Skip to main content

OPINION article

Front. Physiol., 17 October 2013
Sec. Craniofacial Biology and Dental Research
This article is part of the Research Topic Stem Cell Programming and Reprogramming View all 14 articles

SoxB, cell cycle and neurogenesis

  • 1Stem Cell Biology Laboratory, Biomedical Sciences Research Centre “Alexander Fleming”, Vari-Attica, Greece
  • 2The Cyprus Institute of Neurology and Genetics, Nicosia, Cyprus

Sox genes encode for a group of transcription factors that share the presence of an HMG-type (high mobility group) box. This is a 79 amino acid motif that encodes their DNA binding domain. They were identified soon after the discovery of the mammalian testis-determining gene Sry, which also possesses an HMG box, and obtained their name from Sry (“Sox” Sry-related box genes). The members of the Sox family are highly conserved among vertebrates and so far 26 different genes have been identified (Schepers et al., 2002; Wilson and Koopman, 2002). Sox proteins activate or repress target gene transcription, via interactions with a partner factor in a cell-type specific manner (Kondoh and Kamachi, 2010). Based on sequence homology, within and outside the HGM box, they have been classified into seven groups (SoxA to SoxG groups) (Wilson and Koopman, 2002).

SoxB genes, divided in SoxB1 (Sox1-3) and SoxB2 (Sox14 and 21) subgroups, code for a group of highly related transcription factors expressed mainly in the developing and adult Central Nervous System (CNS) in mammals (Wood and Episkopou, 1999; Schepers et al., 2002; Avilion et al., 2003; Sarkar and Hochedlinger, 2013). During embryonic development, the nervous system arises from primitive neuroepithelial cells that give rise to neural stem cells (NSCs), cells that have the potential to self-renew and differentiate into neurons, astrocytes and oligodendrocytes. Neuroepithelial progenitors (NEP) represent the cells with a more restricted potential, while precursors describe cells that exist in an earlier developmental state than others (McKay, 1997). In the developing CNS, Sox1-3 and Sox21 are expressed both in embryonic and adult NEPs and in subsets of differentiated neurons of the adult brain and spinal cord, whereas Sox14 is expressed only in selected differentiated neurons of the embryonic and adult CNS. Proteins encoded by Sox14 and Sox21 genes are very similar to each other but distinct from those encoded by Sox1-3, in areas outside the HMG domain and Group B homology that lies immediately C-proximal of the HMG domain (Uchikawa et al., 1999). Sox1-3 proteins are thought to function as transcriptional activators, while Sox14/21 are believed to function as repressors. However, Sox1 and Sox2 have been shown to be involved in repression of transcription as well (Cavallaro et al., 2008; Elkouris et al., 2011). Despite the wide distribution in the expression of SoxB proteins in the developing CNS, the effort to unravel their distinct function has been hampered by redundancy issues, eventually leading to functional compensation, but insights into the function of each factor have been gained from studies in areas where unique expression of a single factor is detected.

SoxB function has been studied extensively during neurogenesis, the process of neuron formation, which involves continuous cycles of proliferation and differentiation of NEPs. During CNS development, dividing NEPs, residing in the ventricular zone (VZ), either self-renew or exit the cell cycle and differentiate. NEPs make cell fate decisions controlled by extrinsic and/or intrinsic determinants (Ramasamy et al., 2013). Neurons are generated in two sequential steps. In the specification step, which takes place within the VZ, some NEPs up-regulate the Notch receptors and remain as uncommitted progenitors while others down-regulate Notch receptors, up-regulate Notch ligands and then the proneural bHLH genes, Neurogenins 1-3 (Ngn1-3)/Mash1 and progress to terminal differentiation (Guillemot, 2007; Martynoga et al., 2012; Alberi et al., 2013). Accordingly, loss of Notch signaling invariably leads to over-production of neurons and depletion of NEPs (Pierfelice et al., 2011). In the commitment step, which is a transitionary phase and takes place on the lateral margins of the VZ known as intermediate zone (IZ), specified neuronal progenitors switch on another class of bHLH factors like NeuroD, Prox1, and Nscl1, which consolidate the specification step and allow the cells to progress to terminal differentiation (Christie et al., 2013). Ectopic expression of any of the latter can by-pass the specification phase and drive NEPs out of the cell cycle.

The transcription factors Sox1-3 have been proposed to block neuronal commitment and thus prevent differentiation (Bylund et al., 2003; Holmberg et al., 2008; Oosterveen et al., 2013). In contrast Sox21 opposes Sox1-3 function in NEPs, where they are co-expressed, and promotes neurogenesis in vivo (Sandberg et al., 2005). Given that SoxB1 and Sox21 factors are co-expressed in the NEPs and have been shown to bear antagonizing activities, there was an open question how a cell decides to remain uncommitted or exit the cell cycle and how the antagonizing function of these factors is modulated. One model would propose that the relative activities of Sox1-3 vs. Sox21 are thought to determine whether a cell will remain undifferentiated or commit to the neuronal lineage. Accordingly, loss of Sox21 could lead to a predominance of SoxB1 activity leading to a block on neurogenesis in the embryo. However, this is not the case during development, since loss of Sox21 leads to a block in the progression of neuronal specification in the hippocampus only in adult mice (Matsuda et al., 2012). For Sox1, the data had also been conflicting. In vitro studies suggested that Sox1, like Sox21, could promote neurogenesis while Sox2 or Sox3 can block it. Initial reports suggested that over-expression of Sox1, but not Sox2 or Sox3, could promote neurogenesis when over-expressed in primary cultures of neural stem/progenitor cells cultured as neurospheres and the carcinoma cell line P19 (Pevny et al., 1998; Kan et al., 2004, 2007). However, this model was not supported by in vivo loss-of-function analysis in mice. In contrast, the role of Sox1 has been best studied during neuronal differentiation in the developing ventral forebrain and spinal cord, where a unique role has been assigned during neuronal subtype specification (Ekonomou et al., 2005; Panayi et al., 2010). We believe that light into the role of SoxB genes during neuronal specification is likely to be shed by the results from double mutant experiments. For instance, despite the fact that Sox1 and Sox21 single mutant mice are born, double Sox1/Sox21 mutants are embryonic lethal. The cause of lethality is yet to be determined (SM, unpublished data). Neuronal fates are determined around the final cell cycle of NEPs during which neural differentiation factors regulate the cell cycle either directly or indirectly. Only recently, SoxB genes have been implicated in the regulation of cell cycle. Studies in the mouse developing cortex showed that Sox1 acts to maintain the undifferentiated state of NEPs via a mechanism involving suppression of Prox1–mediated cell cycle exit and neurogenesis (Elkouris et al., 2011). These results are consistent with previous data from gain-of-function and loss-of-function analysis in the chick spinal cord, according to which SoxB1 factors block neurogenesis downstream of proneural factors (Bylund et al., 2003; Holmberg et al., 2008).

Sox2 has found itself in the position of the most celebrated of the Sox proteins due to its role in epiblast formation (Avilion et al., 2003; Mandalos et al., 2012), pluripotency (Masui et al., 2007) and reprogramming (Polo et al., 2013; Sarkar and Hochedlinger, 2013). Sox2 null embryos die soon after implantation, masking a potential role for Sox2 in the CNS (Avilion et al., 2003; Mandalos et al., 2012). However, the function of Sox2 during neurogenesis has only recently begun to be elucidated using conditional mutagenesis in mice (Episkopou, 2005). Sox2 is expressed initially in NEPs and later in NSCs in the neurogenic niches of the embryonic and adult CNS (Wood and Episkopou, 1999; Avilion et al., 2003; Favaro et al., 2009; Remboutsika et al., 2011; Kang and Hebert, 2012; Mandalos et al., 2012; Raitano et al., 2013; Ramasamy et al., 2013). Sox2 becomes down-regulated during the final cell cycle of NEPs, immediately before differentiation and is required for the maintenance of NEPs properties functioning partly through the Shh and wnt3a pathways during embryonic and early postnatal life (Favaro et al., 2009). NSCs are completely lost from the early postnatal hippocampus upon loss of Sox2 but they gradually re-appear, an observation attributed to compensatory functions of Sox1 and Sox3 proteins, which are both expressed in an overlapping manner with Sox2. At this point, it is worth to mention that Sox2 maintains the embryonic cortical NSCs cycling, while at the same time it preserves and regulates their Pax6+ cortical radial identity and plasticity (Remboutsika et al., 2011; Raitano et al., 2013). Nevertheless, the exact role of Sox2 during cell cycle regulation of NSCs and NEPs is yet elusive. Recent evidence suggests that the function of SoxB1 factors, with even overlapping activities, is not strictly redundant, as they induce different sets of genes and are likely to partner with different proteins to maintain progenitor identity (Archer et al., 2011). It is reasonable then to suggest that during neurogenesis post-translational modifications, context-specific selection of partner proteins or competition for the same target binding sites modulate the activity of SoxB transcription factors. Thus, despite the discovery of SoxB transcription factors for over fifteen years their role during cell cycle, neurogenesis, neuronal commitment and specification has just started to be elucidated.

Dedication

This review is dedicated to the memory of Prof. Larysa Pevny.

References

Alberi, L., Hoey, S. E., Brai, E., Scotti, A. L., and Marathe, S. (2013). Notch signaling in the brain: in good and bad times. Ageing Res. Rev. 12, 801–814. doi: 10.1016/j.arr.2013.03.004

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Archer, T. C., Jin, J., and Casey, E. S. (2011). Interaction of Sox1, Sox2, Sox3 and Oct4 during primary neurogenesis. Dev. Biol. 350, 429–440. doi: 10.1016/j.ydbio.2010.12.013

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Avilion, A. A., Nicolis, S. K., Pevny, L. H., Perez, L., Vivian, N., and Lovell-Badge, R. (2003). Multipotent cell lineages in early mouse development depend on SOX2 function. Genes Dev. 17, 126–140. doi: 10.1101/gad.224503

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Bylund, M., Andersson, E., Novitch, B. G., and Muhr, J. (2003). Vertebrate neurogenesis is counteracted by Sox1-3 activity. Nat. Neurosci. 6, 1162–1168. doi: 10.1038/nn1131

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Cavallaro, M., Mariani, J., Lancini, C., Latorre, E., Caccia, R., Gullo, F., et al. (2008). Impaired generation of mature neurons by neural stem cells from hypomorphic Sox2 mutants. Development 135, 541–557. doi: 10.1242/dev.010801

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Christie, K. J., Emery, B., Denham, M., Bujalka, H., Cate, H. S., and Turnley, A. M. (2013). Transcriptional regulation and specification of neural stem cells. Adv. Exp. Med. Biol. 786, 129–155. doi: 10.1007/978-94-007-6621-1_8

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Ekonomou, A., Kazanis, I., Malas, S., Wood, H., Alifragis, P., Denaxa, M., et al. (2005). Neuronal migration and ventral subtype identity in the telencephalon depend on SOX1. PLoS Biol. 3:e186. doi: 10.1371/journal.pbio.0030186

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Elkouris, M., Balaskas, N., Poulou, M., Politis, P. K., Panayiotou, E., Malas, S., et al. (2011). Sox1 maintains the undifferentiated state of cortical neural progenitor cells via the suppression of Prox1-mediated cell cycle exit and neurogenesis. Stem Cells 29, 89–98. doi: 10.1002/stem.554

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Episkopou, V. (2005). SOX2 functions in adult neural stem cells. Trends Neurosci. 28, 219–221. doi: 10.1016/j.tins.2005.03.003

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Favaro, R., Valotta, M., Ferri, A. L., Latorre, E., Mariani, J., Giachino, C., et al. (2009). Hippocampal development and neural stem cell maintenance require Sox2-dependent regulation of Shh. Nat. Neurosci. 12, 1248–1256. doi: 10.1038/nn.2397

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Guillemot, F. (2007). Cell fate specification in the mammalian telencephalon. Prog. Neurobiol. 83, 37–52. doi: 10.1016/j.pneurobio.2007.02.009

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Holmberg, J., Hansson, E., Malewicz, M., Sandberg, M., Perlmann, T., Lendahl, U., et al. (2008). SoxB1 transcription factors and Notch signaling use distinct mechanisms to regulate proneural gene function and neural progenitor differentiation. Development 135, 1843–1851. doi: 10.1242/dev.020180

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Kan, L., Israsena, N., Zhang, Z., Hu, M., Zhao, L. R., Jalali, A., et al. (2004). Sox1 acts through multiple independent pathways to promote neurogenesis. Dev. Biol. 269, 580–594. doi: 10.1016/j.ydbio.2004.02.005

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Kan, L., Jalali, A., Zhao, L. R., Zhou, X., McGuire, T., Kazanis, I., et al. (2007). Dual function of Sox1 in telencephalic progenitor cells. Dev. Biol. 310, 85–98. doi: 10.1016/j.ydbio.2007.07.026

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Kang, W., and Hebert, J. M. (2012). A Sox2 BAC transgenic approach for targeting adult neural stem cells. PLoS ONE 7:e49038. doi: 10.1371/journal.pone.0049038

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Kondoh, H., and Kamachi, Y. (2010). SOX-partner code for cell specification: regulatory target selection and underlying molecular mechanisms. Int. J. Biochem. Cell Biol. 42, 391–399. doi: 10.1016/j.biocel.2009.09.003

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Mandalos, N., Saridaki, M., Harper, J. L., Kotsoni, A., Yang, P., Economides, A. N., et al. (2012). Application of a novel strategy of engineering conditional alleles to a single exon gene, Sox2. PLoS ONE 7:e45768. doi: 10.1371/journal.pone.0045768

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Martynoga, B., Drechsel, D., and Guillemot, F. (2012). Molecular control of neurogenesis: a view from the mammalian cerebral cortex. Cold Spring Harb. Perspect. Biol. 4, 10. doi: 10.1101/cshperspect.a008359

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Masui, S., Nakatake, Y., Toyooka, Y., Shimosato, D., Yagi, R., Takahashi, K., et al. (2007). Pluripotency governed by Sox2 via regulation of Oct3/4 expression in mouse embryonic stem cells. Nat. Cell Biol. 9, 625–635. doi: 10.1038/ncb1589

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Matsuda, S., Kuwako, K., Okano, H. J., Tsutsumi, S., Aburatani, H., Saga, Y., et al. (2012). Sox21 promotes hippocampal adult neurogenesis via the transcriptional repression of the Hes5 gene. J. Neurosci. 32, 12543–12557. doi: 10.1523/JNEUROSCI.5803-11.2012

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

McKay, R. (1997). Stem cells in the central nervous system. Science 276, 66–71. doi: 10.1126/science.276.5309.66

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Oosterveen, T., Kurdija, S., Enstero, M., Uhde, C. W., Bergsland, M., Sandberg, M., et al. (2013). SoxB1-driven transcriptional network underlies neural-specific interpretation of morphogen signals. Proc. Natl. Acad. Sci. U.S.A. 110, 7330–7335. doi: 10.1073/pnas.1220010110

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Panayi, H., Panayiotou, E., Orford, M., Genethliou, N., Mean, R., Lapathitis, G., et al. (2010). Sox1 is required for the specification of a novel p2-derived interneuron subtype in the mouse ventral spinal cord. J. Neurosci. 30, 12274–12280. doi: 10.1523/JNEUROSCI.2402-10.2010

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Pevny, L. H., Sockanathan, S., Placzek, M., and Lovell-Badge, R. (1998). A role for SOX1 in neural determination. Development 125, 1967–1978.

Pubmed Abstract | Pubmed Full Text

Pierfelice, T., Alberi, L., and Gaiano, N. (2011). Notch in the vertebrate nervous system: an old dog with new tricks. Neuron 69, 840–855. doi: 10.1016/j.neuron.2011.02.031

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Polo, J. M., Anderssen, E., Walsh, R. M., Schwarz, B. A., Nefzger, C. M., Lim, S. M., et al. (2013). A molecular roadmap of reprogramming somatic cells into iPS cells. Cell 151, 1617–1632. doi: 10.1016/j.cell.2012.11.039

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Raitano, S., Verfaillie, C. M., and Petryk, A. (2013). Self-renewal of neural stem cells: implications for future therapies. Front. Physiol. 4:49. doi: 10.3389/fphys.2013.00049

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Ramasamy, S., Narayanan, G., Sankaran, S., Yu, Y. H., and Ahmed, S. (2013). Neural stem cell survival factors. Arch. Biochem. Biophys. 534, 71–87. doi: 10.1016/j.abb.2013.02.004

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Remboutsika, E., Elkouris, M., Iulianella, A., Andoniadou, C. L., Poulou, M., Mitsiadis, T. A., et al. (2011). Flexibility of neural stem cells. Front. Physiol. 2:16. doi: 10.3389/fphys.2011.00016

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Sandberg, M., Kallstrom, M., and Muhr, J. (2005). Sox21 promotes the progression of vertebrate neurogenesis. Nat. Neurosci. 8, 995–1001. doi: 10.1038/nn1493

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Sarkar, A., and Hochedlinger, K. (2013). The sox family of transcription factors: versatile regulators of stem and progenitor cell fate. Cell Stem Cell 12, 15–30. doi: 10.1016/j.stem.2012.12.007

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Schepers, G. E., Teasdale, R. D., and Koopman, P. (2002). Twenty pairs of sox: extent, homology, and nomenclature of the mouse and human sox transcription factor gene families. Dev. Cell 3, 167–170. doi: 10.1016/S1534-5807(02)00223-X

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Uchikawa, M., Kamachi, Y., and Kondoh, H. (1999). Two distinct subgroups of Group B Sox genes for transcriptional activators and repressors: their expression during embryonic organogenesis of the chicken. Mech. Dev. 84, 103–120. doi: 10.1016/S0925-4773(99)00083-0

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Wilson, M., and Koopman, P. (2002). Matching SOX: partner proteins and co-factors of the SOX family of transcriptional regulators. Curr. Opin. Genet. Dev. 12, 441–446. doi: 10.1016/S0959-437X(02)00323-4

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Wood, H. B., and Episkopou, V. (1999). Comparative expression of the mouse Sox1, Sox2 and Sox3 genes from pre-gastrulation to early somite stages. Mech. Dev. 86, 197–201. doi: 10.1016/S0925-4773(99)00116-1

Pubmed Abstract | Pubmed Full Text | CrossRef Full Text

Keywords: Sox, stem cells, neural stem cells, neural progenitors, neurons, specification, commitment

Citation: Karnavas T, Mandalos N, Malas S and Remboutsika E (2013) SoxB, cell cycle and neurogenesis. Front. Physiol. 4:298. doi: 10.3389/fphys.2013.00298

Received: 09 July 2013; Accepted: 29 September 2013;
Published online: 17 October 2013.

Edited by:

Sachiko Iseki, Tokyo Medical and Dental University, Japan

Copyright © 2013 Karnavas, Mandalos, Malas and Remboutsika. This is an open-access article distributed under the terms of the Creative Commons Attribution License (CC BY). The use, distribution or reproduction in other forums is permitted, provided the original author(s) or licensor are credited and that the original publication in this journal is cited, in accordance with accepted academic practice. No use, distribution or reproduction is permitted which does not comply with these terms.

*Correspondence: remboutsika@gmail.com

These authors have contributed equally to this work.

Disclaimer: All claims expressed in this article are solely those of the authors and do not necessarily represent those of their affiliated organizations, or those of the publisher, the editors and the reviewers. Any product that may be evaluated in this article or claim that may be made by its manufacturer is not guaranteed or endorsed by the publisher.