale bars indicate 100βm. (G-J) Graphs showing changes in the number of GFP+ cells in each layer (as indicated in A) of P8 mice that also expressed SOX2 only (EGL: Nes-GFP vs Nes-Smo CKO: p=N/A; Nes-GFP vs Nes-SmoM2: p= 0.185, t(7)=1.47; Nes-Smo CKO vs Nes-SmoM2: p=0.185, t(7)=1.47; ML: Nes-GFP vs Nes-Smo CKO: p=0.238, t(4)=1.387; Nes-GFP vs Nes-SmoM2: p= 0.203, t(7)=1.403; Nes-Smo CKO vs Nes-SmoM2: p=0.051, t(7)=2.38; PCL: Nes-GFP vs Nes-Smo CKO: p=0.031, t(4)=3.25; Nes-GFP vs Nes-SmoM2: p=0.779, t(7)=0.29; Nes-Smo CKO vs Nes-SmoM2: p=0.053, t(5.24)=2.49; IGL: Nes-GFP vs Nes-Smo CKO: p=0.0036, t(4)=6.13; Nes-GFP vs Nes-SmoM2: p= 0.0017, t(7)=4.92; Nes-Smo CKO vs Nes-SmoM2: p=0.0294, t(7)=2.73; WM: Nes-GFP vs Nes-Smo CKO: p=0.353, t(4)=1.05; Nes-GFP vs Nes-SmoM2: p= 0.54, t(7)=0.64; Nes-Smo CKO vs Nes-SmoM2: p=0.158, t(7)=1.58) (G), PAX2 only (EGL: Nes-GFP vs Nes-Smo CKO: p=0.597, t(4)=0.57; Nes-GFP vs Nes-SmoM2: p= 0.051, t(2.13)=3.99; Nes-Smo CKO vs Nes-SmoM2: p=0.0001, t(7)=7.64; ML: Nes-GFP vs Nes-Smo CKO: p=0.142, t(4)=1.83; Nes-GFP vs Nes-SmoM2: p= 0.0155, t(7)=3.18; Nes-Smo CKO vs Nes-SmoM2: p=0.124, t(7)=1.75; PCL: Nes-GFP vs Nes-Smo CKO: p=0.525, t(4)=0.695; Nes-GFP vs Nes-SmoM2: p=0.966, t(7)=0.044; Nes-Smo CKO vs Nes-SmoM2: p=0.667, t(7)=0.44; IGL: Nes-GFP vs Nes-Smo CKO: p=0.0299, t(4)=3.3; Nes-GFP vs Nes-SmoM2: p= 0.86, t(7)=0.184; Nes-Smo CKO vs Nes-SmoM2: p=0.0127, t(5.46)=3.64; WM: Nes-GFP vs Nes-Smo CKO: p=0.0014, t(4)=7.88; Nes-GFP vs Nes-SmoM2: p= 0.982, t(7)=0.023; Nes-Smo CKO vs Nes-SmoM2: p=0.0034, t(7)=4.34) (H), S100β only (EGL: Nes-GFP vs Nes-Smo CKO: p=0.0963, t(4)=2.165; Nes-GFP vs Nes-SmoM2: p= 0.389, t(7)=0.918; Nes-Smo CKO vs Nes-SmoM2: p=0.901, t(7)=0.129; ML: Nes-GFP vs Nes-Smo CKO: p=0.686, t(4)=0.435; Nes-GFP vs Nes-SmoM2: p=0.406, t(7)=0.884; Nes-Smo CKO vs Nes-SmoM2: p=0.243, t(7)=1.27; PCL: Nes-GFP vs Nes-Smo CKO: p=0.51, t(4)=0.722; Nes-GFP vs Nes-SmoM2: p=0.018, t(7)=3.07; Nes-Smo CKO vs Nes-SmoM2: p=0.032, t(7)=2.68; IGL: Nes-GFP vs Nes-Smo CKO: p=0.379, t(4)=0.99; Nes-GFP vs Nes-SmoM2: p= 0.99, t(7)=0.012; Nes-Smo CKO vs Nes-SmoM2: p=0.571, t(7)=0.594; WM: Nes-GFP vs Nes-Smo CKO: p=0.01, t(3.04)=4.524; Nes-GFP vs Nes-SmoM2: p= 0.725, t(7)=0.366; Nes-Smo CKO vs Nes-SmoM2: p=0.345, t(5.426)=1.458) (I) and both SOX2 and S100β (EGL: Nes-GFP vs Nes-Smo CKO: p>0.9999, t(4)=0; Nes-GFP vs Nes-SmoM2: p= 0.078, t(7)=2.06; Nes-Smo CKO vs Nes-SmoM2: p=0.078, t(7)=2.06; ML: Nes-GFP vs Nes-Smo CKO: p=0.493, t(4)=0.754; Nes-GFP vs Nes-SmoM2: p= 0.082, t(7)=2.028; Nes-Smo CKO vs Nes-SmoM2: p=0.06, t(7)=2.238; PCL: Nes-GFP vs Nes-Smo CKO: p=0.018, t(3.475)=3.87; Nes-GFP vs Nes-SmoM2: p=0.0039, t(7)=4.23; Nes-Smo CKO vs Nes-SmoM2: p=0.0005, t(7)=5.99; IGL: Nes-GFP vs Nes-Smo CKO: p=0.064, t(4)=2.546; Nes-GFP vs Nes-SmoM2: p= 0.61, t(7)=0.532; Nes-Smo CKO vs Nes-SmoM2: p=0.0029, t(7)=4.483; WM: Nes-GFP vs Nes-Smo CKO: p=0.086, t(4)=2.269; Nes-GFP vs Nes-SmoM2: p= 0.174, t(7)=1.514; Nes-Smo CKO vs Nes-SmoM2: p=0.01, t(7)=3.42) (J). All of the analyses were performed on 3 midline sections per brain. All graphical data are presented as means ± SEM and the significance was determined with two-tailed Student’s t test, ***P<0.001, **P<0.01, *P<0.05. Scale bars indicate 100 μm. (K-P) H&E and FIHC detection of the indicated proteins and dapi on sagittal sections of the vermis of P8 (A-C) and P12 (D-F) Nes-FlpoER/+; R26MASTR/SmoM2-YPF (Nes-SmoM2) mice given Tm at P0. (B-C and E-F) are high magnifications of the yellow dotted boxed areas indicated in A and D. Yellow dotted lines indicate the EGL (E) and molecular layer (ML). Black and white scale bars indicate 500 and 50 μm, respectively.

Supplementary Figure 11 Loss of Smo specifically in NEPs results in a lack of expansion and migrate to the EGL after injury.

(A-F) FIHC detection of GFP on cerebellar midsagittal sections at P12 (Lobule IV/V) of control Nes-GFP Non-IR (solid black) and IR (dashed black) and mutant Nes-Smo CKO Non-IR (solid red) and IR (dashed red) mice given Tm at P0. Yellow dotted lines outline the different layers. Scale bars indicate 100μm.

Supplementary Figure 12 Model of cellular responses during regeneration of the developing cerebellum.

Depletion of the External Granule Layer (EGL) during the first days of postnatal cerebellar development results in up-regulation of Hedgehog (HH)-signaling in the Purkinje Cell Layer (PCL), which leads to the expansion and migration of Nestin-Expressing Progenitors (NEPs) in the PCL that normally produce Astrocytes (Astro) and Bergmann Glia (Bg). Once in the EGL, NEP-derived cells progressively lose their Neural Stem Cell (NSC) markers (SOX2 and Nestin), they initiate expression of Granule Cell (GC) lineage-specific genes (Pax6 and Atoh1) and expand to replenish the EGL. Concomitantly, down-regulation of HH-signaling in White Matter (WM) NEPs induces a transient reduction of production of interneurons and astrocytes by the WM bipotent progenitors. Thus, injury of the EGL stimulates a cell-cell communication system that coordinates the responses of the different NEP populations during recovery, leading to a reset of the postnatal developmental clock of the cerebellum to re-establish the correct proportions of cerebellar cell types and ensure normal cerebellar circuit formation.

Supplementary information

Supplementary Text and Figures

Supplementary Figures 1–12 (PDF 4975 kb)

Life Sciences Reporting Summary (PDF 142 kb)

Supplementary Table 1

Statistical details for figure legends (XLSX 17 kb)

Supplementary Table 2

Differentially expressed genes in NEPs from P5 Non-IR et IR cerebella. (XLSX 58 kb)

P6 Non-IR cerebellum shows no obvious movement of NEPs to the EGL.

Detection of native CFP fluorescence on sagittal slices of the vermis (lobule 4/5) of P6 Non-IR Nes-CFP/+ mice showing displacement of CFP+ cells. Image stacks were acquired every 3min for 5h 45min. (MOV 15867 kb)

NEPs are highly motile after irradiation.

Detection of native CFP fluorescence on sagittal slices of the vermis (lobule 4/5) of P6 IR Nes-CFP/+ mice showing movement of CFP+ cells within the EGL and between the Purkinje Cell Layer (PCL) and EGL. Image stacks were acquired every 3min for 5h 45min. (MOV 21908 kb)

PCL NEPs migrate towards the EGL after irradiation.

Magnification of native CFP fluorescence detection on sagittal slices of the vermis (lobule 4/5) of P6 IR Nes-CFP/+ mice showing movement of CFP+ cells into the EGL. Image stacks were acquired every 3min for 5h 45min. Green and purple dots indicate cells located initially in the Purkinje Cell Layer (PCL) and Molecular Layer (ML) respectively. (AVI 10374 kb)

Adult Non-IR Nes-Smo CKO mice have normal motor behavior.

Video showing motor behavior of 5-6 week old Nes-FlpoER/+;R26FSF-GFPcre/+;Smoflox/flox (Nes-Smo CKO) Non-IR mouse administered with Tm at P0. (MP4 18828 kb)

Adult IR Nes-Smo CKO mice show impaired motor behavior.

Video showing motor behavior of 5-6 week old Nes-FlpoER/+;R26FSF-GFPcre/+;Smoflox/flox (Nes-Smo CKO) mouse administered with Tm at P0 and irradiated at P1. (MP4 17017 kb)

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Wojcinski, A., Lawton, A., Bayin, N. et al. Cerebellar granule cell replenishment postinjury by adaptive reprogramming of Nestin+ progenitors. Nat Neurosci 20, 1361–1370 (2017). https://doi.org/10.1038/nn.4621

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/nn.4621

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