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:

Transcriptional repression of Bmp2 by p21Waf1/Cip1 links quiescence to neural stem cell maintenance

Abstract

Relative quiescence and self renewal are defining features of adult stem cells, but their potential coordination remains unclear. Subependymal neural stem cells (NSCs) lacking cyclin-dependent kinase (CDK) inhibitor (CKI) 1a (p21) exhibit rapid expansion that is followed by their permanent loss later in life. Here we demonstrate that transcription of the gene encoding bone morphogenetic protein 2 (Bmp2) in NSCs is under the direct negative control of p21 through actions that are independent of CDK. Loss of p21 in NSCs results in increased levels of secreted BMP2, which induce premature terminal differentiation of multipotent NSCs into mature non-neurogenic astrocytes in an autocrine and/or paracrine manner. We also show that the cell-nonautonomous p21-null phenotype is modulated by the Noggin-rich environment of the subependymal niche. The dual function that we describe here provides a physiological example of combined cell-autonomous and cell-nonautonomous functions of p21 with implications in self renewal, linking the relative quiescence of adult stem cells to their longevity and potentiality.

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: The CKI p21 is expressed in stem cell–like astrocytes and promotes multipotency.
Figure 2: BMP2 production is increased in the absence of p21.
Figure 3: The CKI p21 prevents B cell terminal differentiation.
Figure 4: Enhanced BMP secretion is responsible for the deficit in self renewal of p21-null neurospheres.
Figure 5: p21 acts as a Bmp2 transcriptional repressor.

Similar content being viewed by others

References

  1. Zhao, C., Deng, W. & Gage, F.H. Mechanisms and functional implications of adult neurogenesis. Cell 132, 645–660 (2008).

    Article  CAS  Google Scholar 

  2. Ferri, A.L. et al. Sox2 deficiency causes neurodegeneration and impaired neurogenesis in the adult mouse brain. Development 131, 3805–3819 (2004).

    Article  CAS  Google Scholar 

  3. Maslov, A.Y., Barone, T.A., Plunkett, R.J. & Pruitt, S.C. Neural stem cell detection, characterization, and age-related changes in the subventricular zone of mice. J. Neurosci. 24, 1726–1733 (2004).

    Article  CAS  Google Scholar 

  4. Nam, H.S. & Benezra, R. High levels of Id1 expression define B1 type adult neural stem cells. Cell Stem Cell 5, 515–526 (2009).

    Article  CAS  Google Scholar 

  5. Doetsch, F. et al. Lack of the cell-cycle inhibitor p27Kip1 results in selective increase of transit-amplifying cells for adult neurogenesis. J. Neurosci. 22, 2255–2264 (2002).

    Article  CAS  Google Scholar 

  6. Kippin, T.E., Martens, D.J. & van der Kooy, D. p21 loss compromises the relative quiescence of forebrain stem cell proliferation leading to exhaustion of their proliferation capacity. Genes Dev. 19, 756–767 (2005).

    Article  CAS  Google Scholar 

  7. Besson, A., Dowdy, S.F. & Roberts, J.M. CDK inhibitors: cell cycle regulators and beyond. Dev. Cell 14, 159–169 (2008).

    Article  CAS  Google Scholar 

  8. Cheng, T. et al. Hematopoietic stem cell quiescence maintained by p21cip1/waf1. Science 287, 1804–1808 (2000).

    Article  CAS  Google Scholar 

  9. Yu, H., Yuan, Y., Shen, H. & Cheng, T. Hematopoietic stem cell exhaustion impacted by p18 INK4C and p21 Cip1/Waf1 in opposite manners. Blood 107, 1200–1206 (2006).

    Article  CAS  Google Scholar 

  10. Yuan, Y., Shen, H., Franklin, D.S., Scadden, D.T. & Cheng, T. In vivo self-renewing divisions of haematopoietic stem cells are increased in the absence of the early G1-phase inhibitor, p18INK4C. Nat. Cell Biol. 6, 436–442 (2004).

    Article  CAS  Google Scholar 

  11. Zezula, J. et al. p21cip1 is required for the differentiation of oligodendrocytes independently of cell cycle withdrawal. EMBO Rep. 2, 27–34 (2001).

    Article  CAS  Google Scholar 

  12. Chang, B.D. et al. Effects of p21Waf1/Cip1/Sdi1 on cellular gene expression: implications for carcinogenesis, senescence, and age-related diseases. Proc. Natl. Acad. Sci. USA 97, 4291–4296 (2000).

    Article  CAS  Google Scholar 

  13. Coqueret, O. & Gascan, H. Functional interaction of STAT3 transcription factor with the cell cycle inhibitor p21WAF1/CIP1/SDI1. J. Biol. Chem. 275, 18794–18800 (2000).

    Article  CAS  Google Scholar 

  14. Delavaine, L. & La Thangue, N.B. Control of E2F activity by p21Waf1/Cip1. Oncogene 18, 5381–5392 (1999).

    Article  CAS  Google Scholar 

  15. Devgan, V., Mammucari, C., Millar, S.E., Brisken, C. & Dotto, G.P. p21WAF1/Cip1 is a negative transcriptional regulator of Wnt4 expression downstream of Notch1 activation. Genes Dev. 19, 1485–1495 (2005).

    Article  CAS  Google Scholar 

  16. Snowden, A.W., Anderson, L.A., Webster, G.A. & Perkins, N.D. A novel transcriptional repression domain mediates p21WAF1/CIP1 induction of p300 transactivation. Mol. Cell Biol. 20, 2676–2686 (2000).

    Article  CAS  Google Scholar 

  17. Marqués-Torrejón, M.A. et al. Cyclin-dependent kinase inhibitor p21 controls adult neural stem cell expansion by regulating Sox2 gene expression. Cell Stem Cell 12, 88–100 (2013).

    Article  Google Scholar 

  18. Brugarolas, J. et al. Radiation-induced cell cycle arrest compromised by p21 deficiency. Nature 377, 552–557 (1995).

    Article  CAS  Google Scholar 

  19. Raponi, E. et al. S100B expression defines a state in which GFAP-expressing cells lose their neural stem cell potential and acquire a more mature developmental stage. Glia 55, 165–177 (2007).

    Article  Google Scholar 

  20. Encinas, J.M. et al. Division-coupled astrocytic differentiation and age-related depletion of neural stem cells in the adult hippocampus. Cell Stem Cell 8, 566–579 (2011).

    Article  CAS  Google Scholar 

  21. Bond, A.M., Bhalala, O.G. & Kessler, J.A. The dynamic role of bone morphogenetic proteins in neural stem cell fate and maturation. Dev. Neurobiol. 72, 1068–1084 (2012).

    Article  CAS  Google Scholar 

  22. Gomes, W.A., Mehler, M.F. & Kessler, J.A. Transgenic overexpression of BMP4 increases astroglial and decreases oligodendroglial lineage commitment. Dev. Biol. 255, 164–177 (2003).

    Article  CAS  Google Scholar 

  23. Gross, R.E. et al. Bone morphogenetic proteins promote astroglial lineage commitment by mammalian subventricular zone progenitor cells. Neuron 17, 595–606 (1996).

    Article  CAS  Google Scholar 

  24. Snyder, E.Y. et al. Multipotent neural cell lines can engraft and participate in development of mouse cerebellum. Cell 68, 33–51 (1992).

    Article  CAS  Google Scholar 

  25. Wilson, P.A. & Hemmati-Brivanlou, A. Vertebrate neural induction: inducers, inhibitors, and a new synthesis. Neuron 18, 699–710 (1997).

    Article  CAS  Google Scholar 

  26. Korchynskyi, O. & ten Dijke, P. Identification and functional characterization of distinct critically important bone morphogenetic protein-specific response elements in the Id1 promoter. J. Biol. Chem. 277, 4883–4891 (2002).

    Article  CAS  Google Scholar 

  27. Le Dréau, G. et al. Canonical BMP7 activity is required for the generation of discrete neuronal populations in the dorsal spinal cord. Development 139, 259–268 (2012).

    Article  Google Scholar 

  28. Ferrón, S.R. et al. A combined ex/in vivo assay to detect effects of exogenously added factors in neural stem cells. Nat. Protoc. 2, 849–859 (2007).

    Article  Google Scholar 

  29. Ghosh-Choudhury, N. et al. Autoregulation of mouse BMP-2 gene transcription is directed by the proximal promoter element. Biochem. Biophys. Res. Commun. 286, 101–108 (2001).

    Article  CAS  Google Scholar 

  30. Müller, H. et al. E2Fs regulate the expression of genes involved in differentiation, development, proliferation, and apoptosis. Genes Dev. 15, 267–285 (2001).

    Article  Google Scholar 

  31. Lin, J., Reichner, C., Wu, X. & Levine, A.J. Analysis of wild-type and mutant p21WAF-1 gene activities. Mol. Cell Biol. 16, 1786–1793 (1996).

    Article  CAS  Google Scholar 

  32. Nakanishi, M., Robetorye, R.S., Pereira-Smith, O.M. & Smith, J.R. The C-terminal region of p21SDI1/WAF1/CIP1 is involved in proliferating cell nuclear antigen binding but does not appear to be required for growth inhibition. J. Biol. Chem. 270, 17060–17063 (1995).

    Article  CAS  Google Scholar 

  33. Ma, Y. et al. A small-molecule E2F inhibitor blocks growth in a melanoma culture model. Cancer Res. 68, 6292–6299 (2008).

    Article  CAS  Google Scholar 

  34. Mabie, P.C. et al. Bone morphogenetic proteins induce astroglial differentiation of oligodendroglial-astroglial progenitor cells. J. Neurosci. 17, 4112–4120 (1997).

    Article  CAS  Google Scholar 

  35. Lim, D.A. et al. Noggin antagonizes BMP signaling to create a niche for adult neurogenesis. Neuron 28, 713–726 (2000).

    Article  CAS  Google Scholar 

  36. Colak, D. et al. Adult neurogenesis requires Smad4-mediated bone morphogenic protein signaling in stem cells. J. Neurosci. 28, 434–446 (2008).

    Article  CAS  Google Scholar 

  37. Bonaguidi, M.A. et al. LIF and BMP signaling generate separate and discrete types of GFAP-expressing cells. Development 132, 5503–5514 (2005).

    Article  CAS  Google Scholar 

  38. Gobeske, K.T. et al. BMP signaling mediates effects of exercise on hippocampal neurogenesis and cognition in mice. PLoS ONE 4, e7506 (2009).

    Article  Google Scholar 

  39. Mira, H. et al. Signaling through BMPR-IA regulates quiescence and long-term activity of neural stem cells in the adult hippocampus. Cell Stem Cell 7, 78–89 (2010).

    Article  CAS  Google Scholar 

  40. Ferrándiz, N. et al. p21 as a transcriptional co-repressor of S-phase and mitotic control genes. PLoS ONE 7, e37759 (2012).

    Article  Google Scholar 

  41. Cooper-Kuhn, C.M. et al. Impaired adult neurogenesis in mice lacking the transcription factor E2F1. Mol. Cell Neurosci. 21, 312–323 (2002).

    Article  CAS  Google Scholar 

  42. Tevosian, S.G., Paulson, K.E., Bronson, R. & Yee, A.S. Expression of the E2F–1/DP-1 transcription factor in murine development. Cell Growth Differ. 7, 43–52 (1996).

    CAS  PubMed  Google Scholar 

  43. Vidal, A., Millard, S.S., Miller, J.P. & Koff, A. Rho activity can alter the translation of p27 mRNA and is important for RasV12-induced transformation in a manner dependent on p27 status. J. Biol. Chem. 277, 16433–16440 (2002).

    Article  CAS  Google Scholar 

Download references

Acknowledgements

We thank C. Gallego (Institut de Biologia Molecular de Barcelona), M.R. Campanero (Instituto de Investigaciones Biomédicas Alberto Sols, Madrid), S. Harris (University of Texas Health Science Center at San Antonio), S. Hitoshi (Shiga University of Medical Science), D. van der Kooy (University of Toronto) and W. Zhang (The Johns Hopkins University School of Medicine) for generously providing constructs. We also thank the Servicios Centrales de Soporte a la Investigación Experimental of the Universidad de Valencia for technical support. This work was supported by grants from the Ministerio de Economía y Competitividad (MINECO) (SAF program) to I.F., S.R.F. and A.V. and the Ministerio de Sanidad (Red TerCel and CIBERNED) and Generalitat Valenciana (Programas Prometeo, ISIC and ACOMP) to I.F. M.A.M.-T. was a recipient of a predoctoral fellowship from the Ministerio de Educación y Ciencia (FPI program), C.A-A. was a recipient of a predoctoral fellowship from the Ministerio de Educación y Ciencia (FPU program), E.G.-I. is supported by a predoctoral fellowship from Xunta de Galicia, and C.C. was a 'Parga Pondal' investigator supported by Xunta de Galicia.

Author information

Authors and Affiliations

Authors

Contributions

E.P. performed blots and ChIP in neurospheres and contributed to transcription assays. J.M.M.-R. performed transcription assays and contributed to the expression analyses. E.P. and J.M.M.-R. performed ELISA, intracerebro-ventricular infusion of Noggin and rescue experiments in vitro. M.A.M.-T. contributed to the analysis of the mouse phenotype in vivo and in vitro and to the multipotency analyses. C.A.-A. contributed to the coculture experiments and expression analysis. A.V. designed the generation and validation of p21 mutant constructs and the interference of p21 with shRNA on c17.2 cells. C.C., E.G.-I., A.S. and A.V. generated and validated p21 mutant constructs and the shRNA to p21 in c17.2 and 3T3 cells and performed ChIP in c17.2 cells. S.R.F. performed the analysis of the phenotype in vivo and in vitro, performed terminal astrocyte differentiation experiments and contributed to expression analyses and rescue experiments. I.F. designed and supervised the study, secured funding and analyzed the data. All authors contributed to experimental design, discussions and data analysis. E.P., J.M.M.-R., A.V., S.R.F. and I.F. wrote the manuscript.

Corresponding authors

Correspondence to Sacri R Ferrón or Isabel Fariñas.

Ethics declarations

Competing interests

The authors declare no competing financial interests.

Integrated supplementary information

Supplementary Figure 1 Primary neurosphere-forming cells in p21-null mice.

(a) Phase contrast micrographs of primary neurospheres formed by cells directly obtained from 2-m Cdkn1a wild-type and mutant mice. (b) Graph representing the numbers of primary neurospheres obtained from individual SEZs of wild-type and Cdkn1a mutant mice at different adult ages, from 1-m to 12-m. Notice that between 1-m and 5-m of age more neurospheres are recovered from mutant tissue but that, after 5-m, the mutant yield decreases dramatically (2-m: n = 5; 4-m: n = 4; 5-m: n = 3; 6-m: n = 6; 12-m: n = 4 animals per genotype). (c) Senescence-associated (SA) β-galactosidase staining in early and late passages of wild-type and Cdkn1a-null cultures. (d) Self-renewal assays in early and late-passage secondary neurospheres derived from wild-type and Cdkn1a-null animals. F/DFn/ Dfd: 1.833/7/8; 4.770/6/6. P=0.4143, 0.0789. (e) γH2AX and 53BP1 simultaneous immunofluorescent detection in early and late-passage secondary neurospheres cells derived from wild-type and Cdkn1a-null animals. (f) Percentage of Sox2-high expressing cells and (g) percentage of γH2AX-53BP1 double-positive cells in early and late passages of wild-type and Cdkn1a-null neurosphere-derived cells. F/DFn/Dfd: 3.829/2/1; P=0.0229 (f); F/DFn/Dfd: 5.051/2/1; P=0.6003 (g). Data are shown as mean values ± s.e.m.; *p<0.05. Scale bar in a, c: 50 μm; in e: 25 μm; insert in f: 10 μm.

Supplementary Figure 2 Analysis of B-cell derivatives in the SEZ of p21-null 2-month old mice.

(a) Immunohistochemistry for Mash1 (red) and Ki67 (green) in coronal sections through the SEZ of 2-m wild-type and Cdkn1a mutant mice, as seen by confocal microscopy. White arrows indicate double-positive cells. (b) Immunohistochemistry for βIII-tubulin (green) and Ki67 (red) in the SEZ of 2-m Cdkn1a wild-type and mutant mice. White arrows indicate double-positive cells. (c) Percentage of Mash1+ and βIII-tubulin+ cells and (d) Mash1-Ki67 and βIIItubulin-Ki67 double positive cells found in the SEZ of young animals of the two genotypes. Data are shown as mean values ± s.e.m. (n = 3); *p<0.05, **p<0.01. Scale bar in a: 10 μm; in b: 10 μm.

Supplementary Figure 3 Enhanced astroglial differentiation in p21-null mice.

(a) Coronal sections of the SEZ of 12-m wild-type and Cdkn1a mutant mice stained for S100β (blue) and p-SMAD1 (red). Notice increased numbers of double positive cells, indicated by white arrows, in the mutant sample. (b) Percentage of pSMAD1+ and pSMAD1+S100β+ cells in the SEZ of 12-m wild-type and Cdkn1a mutant mice. Data are shown as mean values ± s.e.m. (n = 3); *p<0.05. Scale bars: in a, 20 μm.

Supplementary Figure 4 BMP2-induced astroglial differentiation in p21 null mice is independent of Sox2-induced replicative stress.

(a) Percentage of GFAP+ cells expressing high levels of Sox2 (Sox2high) amongst GFAP-Sox2 double positive cells, in saline (−) vs. Noggin (+) infused p21-null mice. F/DFn/Dfd: 17.30/2/2; P=0.1093. (b) Percentage of S100β/γH2AX double positive cells in wild-type and saline and noggin-infused p21-null animals. F/DFn/Dfd: 2.046/4/2; 12.63/2/1 P=0.7085; 0.3902. (c) Examples of the simultaneous detection of S100β and γH2AX in the SEZ of wild-type and saline and Noggin-infused p21-null animals. (d) Percentage of cells expressing high levels of Sox2 and (e) percentages of γH2AX and 53BP1 doubly positive cells in early and late-passage, wild-type and p21-null Noggin treated cells. (f) Fold change in neurosphere numbers produced by FACS-sorted cells from neurospheres infected with a retroviral construct for the expression of Sox2 (pMY-sox2) or empty vector (pMY), treated or not with 100 ng/ml Noggin. Data are shown as mean values ± s.e.m; *p<0.05. Scale bar in c: 35 μm.

Supplementary Figure 5 Bmp2 proximal promoter is preferentially used in neurospheres.

(a) Schematic representation of the murine Bmp2 promoter (grey bar) showing two alternative transcription initiation sites (TSS) at +1 and −736, and the two fragments used for reporter experiments: distal, (−2712/−700)-luciferase, red bar, and proximal, (−643/+165)-luciferase, purple bar. Neurospheres were transfected with either of these reporter constructs and luciferase activity was measured 24 h later. (b) Interfered shRNAp21-c17.2 cells were transfected with an E2F-luciferase reporter bearing 3 tandem repeats of the hydropholate reductase E2F-binding site (E2F-luc) and treated with 2.5 μM CDK inhibitor 2-bromo-12,13-dihydro-5H-indolo [2,3-a] pyrrolo [3,4-c] carbazole-5,7(6H)-dione. Interfered shRNAp21-c17.2 cells were transfected in parallel with the proximal Bmp2 promoter in the presence or in the absence of the inhibitor. (c) p21-null neurospheres were transfected with the E2F-luciferase reporter (E2F-luc), or co-transfected with pCDNA-p21 and the proximal Bmp2-promoter, and cells were treated or not with 40 μM E2F inhibitor HLM006474. Luciferase activity was measured 24h after transfection. (d) Immunoblot for Flag epitope showing the expression of the p21-mutants used (upper panel) and co-immunoprecipitation of p21-Flag and CDK2 or Cyclin A in mouse fibroblasts transfected with the different p21-mutants used (bottom panels). (e) Percentage of cells transfected with the different p21-mutant constructs that are retained in G1 phase of the cell cycle shown as increase relative to empty vector values. Data are shown as mean values ± s.e.m. (n = 4, panels a–c; n =2, panel e); **p<0.05,**p<0.01, **p<0.001.

Supplementary Figure 6 Uncropped protein immunoblots and DNA agarose gels.

(a) Immunoblots in Fig. 2d. (b) Immunoblot in Fig.4d. (c) ChIP agarose gels in Fig. 5d. (d) Immunoblots in Fig. S5d. Dashed squares outline the cropped area presented in the corresponding figure. MWM: molecular weight marker, IB: immunoblot, IP: immunoprecipitation, e.v.: empty vector, NTC: non template control.

Supplementary information

Supplementary Text and Figures

Supplementary Figures 1–6 (PDF 7396 kb)

Rights and permissions

Reprints and permissions

About this article

Cite this article

Porlan, E., Morante-Redolat, J., Marqués-Torrejón, M. et al. Transcriptional repression of Bmp2 by p21Waf1/Cip1 links quiescence to neural stem cell maintenance. Nat Neurosci 16, 1567–1575 (2013). https://doi.org/10.1038/nn.3545

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/nn.3545

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