Next Article in Journal
The Emerging Roles of Long Non-Coding RNAs in Intellectual Disability and Related Neurodevelopmental Disorders
Next Article in Special Issue
PAX8 in the Junction between Development and Tumorigenesis
Previous Article in Journal
APTC-C-SA01: A Novel Bacteriophage Cocktail Targeting Staphylococcus aureus and MRSA Biofilms
Previous Article in Special Issue
PAX9 in Cancer Development
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Thirty Years’ History since the Discovery of Pax6: From Central Nervous System Development to Neurodevelopmental Disorders

Department of Developmental Neuroscience, Tohoku University Graduate School of Medicine, Sendai 980-8575, Japan
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2022, 23(11), 6115; https://doi.org/10.3390/ijms23116115
Submission received: 3 May 2022 / Revised: 19 May 2022 / Accepted: 27 May 2022 / Published: 30 May 2022
(This article belongs to the Special Issue PAX Genes in Health and Diseases)

Abstract

:
Pax6 is a sequence-specific DNA binding transcription factor that positively and negatively regulates transcription and is expressed in multiple cell types in the developing and adult central nervous system (CNS). As indicated by the morphological and functional abnormalities in spontaneous Pax6 mutant rodents, Pax6 plays pivotal roles in various biological processes in the CNS. At the initial stage of CNS development, Pax6 is responsible for brain patterning along the anteroposterior and dorsoventral axes of the telencephalon. Regarding the anteroposterior axis, Pax6 is expressed inversely to Emx2 and Coup-TF1, and Pax6 mutant mice exhibit a rostral shift, resulting in an alteration of the size of certain cortical areas. Pax6 and its downstream genes play important roles in balancing the proliferation and differentiation of neural stem cells. The Pax6 gene was originally identified in mice and humans 30 years ago via genetic analyses of the eye phenotypes. The human PAX6 gene was discovered in patients who suffer from WAGR syndrome (i.e., Wilms tumor, aniridia, genital ridge defects, mental retardation). Mutations of the human PAX6 gene have also been reported to be associated with autism spectrum disorder (ASD) and intellectual disability. Rodents that lack the Pax6 gene exhibit diverse neural phenotypes, which might lead to a better understanding of human pathology and neurodevelopmental disorders. This review describes the expression and function of Pax6 during brain development, and their implications for neuropathology.

1. Introduction

It has been 30 years since Pax6/PAX6 was identified as a gene responsible for congenital anomalies of the eye in mice and humans [1,2]. The gene turned out to be a member of a paired box (Pax) family encoding transcriptional factors that also work in brain development [3]. Regarding structure, Pax proteins share a common DNA-binding domain called the paired domain (PD) [4]. Some members of the Pax family, including Pax6, have a homeodomain (HD), i.e., another DNA-binding domain (Figure 1A) [5,6]. The molecular structure and function of Pax6 are well preserved from Drosophila to mammals [7,8]. It has been reported that the PD is necessary for the regulation of embryonic neurogenesis, in which a mutant lacking the HD of Pax6 showed only subtle defects [9]. Therefore, the PD exerts a key role during brain development [9].
Pax6 first garnered attention due to its role in eye development. Spontaneous Small eye (Sey) mutant mice were identified from the phenotype of microphthalmia as a heterozygous phenotype, while Pax6 homozygous mutant (Sey/Sey) mice completely lacked the formation of the eyes and nose and died soon after birth [10]. The Pax6 gene mutated in Sey/Sey contains a stop codon in the coding region before the homeodomain (Figure 1A) [1]. The Pax6 gene was also identified in spontaneous mutant mice (Sey/Sey) and rats (rSey2/rSey2) showing similar ocular phenotypes [11,12,13]. The gene responsible for the eyeless mutant in Drosophila was found to be a Pax6 homolog [14]. The ectopic expression of eyeless induced the formation of compound eyes outside the head region [7]. Thus, Pax6 is required and sufficient for eye formation.
Further study has revealed that Pax6 is expressed in various regions in the developing central nervous system (CNS) from the initial stage when the neural plate is induced [15]. In the developing CNS, Pax6 is expressed in the telencephalon [16], diencephalon [17,18], rhombomeres [19,20], and spinal cord [20,21,22]. At early stages, Pax6 is expressed in the ventricular zone (VZ), where neural stem cells (NSCs) are located (see “Pax6 in neurogenesis”) [23,24]. At later stages, Pax6 is expressed in the neurons of specific brain regions such as the olfactory bulb [20,25], amygdala [13,26], thalamus [13,27], and cerebellum [13,20,28]. rSey2/rSey2 rats lack the olfactory bulb, yet they have an olfactory bulb-like structure at the lateral position in the neocortex [20]. Sey/Sey mice and rSey2/rSey2 rats exhibit a reduction in the size of the forebrain and the cortical thickness (Figure 1B,C) [29].
Outside the CNS, PAX6/Pax6 is expressed in the lens [20,30,31], corneal epithelium [32,33], retinal neuroepithelium [34,35], and olfactory placodes/epithelium [25,36,37]. Pax6 is also involved in the development of other tissues such as the pancreas, pituitary gland, and even in the testes [21,38,39,40] (reviewed in [13,20,37,41,42,43,44]). In this review article, we would like to focus on the role of Pax6 in brain development, taking neurodevelopmental disorders into account.
Figure 1. Schematic structure of Pax6 protein and cortical phenotypes in the Pax6 spontaneous homozygous mutant (Sey/Sey) and Pax6 splice variant (Pax6(5a)) mice. (A) Schematic structure of the Pax6 protein in the wild-type (WT), Sey/Sey, and Pax6(5a) mice. One base-pair substitution in exon 8 of the Pax6 gene causes a Gly194Stop nonsense mutation, causing a stop codon downstream at the PST site, thereby resulting in the truncated Pax6Sey protein. The canonical PD binds via its N-terminal PAI domain to the DNA; the insertion of the 14 amino acids from exon 5a into the PAI domain leads to the mutant PD(5a) [45]. PD: paired domain, HD: homeodomain, PST: Pro-Ser-Thr rich region. (B) Schematic illustration of the head of WT and Sey/Sey mice. Pax6 shows a gradient of expression in the AP axis in WT. Sey/Sey mice cause a failure of eye and olfactory bulb formation and rostralization in the telencephalon. OB: olfactory bulb, Tel: telencephalon, M1: motor, S1: sensory, A1: auditory and V1: visual areas. (C) Schematic illustration of a WT and Sey/Sey mouse telencephalon. Sey/Sey mice cause dorsalization in the telencephalon. Cx: cortex, BG: basal ganglia. The graphical diagram has been redrawn from [9,20,46].
Figure 1. Schematic structure of Pax6 protein and cortical phenotypes in the Pax6 spontaneous homozygous mutant (Sey/Sey) and Pax6 splice variant (Pax6(5a)) mice. (A) Schematic structure of the Pax6 protein in the wild-type (WT), Sey/Sey, and Pax6(5a) mice. One base-pair substitution in exon 8 of the Pax6 gene causes a Gly194Stop nonsense mutation, causing a stop codon downstream at the PST site, thereby resulting in the truncated Pax6Sey protein. The canonical PD binds via its N-terminal PAI domain to the DNA; the insertion of the 14 amino acids from exon 5a into the PAI domain leads to the mutant PD(5a) [45]. PD: paired domain, HD: homeodomain, PST: Pro-Ser-Thr rich region. (B) Schematic illustration of the head of WT and Sey/Sey mice. Pax6 shows a gradient of expression in the AP axis in WT. Sey/Sey mice cause a failure of eye and olfactory bulb formation and rostralization in the telencephalon. OB: olfactory bulb, Tel: telencephalon, M1: motor, S1: sensory, A1: auditory and V1: visual areas. (C) Schematic illustration of a WT and Sey/Sey mouse telencephalon. Sey/Sey mice cause dorsalization in the telencephalon. Cx: cortex, BG: basal ganglia. The graphical diagram has been redrawn from [9,20,46].
Ijms 23 06115 g001

2. Pax6 in Cortical Patterning

As mentioned above, Pax6 is considered to be expressed from the initial developmental stage of the CNS, when it plays key roles in the regionalization of the neuroectoderm and neural tube, i.e., the primordium of the CNS [47]. Each region of the CNS is specialized before and during neurogenesis [48]. Although Pax6 is involved in the patterning of the ventral regions of the brain stem and spinal cord [20,49,50], we focus here on the cortical patterning. As described later, the initial patterning can also influence the ontology of neurodevelopmental disorders.

2.1. Anteroposterior Axis

The telencephalon is a highly regionalized organ, which is functionally and morphologically diverse. In early brain development, the telencephalon is patterned along the anteroposterior (AP) and dorsoventral (DV) axes, resulting in the formation of various cortical and subcortical areas [51].
This patterning is first attributed to the action of secreted signaling factors such as Wnt, bone morphogenetic protein (BMP), and FGF8, and secondarily to the transcription factors such as Pax6, Emx2, Coup-TFI, and Sp8 [52,53,54,55,56]. It is of note that these key transcription factors are expressed in a gradient (Figure 2A). Pax6 and Sp8 are expressed in a rostro-lateralhigh to caudo-mediallow manner along the AP axis. Emx2 and Coup-TF1 are expressed inversely to Pax6 and Sp8 [55,57,58,59]. Changes in the expression of these transcription factors affect brain patterning. For example, the overall cortical regions in Sey/Sey mice shift to rostral, resulting in a reduction in the motor (M1) and sensory (S1) cortical areas, and conversely an enlargement of the auditory (A1) and visual (V1) areas (Figure 2B) [57,60]. In contrast, the cortical areas in Emx2 mutant mice shift to caudal, resulting in enlarged M1 and S1 areas and reduced A1 and V1 areas [57,60,61]. Yac-Pax6 Tg mice, in which Pax6 is overexpressed, show a slight decrease in the S1 area but no major changes in other brain areas (Figure 2C) [62,63]. These results suggest that the proper amount of Pax6 regulates the arealization of the cortex. Furthermore, Pax6 is regulated by Emx2 and Coup-TF1, which determine the caudal area of the cortex (Figure 2B–D) (see [51] and the references therein), indicating that regulatory networks including Pax6 and other transcriptional factors contribute to the AP patterning of the telencephalon.
The expression of the transcription factors is induced by specific secreted molecules (Figure 2D). For example, BMP and Wnt, being released from the dorso-medial telencephalon, increase the expression of Emx2 [52], while Fgf15 from the rostral side induces that of Coup-TF1 [64]. In contrast, Fgf8 from the rostral side induces Sp8 [55,59]. A mathematical analysis of gene regulatory networks using Boolean and computational models has revealed an underpinning expression gradient in the mammalian cortex [65,66]. Although upstream secretary molecules that can induce the expression of Pax6 have not been identified yet, it is of interest that Pax6 controls the responsiveness to secreted molecules such as sonic hedgehog and BMP in the cortex [67]. Future analyses may reveal more about the relationship between Pax6 and secreted molecules.
Figure 2. Gradient expression of key transcription factors in cortical patterning. (A) Gradient expression patterns of the major transcription factors Pax6, Emx2, Coup-TFI, Sp8, Fgf8, and Emx1 along the anteroposterior (AP) and lateromedial (LM); anterolateral (AL) and posteromedial (PM) axes. Pax6 is highly expressed along the AL axis, and weakly along the PM axis. Emx2 is expressed inversely to Pax6. (B,C) Summary of loss-of-function (B) and gain-of-function (C) by the alteration of transcription factor expression in regard to cortical patterning. Analyses of brain formation indicate that Sey/Sey mice show a reduction in the M1 and S1 areas, as well as enlargement of the A1 and V1 areas. Embryonic analyses imply that Emx2 mutants show an inverse relationship to Sey/Sey mice. The overexpression of Emx2 under the control of a Nestin promoter increases the size of the V1 area. In contrast, the overexpression of Pax6 slightly reduces the size of the S1 area. (D) The regulatory network of transcription factors and secreted molecules. The BMP and Wnt gradients increase the expression of Emx2. Fgf15 enhances the expression of Coup-TF1. Fgf8 inhibits the expression of Emx2 and Coup-TF1. TF: transcriptional factor. The graphical diagram has been redrawn from [51,54,55,56,60,61,63,66,68,69,70].
Figure 2. Gradient expression of key transcription factors in cortical patterning. (A) Gradient expression patterns of the major transcription factors Pax6, Emx2, Coup-TFI, Sp8, Fgf8, and Emx1 along the anteroposterior (AP) and lateromedial (LM); anterolateral (AL) and posteromedial (PM) axes. Pax6 is highly expressed along the AL axis, and weakly along the PM axis. Emx2 is expressed inversely to Pax6. (B,C) Summary of loss-of-function (B) and gain-of-function (C) by the alteration of transcription factor expression in regard to cortical patterning. Analyses of brain formation indicate that Sey/Sey mice show a reduction in the M1 and S1 areas, as well as enlargement of the A1 and V1 areas. Embryonic analyses imply that Emx2 mutants show an inverse relationship to Sey/Sey mice. The overexpression of Emx2 under the control of a Nestin promoter increases the size of the V1 area. In contrast, the overexpression of Pax6 slightly reduces the size of the S1 area. (D) The regulatory network of transcription factors and secreted molecules. The BMP and Wnt gradients increase the expression of Emx2. Fgf15 enhances the expression of Coup-TF1. Fgf8 inhibits the expression of Emx2 and Coup-TF1. TF: transcriptional factor. The graphical diagram has been redrawn from [51,54,55,56,60,61,63,66,68,69,70].
Ijms 23 06115 g002

2.2. Dorsoventral Axis

Dorsoventrally, the telencephalon is divided into two compartments: the dorsal cortex and the ventral basal ganglia, each of which has specific molecular features [71]. For proper patterning along the DV axis, the coordination of multiple transcription factors is required. In this section, we will review that Pax6 also influences DV patterning in the telencephalon.
The homeobox genes Pax6 and Gsh2 are expressed in the cortex and lateral ganglionic eminence (LGE), respectively. It is of note that these two transcriptional factors play complementary roles in the DV patterning of the mammalian telencephalon [72,73]. In Pax6-deficient mice, the region expressing Ascl1 (Mash1) and Dlx, which is originally localized in the basal ganglia, shifts dorsally, while that expressing Neurogenin1/2 (Neurog1/2), which is originally localized in the dorsal cortex, shifts dorsomedially [74]. In Gsh2-deficient mice, conversely, the region expressing Ascl1/Dlx and Neurog1/2 shifts ventrally (Figure 1C) [74,75]. It has been indicated that Gsh2 is required for repressing Pax6 and vice versa, and Pax6 is required for maintaining Neurog1/2, which means, in turn, Pax6 is responsible for the repression of Ascl1/Dlx [74,76]. These results indicate that Pax6 defines the DV boundary region and contributes to DV progenitor identity acquisition.
In addition to the dorsalization of the embryonic telencephalon, our group has found another key transcription factor under the control of Pax6. From microarray analyses using rSey2/rSey2 rat embryos, we have identified Dmrta1 (doublesex and mab-3-related transcription factor-like family A1) as a Pax6 downstream target gene in the rat telencephalon [77]. Dmrta1 is specifically expressed in the dorsal telencephalon and contributes to DV patterning [77]. Dmrta1 overexpression in the rat ventral telencephalon induced the mis-expression of the dorsal marker Neurog2 and repressed that of the ventral marker Ascl1 [77]. These novel pathways, i.e., Pax6 → Dmrta1 → Ngn2-|Ascl1 and/or Pax6 → Dmrta1-|Ascl1, could determine progenitor cell DV identity by repressing ventralization. For more information on the roles of Dmrt family members in brain patterning, see the review in [78]. As described in this section, Pax6 promotes brain development via numerous downstream factors to maintain proper AP and DV patterning in the telencephalon.

3. Pax6 in Neurogenesis

During embryonic brain development, neuroepithelial cells initially expand their population by symmetric cell division (Figure 3A, the proliferation phase) [79,80]. Later, the neuroepithelial cells become thinner and longer and are called radial glial (RG) cells. The RG cells divide symmetrically or asymmetrically to self-renew themselves or to produce neurons, respectively (Figure 3A, the neurogenetic phase). After neural differentiation, RG cells produce glial cells, i.e., first astrocytes and then oligodendrocytes, in the cortical primordium (Figure 3A, the gliogenic phase). The neuroepithelial cells and RG cells function as NSCs.
The relative proportion of Pax6-positive VZ in the telencephalon is largest at around embryonic day (E) 12.5 [81], at the time of the transition between the proliferation of NSCs and the induction of neurons (Figure 3B,C) [20,71,82]. Pax6-positive NSCs sequentially differentiate into neurons via neural progenitor cells, so-called intermediate progenitors (IPCs) (see below) [83].
Neurogenesis, i.e., production of neurons, occurs simultaneously with cortical patterning, and these two phenomena are difficult to separate. It is currently believed that neurogenesis continues throughout life in certain brain regions such as the hippocampus dentate gyrus and the subventricular zone of the lateral ventricle [84,85,86,87]. Pax6 is expressed in NSCs at all stages of embryonic, postnatal, and adult neurogenesis (see “Pax6 in neurogenesis” and “Pax6 in Relation to Neurodevelopmental Disorders”) (Figure 3B) [88,89,90]. In addition, Pax6 is expressed in astrocytes [91,92], a type of glial cell sharing some features with NSCs, maintaining a good balance between the maintenance and differentiation of astrocyte progenitors [91]. The determination of the proper neuronal arrangement in the telencephalon proceeds by the following two steps; (i) cell proliferation, and (ii) neural differentiation. In this section, the two different steps regulated by the downstream genes of Pax6 will be reviewed.

3.1. Downstream Genes of Pax6 Contribute to Self-Renewal of NSCs

As previously mentioned in the introduction of this paper, NSCs proliferate before neural differentiation proceeds (Figure 3A). The cell cycle consists of a series of intracellular processes, i.e., promoting cell division and producing two daughter cells. Cyclin D2 is essential for the transition from the G1 phase (when the cellular and extracellular environment is checked) to the S phase (DNA synthesis phase). The daughter cells expressing cyclin D2 are maintained in an undifferentiated state with proliferative potential, whereas those that lack cyclin D2 exit the cell cycle and give rise to neurons [93]. Cyclin D2 might be negatively regulated by Pax6, since cyclin D2 expression is increased in the Sey/Sey telencephalon at E12.5 [94] and in the cortex at later stages (our unpublished data). These cyclin D2-expressing cells might promote cell cycle re-entry because of an increased S phase population in the Sey/Sey telencephalon at E15.5 [95]. Interestingly, the S phase length in Sey/Sey mice is reported to be shorter at E12.5 but longer at E15.5 [95]. Cyclin D2 is also expressed in NSCs in the ganglionic eminence, and these cyclin D2-expressing cells gradually migrate toward the dorsal cortex. Based on the information above, the effects of cyclin D2 in the Sey/Sey cortex should be carefully examined regarding the embryonic stages and radial versus tangential migration of neurons.
One unique phenomenon that occurs during cell proliferation in the cortex is called interkinetic nuclear migration (INM), originally named as “elevator movement” by Fujita et al. [96,97], in which the nucleus of the neuroepithelial/RG cell goes up and down, from the apical to the basal side within the VZ according to the cell cycle [79,98]. rSey2/Sey2 rat cortexes shows ectopic INM, suggesting that Pax6 is involved in its regulation [20]. We also found that Ninein, which is a centrosome protein downstream of Pax6, regulates the dynamics of INM by anchoring to microtubules (Figure 4A–C) [99]. Pax6 and its downstream gene Ninein are involved in the regulation of INM, thereby serving to ensure the proliferation of NSCs.
In addition to the cell cycle-related regulation described above, chromatin structures are also key for maintaining NSCs in an undifferentiated state [100,101,102,103]. Pax6 is reported to bind to chromatin remodeling complexes including the RE1 silencing transcription factor (REST) and BRG1/BRM-associated factor (BAF) [104,105,106]. From chromatin immunoprecipitation (ChIP) analyses, it has been shown that the Baf170 subunit, a direct Pax6-interacting protein, recruits the REST-corepressor complex, thereby regulating the binding efficiency of target genes that induce IPCs and early-born neurons [105]. Since REST is expressed in NSCs in the cortical primordia (Figure 4A) [107,108], the Pax6/REST complex is responsible for the regulation of maintenance of NSCs and neural differentiation at the appropriate time.
Another interesting gene downstream of Pax6 is Fmr1, a causative gene for fragile X syndrome, one of the neurodevelopmental disorders. A previous ChIP-chip analysis suggested that Pax6 binds to the promoter of the Fmr1 gene [109]. Fmr1 encodes an RNA-binding protein, FMRP, which maintains NSCs [110]. A knockdown of Fmr1 and Fmr1 knock-out mice resulted in a reduction in the size of the Pax6+ NSC pool due to an NSC-to-IPC cell fate change [110]. A novel finding is that FMRP localized in the basal endfeet of NSCs transports its target mRNAs, leading to the regulation of localized mRNAs (Figure 4A) [111]. Moreover, FMRP-target mRNAs during corticogenesis include various genes related to neurodevelopmental disorders [112]; some of them could regulate the proper maintenance of NSCs and might explain mental retardation, one of the symptoms of fragile X syndrome. Since Pax6 is considered as an ASD risk gene (see “Pax6 in neurodevelopmental disorders” and [42]), the Pax6–Fmr1 regulation in corticogenesis may provide new insights into the pathogenesis of ASD.

3.2. Downstream Genes of Pax6 Contribute to Neural Differentiation

The transition from cell proliferation to neural differentiation of NSCs is governed by the expression of pro-neural genes [80,113,114,115]. Pax6 binds to the enhancer region of a pro-neural gene, Neurog2, and directly induces its expression, which in turn downregulates expression of Pax6 itself (Figure 4B) [116,117]. NSCs differentiate into IPCs with the up-regulated expression of Tbr2 (EOMES), a T-domain transcription factor that is induced by Pax6 [83]. Then, the induction of Tbr2 promotes the differentiation of IPs into postmitotic projection neurons via up-regulation of Tbr1 expression [83]. These facts indicate that a sequential molecular cascade of “Pax6 → Tbr2 → Tbr1” correlates with the cell type change of “RGs → IPs → postmitotic projection neurons” according to differentiation [83]. Another group has reported that Tbr2 is a downstream gene of Neurog2 [118]. Pax6 thus regulates multiple downstream genes, i.e., Pax6 → Tbr2 → Tbr1, Pax6 → Neurog2 → Tbr2 → Tbr1, and Pax6 → Neurog2 → Insm1 → Neurod1 → Tbr1 (Figure 4D) [119,120]. Although further fate mapping analyses at the single cell level are needed, the molecular machinery underlying Pax6 downstream cascades may regulate cell fate determination in the proper timing of neural development. It is of note that the transition of transcription factor expression is also recapitulated in adult neurogenesis (Figure 5).
Neurogenesis proceeds through positive and negative regulation of multiple transcription factors. One major signal that regulates many downstream transcriptomic factors is Notch signaling, which is a widely conserved signaling pathway from Drosophila to mammals [121]. Notch signaling is well known to determine cell fate choice, such as cell proliferation versus differentiation, during the formation of various tissues including the CNS [122,123,124,125,126]. Although Pax6 is conserved in both the chick and mouse telencephalons, the mechanism of neural differentiation provided by Pax6 differs among species and developmental stages [127]. In chicks and in the early stage of the mouse cortex, Pax6 transiently suppresses Notch signaling and induces neural differentiation (Figure 4D), while in the mid/late stages of the mouse cortex, Pax6 maintains NSCs. Therefore, it is speculated that a spatiotemporal dual function of Pax6 in the mouse cortex could lead to the generation of the sophisticated mammalian brain architecture.
PAX6 is also expressed in human neuroectoderm (NE) tissue [128]. Interestingly, PAX6 is uniformly expressed in the NE cells of human fetuses and those of differentiated cells [128]. However, Pax6 is expressed in restricted mouse brain regions during later development [128]. The human neocortex is enlarged compared to that of the mouse, possibly due to the function of basal RG cells [129]. Pax6 is highly expressed in primate but not mouse basal progenitors. Wong et al. demonstrated that sustained Pax6 expression specifically in BP-producing apical RG cells induces primate-like progenitor cells, indicating that sustained PAX6 expression in basal progenitors could be a key feature of subventricular enlargement in the human brain [130].
PAX6 dysfunction is also involved in the pathogenesis of glioblastoma. Transcriptomic and epigenomic analyses have revealed that PAX6/DLX6 promotes the differentiation of WNT5A-mediated glioblastoma stem cells into endothelial-like cells, which serve as an environmental niche supporting the growth of invasive glioma cells throughout the brain parenchyma [131]. In a controversial study, PAX6-knock out human neuronal epithelioma cells display increased proliferation and colony-forming abilities, indicating that PAX6 functions as a tumor suppressor [132]. Actually, there are case reports showing downregulation of PAX6 in glioma samples (see reviews [133] and the references therein). In this way, the formation/prevention of brain tumors is another example of the multiple functions of PAX6 in cell fate determination.
In this section, we have discussed that Pax6 and its downstream transcriptional factors and signaling molecules are important for neural fate determination, where these transcriptional factors have stage- and cell type-specific regulatory mechanisms. A recent preprint article has reported that a conditional knock-out of Pax6 ectopically induces GABAergic interneurons, as well as oligodendrocyte precursor cells in the cortex [67]. This might be reasonable because cyclin D2 is upregulated in the Sey/Sey mouse cortex ([94,134] and our unpublished data). Taken together, Pax6-regulated cell proliferation and neural differentiation plays pivotal roles in proper timing during brain formation.
Figure 4. The function of Pax6 and downstream genes. (A) Subcellular gene localization in radial glial (RG) cells. Genes localized in the cell nucleus, cytoplasm, and apical and basal endfeet. The graphical diagram has been redrawn from [135]. (B) Downstream genes regulated by Pax6 and their roles. Pax6 regulates Fabp7, Ninein, and Lewis X, which are involved in stem cell self-renewal. Pax6 induces Ngn2, Tbr2, and Dmrta1, which induce neural differentiation. (C) The Pax6–Ninein network regulates interkinetic nuclear migration (INM) during cell-cycle progression in neuroepithelial cells. The microtubule cytoskeleton plays an important role in INM, in which the nuclei of neuroepithelial cells move apically during G2 phase and basally during G1 phase. Ninein, downstream of Pax6, anchors microtubules during elevator movements to control the dynamics of INM. The graphical diagram has been redrawn from [99]. (D) Pax6–Notch interaction for neurogenic programs in the developing mouse cortex. Pax6-dependent neural differentiation by Notch signaling inhibition generates deeper layer (DL) neurons in the early neurogenic phase. Pax6-dependent self-renewal of the RG cells which give rise to upper layer (UL) neurons occurs in the middle/late neurogenic phase when Notch signaling is absent. In this way, the Pax6–Notch pathway coordinates the balance between self-renewal and neural differentiation. The graphical diagram has been redrawn from [127].
Figure 4. The function of Pax6 and downstream genes. (A) Subcellular gene localization in radial glial (RG) cells. Genes localized in the cell nucleus, cytoplasm, and apical and basal endfeet. The graphical diagram has been redrawn from [135]. (B) Downstream genes regulated by Pax6 and their roles. Pax6 regulates Fabp7, Ninein, and Lewis X, which are involved in stem cell self-renewal. Pax6 induces Ngn2, Tbr2, and Dmrta1, which induce neural differentiation. (C) The Pax6–Ninein network regulates interkinetic nuclear migration (INM) during cell-cycle progression in neuroepithelial cells. The microtubule cytoskeleton plays an important role in INM, in which the nuclei of neuroepithelial cells move apically during G2 phase and basally during G1 phase. Ninein, downstream of Pax6, anchors microtubules during elevator movements to control the dynamics of INM. The graphical diagram has been redrawn from [99]. (D) Pax6–Notch interaction for neurogenic programs in the developing mouse cortex. Pax6-dependent neural differentiation by Notch signaling inhibition generates deeper layer (DL) neurons in the early neurogenic phase. Pax6-dependent self-renewal of the RG cells which give rise to upper layer (UL) neurons occurs in the middle/late neurogenic phase when Notch signaling is absent. In this way, the Pax6–Notch pathway coordinates the balance between self-renewal and neural differentiation. The graphical diagram has been redrawn from [127].
Ijms 23 06115 g004
Figure 5. Linage and gene expression of NSCs in the adult hippocampal dentate gyrus. Pax6 is expressed in quiescent and active NSCs and intermediate progenitor cells (IPCs). The graphical diagram has been redrawn from [20,88,91,136,137,138,139,140,141,142,143,144].
Figure 5. Linage and gene expression of NSCs in the adult hippocampal dentate gyrus. Pax6 is expressed in quiescent and active NSCs and intermediate progenitor cells (IPCs). The graphical diagram has been redrawn from [20,88,91,136,137,138,139,140,141,142,143,144].
Ijms 23 06115 g005

4. Pax6 in Relation to Neurodevelopmental Disorders

4.1. Human PAX6 Gene Is Related to Neurodevelopmental Disorders

The human PAX6 gene was discovered during the search for the gene responsible for WAGR (Wilms tumor, aniridia, genital ridge defects, mental retardation) syndrome; patients with the syndrome often show a deletion in chromosome region 11p13 [145]. PAX6 was originally identified as AN2 (aniridia type II protein) [2] and found to be a homolog of the causative gene in Sey/Sey mice [1], as mentioned above. Another aniridia gene AN1 was previously mapped to chromosome two, although this mapping was disproven in 1992 [146]. Therefore, AN2 is now designated AN1, and PAX6 is currently shown as AN1 in the OMIN database ([147,148], https://omin.org/entry/106210 (accessed on 19 May 2022)).
A causative gene for a kidney disease (Wilms tumor), WT1, is located 0.7 Mb away from PAX6 [149]. Interestingly, a case report has suggested that a 1.6 Mb region containing PAX6, WT1, and PRRG4 is responsible for the severe developmental delays and autistic behaviors seen in WAGR syndrome [150]. There is another report showing that WAGR patients sometimes show symptoms of autism in addition to mental retardation [151].
Since aniridia is an obvious congenital disease, patients are often diagnosed genetically. There are many clinical reports showing that aniridia patients with mutations in the PAX6 gene often exhibit neural phenotypes at the structural and functional level (see [152,153] and the references therein). It has been reported, for example, that patients with a deficiency in the PAX6 3′ region containing its enhancer confer ASD and moderate mental retardation, indicating the role of PAX6 in neural phenotypes in addition to aniridia. Our group has also reported 15 single nucleotide polymorphisms within the PAX6 locus in Japanese autistic patients [154]. Currently, PAX6 is listed in the Simons Foundation Autism Research Initiative (SFARI) database in the category of “syndromic” [155,156].
Additional evidence for the involvement of the human PAX6 gene in neurodevelopmental disorders comes from a genome-wide association study (GWAS) of ASD patients [157]. It is of note that the ASD subject group, consisting of both males and females, showed a high odds ratio in the 11p13 region where PAX6 is localized, but a group consisting of only males did not show a significant odds ratio in the same region. There are not many genetic studies focusing on gender differences in ASD, but a missense mutation of CTNND2 (δ-catenin) has been identified in female-enriched families containing multiple patients [158]. Since δ-catenin is proven to be downstream of Pax6 in the embryonic mouse [159], it is speculated that PAX6 and CTNND2 might be involved in the neuropathology of a severe type of ASD often reported in girls. Our group has found that the reduction in the brain regions of rSey2/+ rats is more severe in the female than the male (Figure 6B,C) [160]. It is reasonable to assume that the functional impairment of Pax6, as well as δ-catenin—both of which are expressed in NSCs in the initial stage of cortical development—may cause a severe reduction in neurons in the cortex, resulting in mental retardation or intellectual disability.

4.2. Pax6 Deficient Rodent Models for Neurodevelopmental Disorders

The responsibility of Pax6 in ASD-like phenotypes is modeled in rodents. Cortex-specific Pax6 knock-out (Pax6fl/fl; Emx1-Cre) mice exhibit deficiencies in sensorimotor information integration, as well as both hippocampus-dependent short-term and neocortex-dependent long-term memory recalls [163]. We have identified that rSey2/+ rats show impaired sensorimotor gating, abnormal social interaction, and impaired rearing activity, fear-conditioning, memory, and vocal behavior in pups [164,165]. We have further analyzed the behaviors of Sey/+ mice and found abnormal vocalization in Sey/+ pups derived from young fathers and an increase in hyperactivity in those derived from aged fathers [166]. The fact that mice with a single genetic risk factor, Pax6, can develop different phenotypes depending on paternal age has alerted basic researchers to the need for considering not only genetic factors but also non-genetic factors in animal models. Researchers had been careful about the age of female mice yet forgot about that of male mice. Another lesson learned here is that non-genetic factors (e.g., paternal age) can mask causative genes in the genetic analysis of neurodevelopmental disorders such as ASD and attention deficit/hyperactivity.
Another Pax6 mutant strain, Pax6Leca2, exhibits an impaired retinal structure [167], alteration of circadian clock, and hyperactivity during the light period [168], which is partly consistent with our study in Sey/+ mice derived from aged fathers [166]. Interestingly, a core circadian clock gene, Clock, positively regulates Pax6 [169], and Pax6 mutants show an altered expression of circadian clock genes [168]. Our group has also elucidated that Pax6 regulates expression of the Fabp7 gene [170], another candidate associated with circadian rhythm [171] (also involved in the dataset by [172]). Sleep is regulated by circadian rhythm, which is considered as a translationally relevant endpoint in studies of ASD [173]. Therefore, Pax6-circadian clock gene regulation in the brain may influence the pathogenesis of ASD. As supportive evidence, patients with PAX6 mutation who exhibit aniridia also have a smaller pineal gland, where melatonin is produced, and are often diagnosed with sleep disorders [174]. Analyses using Pax6 mutant mice suggest that the dysfunction of the eyes might be the direct cause of the alteration of the circadian clock, rather than an impairment of the pineal gland in the brain [168]. Detailed investigation of Pax6-circadian clock gene regulation in the eyes and brain will shed further light on the pathogenesis of ASD and comorbid sleep disorders.

5. Closing Remark

Pax6 is one of the key molecules from the initial stage of neural development. It regulates brain patterning and the balance between cell proliferation and neural differentiation in an appropriate manner to developmental timing. Multiple context-dependent functions of Pax6 are governed by numerous downstream factors. Patients with mutations in the PAX6 gene may exhibit symptoms of neurodevelopmental disorders, including ASD. Since Pax6 mutant rodents exhibit diverse neurodevelopmental phenotypes, they could be used as a tool to elucidate the human pathology of neurodevelopmental disorders. It has also been reported that a Pax6 partner, REST, is expressed from the young to the aged human brain and that the expression level of REST is proportionally related to cognitive function [175]. Thus, the PAX6/REST complex might also be involved in neurodevelopmental and psychiatric disorders, possibly through the regulation of the cell fate decisions of NSCs during embryonic, postnatal, and adult neurogenesis. Further studies on Pax6 and its partner/downstream molecules may lead to the elucidation of the basic molecular mechanisms of neuropsychiatric disorders for future therapeutics.

Author Contributions

Conceptualization, S.O. and N.O.; immunohistochemical analysis, S.O. and S.M.; figure preparation, S.O., S.M. and T.K.; writing—review and editing, S.O., T.K. and N.O.; supervision, N.O.; project administration, S.O. and N.O.; funding acquisition, S.O. and N.O. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the JSPS KAKENHI (grant-in-aid for early-career scientists) under grant number JP30884217, and AMED under grant number JP21wm0425003.

Institutional Review Board Statement

This animal study protocol was approved by the Institutional for Animal Experimentation Tohoku University Graduate School of Medicine (protocol code, 2020 Medical Zoology-109-04; date of approval, 26 January 2022).

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

The authors are grateful to S. Makino for maintaining experimental animals and J. Germeraad for English proofreading. This work is supported by Neuro Global Program, Tohoku University. The immunostaining images were acquired in the Biomedical Research Core of Tohoku University Graduate School of Medicine.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Hill, R.E.; Favor, J.; Hogan, B.L.; Ton, C.C.; Saunders, G.F.; Hanson, I.M.; Prosser, J.; Jordan, T.; Hastie, N.D.; van Heyningen, V. Mouse Small eye results from mutations in a paired-like homeobox-containing gene. Nature 1991, 355, 750. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. Ton, C.C.; Hirvonen, H.; Miwa, H.; Weil, M.M.; Monaghan, P.; Jordan, T.; van Heyningen, V.; Hastie, N.D.; Meijers-Heijboer, H.; Drechsler, M.; et al. Positional cloning and characterization of a paired box- and homeobox-containing gene from the aniridia region. Cell 1991, 67, 1059–1074. [Google Scholar] [CrossRef] [Green Version]
  3. Walther, C.; Gruss, P. Pax-6, a murine paired box gene, is expressed in the developing CNS. Development 1991, 113, 1435–1449. [Google Scholar] [CrossRef] [PubMed]
  4. Callaerts, P.; Halder, G.; Gehring, W.J. PAX-6 in development and evolution. Annu. Rev. Neurosci. 1997, 20, 483–532. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  5. Chalepakis, G.; Fritsch, R.; Fickenscher, H.; Deutsch, U.; Goulding, M.; Gruss, P. The molecular basis of the undulated/Pax-1 mutation. Cell 1991, 66, 873–884. [Google Scholar] [CrossRef]
  6. Treisman, J.; Harris, E.; Desplan, C. The paired box encodes a second DNA-binding domain in the paired homeo domain protein. Genes Dev. 1991, 5, 594–604. [Google Scholar] [CrossRef] [Green Version]
  7. Halder, G.; Callaerts, P.; Gehring, W.J. Induction of ectopic eyes by targeted expression of the eyeless gene in Drosophila. Science 1995, 267, 1788–1792. [Google Scholar] [CrossRef] [Green Version]
  8. Kozmik, Z. Pax genes in eye development and evolution. Curr. Opin. Genet. Dev. 2005, 15, 430–438. [Google Scholar] [CrossRef]
  9. Haubst, N.; Berger, J.; Radjendirane, V.; Graw, J.; Favor, J.; Saunders, G.F.; Stoykova, A.; Götz, M. Molecular dissection of Pax6 function: The specific roles of the paired domain and homeodomain in brain development. Development 2004, 131, 6131–6140. [Google Scholar] [CrossRef] [Green Version]
  10. Roberts, R.C. Small eyes—A new dominant eye mutant in the mouse. Genet. Res. 1967, 9, 121–122. [Google Scholar] [CrossRef] [Green Version]
  11. Hogan, B.L.; Hirst, E.M.; Horsburgh, G.; Hetherington, C.M. Small eye (Sey): A mouse model for the genetic analysis of craniofacial abnormalities. Development 1988, 103, 115–119. [Google Scholar] [CrossRef] [PubMed]
  12. Matsuo, T.; Osumi-Yamashita, N.; Noji, S.; Ohuchi, H.; Koyama, E.; Myokai, F.; Matsuo, N.; Taniguchi, S.; Doi, H.; Iseki, S.; et al. A mutation in the Pax-6 gene in rat small eye is associated with impaired migration of midbrain crest cells. Nat. Genet. 1993, 3, 299–304. [Google Scholar] [CrossRef] [PubMed]
  13. Osumi, N. The role of Pax6 in brain patterning. Tohoku J. Exp. Med. 2001, 193, 163–174. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  14. Quiring, R.; Walldorf, U.; Kloter, U.; Gehring, W.J. Homology of the eyeless gene of Drosophila to the Small eye gene in mice and Aniridia in humans. Science 1994, 265, 785–789. [Google Scholar] [CrossRef]
  15. Inoue, T.; Nakamura, S.; Osumi, N. Fate mapping of the mouse prosencephalic neural plate. Dev. Biol. 2000, 219, 373–383. [Google Scholar] [CrossRef] [Green Version]
  16. Stoykova, A.; Treichel, D.; Hallonet, M.; Gruss, P. Pax6 modulates the dorsoventral patterning of the mammalian telencephalon. J. Neurosci. 2000, 20, 8042–8050. [Google Scholar] [CrossRef] [Green Version]
  17. Mastick, G.S.; Andrews, G.L. Pax6 regulates the identity of embryonic diencephalic neurons. Mol. Cell Neurosci. 2001, 17, 190–207. [Google Scholar] [CrossRef]
  18. Pinson, J.; Mason, J.O.; Simpson, T.I.; Price, D.J. Regulation of the Pax6: Pax6(5a) mRNA ratio in the developing mammalian brain. BMC Dev. Biol. 2005, 5, 13. [Google Scholar] [CrossRef] [Green Version]
  19. Heyman, I.; Faissner, A.; Lumsden, A. Cell and matrix specialisations of rhombomere boundaries. Dev. Dyn. 1995, 204, 301–315. [Google Scholar] [CrossRef]
  20. Osumi, N.; Shinohara, H.; Numayama-Tsuruta, K.; Maekawa, M. Concise review: Pax6 transcription factor contributes to both embryonic and adult neurogenesis as a multifunctional regulator. Stem Cells 2008, 26, 1663–1672. [Google Scholar] [CrossRef]
  21. Püschel, A.W.; Gruss, P.; Westerfield, M. Sequence and expression pattern of pax-6 are highly conserved between zebrafish and mice. Development 1992, 114, 643–651. [Google Scholar] [CrossRef] [PubMed]
  22. Ikeda, K.; Onimaru, H.; Inada, H.; Tien Lin, S.; Arata, S.; Osumi, N. Structural and functional defects of the respiratory neural system in the medulla and spinal cord of Pax6 mutant rats. Brain Res. Bull. 2019, 152, 107–116. [Google Scholar] [CrossRef] [PubMed]
  23. Fernandez, A.S.; Pieau, C.; Repérant, J.; Boncinelli, E.; Wassef, M. Expression of the Emx-1 and Dlx-1 homeobox genes define three molecularly distinct domains in the telencephalon of mouse, chick, turtle and frog embryos: Implications for the evolution of telencephalic subdivisions in amniotes. Development 1998, 125, 2099–2111. [Google Scholar] [CrossRef] [PubMed]
  24. Götz, M.; Stoykova, A.; Gruss, P. Pax6 controls radial glia differentiation in the cerebral cortex. Neuron 1998, 21, 1031–1044. [Google Scholar] [CrossRef] [Green Version]
  25. Kohwi, M.; Osumi, N.; Rubenstein, J.L.; Alvarez-Buylla, A. Pax6 is required for making specific subpopulations of granule and periglomerular neurons in the olfactory bulb. J. Neurosci. 2005, 25, 6997–7003. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  26. Soma, M.; Aizawa, H.; Ito, Y.; Maekawa, M.; Osumi, N.; Nakahira, E.; Okamoto, H.; Tanaka, K.; Yuasa, S. Development of the mouse amygdala as revealed by enhanced green fluorescent protein gene transfer by means of in utero electroporation. J. Comp. Neurol. 2009, 513, 113–128. [Google Scholar] [CrossRef] [PubMed]
  27. Pratt, T.; Vitalis, T.; Warren, N.; Edgar, J.M.; Mason, J.O.; Price, D.J. A role for Pax6 in the normal development of dorsal thalamus and its cortical connections. Development 2000, 127, 5167–5178. [Google Scholar] [CrossRef]
  28. Engelkamp, D.; Rashbass, P.; Seawright, A.; van Heyningen, V. Role of Pax6 in development of the cerebellar system. Development 1999, 126, 3585–3596. [Google Scholar] [CrossRef]
  29. Stoykova, A.; Fritsch, R.; Walther, C.; Gruss, P. Forebrain patterning defects in Small eye mutant mice. Development 1996, 122, 3453–3465. [Google Scholar] [CrossRef]
  30. Ashery-Padan, R.; Marquardt, T.; Zhou, X.; Gruss, P. Pax6 activity in the lens primordium is required for lens formation and for correct placement of a single retina in the eye. Genes Dev. 2000, 14, 2701–2711. [Google Scholar] [CrossRef] [Green Version]
  31. Kim, J.; Lauderdale, J.D. Overexpression of pairedless Pax6 in the retina disrupts corneal development and affects lens cell survival. Dev. Biol. 2008, 313, 434–454. [Google Scholar] [CrossRef] [PubMed]
  32. Sivak, J.M.; Mohan, R.; Rinehart, W.B.; Xu, P.X.; Maas, R.L.; Fini, M.E. Pax-6 expression and activity are induced in the reepithelializing cornea and control activity of the transcriptional promoter for matrix metalloproteinase gelatinase B. Dev. Biol. 2000, 222, 41–54. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  33. Davis, J.; Duncan, M.K.; Robison, W.G., Jr.; Piatigorsky, J. Requirement for Pax6 in corneal morphogenesis: A role in adhesion. J. Cell Sci. 2003, 116, 2157–2167. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  34. Macdonald, R.; Barth, K.A.; Xu, Q.; Holder, N.; Mikkola, I.; Wilson, S.W. Midline signalling is required for Pax gene regulation and patterning of the eyes. Development 1995, 121, 3267–3278. [Google Scholar] [CrossRef] [PubMed]
  35. Mui, S.H.; Kim, J.W.; Lemke, G.; Bertuzzi, S. Vax genes ventralize the embryonic eye. Genes Dev. 2005, 19, 1249–1259. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  36. Suzuki, J.; Sakurai, K.; Yamazaki, M.; Abe, M.; Inada, H.; Sakimura, K.; Katori, Y.; Osumi, N. Horizontal Basal Cell-Specific Deletion of Pax6 Impedes Recovery of the Olfactory Neuroepithelium Following Severe Injury. Stem Cells Dev. 2015, 24, 1923–1933. [Google Scholar] [CrossRef] [Green Version]
  37. Cvekl, A.; Callaerts, P. PAX6: 25th anniversary and more to learn. Exp. Eye Res. 2017, 156, 10–21. [Google Scholar] [CrossRef]
  38. Kioussi, C.; O’Connell, S.; St-Onge, L.; Treier, M.; Gleiberman, A.S.; Gruss, P.; Rosenfeld, M.G. Pax6 is essential for establishing ventral-dorsal cell boundaries in pituitary gland development. Proc. Natl. Acad. Sci. USA 1999, 96, 14378–14382. [Google Scholar] [CrossRef] [Green Version]
  39. Dohrmann, C.; Gruss, P.; Lemaire, L. Pax genes and the differentiation of hormone-producing endocrine cells in the pancreas. Mech. Dev. 2000, 92, 47–54. [Google Scholar] [CrossRef]
  40. Kimura, R.; Yoshizaki, K.; Osumi, N. Dynamic expression patterns of Pax6 during spermatogenesis in the mouse. J. Anat. 2015, 227, 1–9. [Google Scholar] [CrossRef]
  41. Shaham, O.; Menuchin, Y.; Farhy, C.; Ashery-Padan, R. Pax6: A multi-level regulator of ocular development. Prog. Retin. Eye Res. 2012, 31, 351–376. [Google Scholar] [CrossRef] [PubMed]
  42. Kikkawa, T.; Casingal, C.R.; Chun, S.H.; Shinohara, H.; Hiraoka, K.; Osumi, N. The role of Pax6 in brain development and its impact on pathogenesis of autism spectrum disorder. Brain Res. 2019, 1705, 95–103. [Google Scholar] [CrossRef] [PubMed]
  43. Panneerselvam, A.; Kannan, A.; Mariajoseph-Antony, L.F.; Prahalathan, C. PAX proteins and their role in pancreas. Diabetes Res. Clin. Pract. 2019, 155, 107792. [Google Scholar] [CrossRef] [PubMed]
  44. Manuel, M.N.; Mi, D.; Mason, J.O.; Price, D.J. Regulation of cerebral cortical neurogenesis by the Pax6 transcription factor. Front. Cell Neurosci. 2015, 9, 70. [Google Scholar] [CrossRef] [Green Version]
  45. Epstein, J.A.; Glaser, T.; Cai, J.; Jepeal, L.; Walton, D.S.; Maas, R.L. Two independent and interactive DNA-binding subdomains of the Pax6 paired domain are regulated by alternative splicing. Genes Dev. 1994, 8, 2022–2034. [Google Scholar] [CrossRef] [Green Version]
  46. Ypsilanti, A.R.; Rubenstein, J.L. Transcriptional and epigenetic mechanisms of early cortical development: An examination of how Pax6 coordinates cortical development. J. Comp. Neurol. 2016, 524, 609–629. [Google Scholar] [CrossRef] [Green Version]
  47. Peng, G.; Suo, S.; Cui, G.; Yu, F.; Wang, R.; Chen, J.; Chen, S.; Liu, Z.; Chen, G.; Qian, Y.; et al. Molecular architecture of lineage allocation and tissue organization in early mouse embryo. Nature 2019, 572, 528–532. [Google Scholar] [CrossRef]
  48. Sagner, A.; Briscoe, J. Establishing neuronal diversity in the spinal cord: A time and a place. Development 2019, 146, dev182154. [Google Scholar] [CrossRef] [Green Version]
  49. Ericson, J.; Rashbass, P.; Schedl, A.; Brenner-Morton, S.; Kawakami, A.; van Heyningen, V.; Jessell, T.M.; Briscoe, J. Pax6 controls progenitor cell identity and neuronal fate in response to graded Shh signaling. Cell 1997, 90, 169–180. [Google Scholar] [CrossRef] [Green Version]
  50. Osumi, N.; Hirota, A.; Ohuchi, H.; Nakafuku, M.; Iimura, T.; Kuratani, S.; Fujiwara, M.; Noji, S.; Eto, K. Pax-6 is involved in the specification of hindbrain motor neuron subtype. Development 1997, 124, 2961–2972. [Google Scholar] [CrossRef]
  51. O’Leary, D.D.; Chou, S.J.; Sahara, S. Area patterning of the mammalian cortex. Neuron 2007, 56, 252–269. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  52. Theil, T.; Aydin, S.; Koch, S.; Grotewold, L.; Rüther, U. Wnt and Bmp signalling cooperatively regulate graded Emx2 expression in the dorsal telencephalon. Development 2002, 129, 3045–3054. [Google Scholar] [CrossRef] [PubMed]
  53. Fukuchi-Shimogori, T.; Grove, E.A. Emx2 patterns the neocortex by regulating FGF positional signaling. Nat. Neurosci. 2003, 6, 825–831. [Google Scholar] [CrossRef] [PubMed]
  54. Garel, S.; Huffman, K.J.; Rubenstein, J.L. Molecular regionalization of the neocortex is disrupted in Fgf8 hypomorphic mutants. Development 2003, 130, 1903–1914. [Google Scholar] [CrossRef] [Green Version]
  55. Sahara, S.; Kawakami, Y.; Izpisua Belmonte, J.C.; O’Leary, D.D. Sp8 exhibits reciprocal induction with Fgf8 but has an opposing effect on anterior-posterior cortical area patterning. Neural Dev. 2007, 2, 10. [Google Scholar] [CrossRef] [Green Version]
  56. Cholfin, J.A.; Rubenstein, J.L. Frontal cortex subdivision patterning is coordinately regulated by Fgf8, Fgf17, and Emx2. J. Comp. Neurol. 2008, 509, 144–155. [Google Scholar] [CrossRef] [Green Version]
  57. Bishop, K.M.; Goudreau, G.; O’Leary, D.D. Regulation of area identity in the mammalian neocortex by Emx2 and Pax6. Science 2000, 288, 344–349. [Google Scholar] [CrossRef]
  58. Zhou, C.; Tsai, S.Y.; Tsai, M.J. COUP-TFI: An intrinsic factor for early regionalization of the neocortex. Genes Dev. 2001, 15, 2054–2059. [Google Scholar] [CrossRef] [Green Version]
  59. Zembrzycki, A.; Griesel, G.; Stoykova, A.; Mansouri, A. Genetic interplay between the transcription factors Sp8 and Emx2 in the patterning of the forebrain. Neural Dev. 2007, 2, 8. [Google Scholar] [CrossRef] [Green Version]
  60. Bishop, K.M.; Rubenstein, J.L.; O’Leary, D.D. Distinct actions of Emx1, Emx2, and Pax6 in regulating the specification of areas in the developing neocortex. J. Neurosci. 2002, 22, 7627–7638. [Google Scholar] [CrossRef] [Green Version]
  61. Hamasaki, T.; Leingärtner, A.; Ringstedt, T.; O’Leary, D.D. EMX2 regulates sizes and positioning of the primary sensory and motor areas in neocortex by direct specification of cortical progenitors. Neuron 2004, 43, 359–372. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  62. Tyas, D.A.; Simpson, T.I.; Carr, C.B.; Kleinjan, D.A.; van Heyningen, V.; Mason, J.O.; Price, D.J. Functional conservation of Pax6 regulatory elements in humans and mice demonstrated with a novel transgenic reporter mouse. BMC Dev. Biol. 2006, 6, 21. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  63. Manuel, M.; Georgala, P.A.; Carr, C.B.; Chanas, S.; Kleinjan, D.A.; Martynoga, B.; Mason, J.O.; Molinek, M.; Pinson, J.; Pratt, T.; et al. Controlled overexpression of Pax6 in vivo negatively autoregulates the Pax6 locus, causing cell-autonomous defects of late cortical progenitor proliferation with little effect on cortical arealization. Development 2007, 134, 545–555. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  64. Borello, U.; Cobos, I.; Long, J.E.; McWhirter, J.R.; Murre, C.; Rubenstein, J.L. FGF15 promotes neurogenesis and opposes FGF8 function during neocortical development. Neural Dev. 2008, 3, 17. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  65. Giacomantonio, C.E.; Goodhill, G.J. A Boolean model of the gene regulatory network underlying Mammalian cortical area development. PLoS Comput. Biol. 2010, 6, e1000936. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  66. Giacomantonio, C.E.; Goodhill, G.J. A computational model of the effect of gene misexpression on the development of cortical areas. Biol. Cybern. 2014, 108, 203–221. [Google Scholar] [CrossRef]
  67. Manuel, M.; Tan, B.K.; Kozic, Z.; Molinek, M.; Marcos, S.T.; Razak, A.F.M.; Dobolyi, D.; Dobie, R.; Henderson, P.E.B.; Henderson, C.N.; et al. Pax6 limits the competence of developing cerebral cortical cells. bioRxiv 2022. [Google Scholar] [CrossRef]
  68. Armentano, M.; Chou, S.J.; Tomassy, G.S.; Leingärtner, A.; O’Leary, D.D.; Studer, M. COUP-TFI regulates the balance of cortical patterning between frontal/motor and sensory areas. Nat. Neurosci. 2007, 10, 1277–1286. [Google Scholar] [CrossRef]
  69. McBride, D.J.; Buckle, A.; van Heyningen, V.; Kleinjan, D.A. DNaseI hypersensitivity and ultraconservation reveal novel, interdependent long-range enhancers at the complex Pax6 cis-regulatory region. PLoS ONE 2011, 6, e28616. [Google Scholar] [CrossRef] [Green Version]
  70. Clowry, G.J.; Alzu’bi, A.; Harkin, L.F.; Sarma, S.; Kerwin, J.; Lindsay, S.J. Charting the protomap of the human telencephalon. Semin. Cell Dev. Biol. 2018, 76, 3–14. [Google Scholar] [CrossRef] [Green Version]
  71. Puelles, L.; Kuwana, E.; Puelles, E.; Bulfone, A.; Shimamura, K.; Keleher, J.; Smiga, S.; Rubenstein, J.L. Pallial and subpallial derivatives in the embryonic chick and mouse telencephalon, traced by the expression of the genes Dlx-2, Emx-1, Nkx-2.1, Pax-6, and Tbr-1. J. Comp. Neurol. 2000, 424, 409–438. [Google Scholar] [CrossRef]
  72. Corbin, J.G.; Gaiano, N.; Machold, R.P.; Langston, A.; Fishell, G. The Gsh2 homeodomain gene controls multiple aspects of telencephalic development. Development 2000, 127, 5007–5020. [Google Scholar] [CrossRef] [PubMed]
  73. Carney, R.S.; Cocas, L.A.; Hirata, T.; Mansfield, K.; Corbin, J.G. Differential regulation of telencephalic pallial-subpallial boundary patterning by Pax6 and Gsh2. Cereb. Cortex 2009, 19, 745–759. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  74. Toresson, H.; Potter, S.S.; Campbell, K. Genetic control of dorsal-ventral identity in the telencephalon: Opposing roles for Pax6 and Gsh2. Development 2000, 127, 4361–4371. [Google Scholar] [CrossRef]
  75. Yun, K.; Potter, S.; Rubenstein, J.L. Gsh2 and Pax6 play complementary roles in dorsoventral patterning of the mammalian telencephalon. Development 2001, 128, 193–205. [Google Scholar] [CrossRef]
  76. Fode, C.; Ma, Q.; Casarosa, S.; Ang, S.L.; Anderson, D.J.; Guillemot, F. A role for neural determination genes in specifying the dorsoventral identity of telencephalic neurons. Genes Dev. 2000, 14, 67–80. [Google Scholar] [CrossRef]
  77. Kikkawa, T.; Obayashi, T.; Takahashi, M.; Fukuzaki-Dohi, U.; Numayama-Tsuruta, K.; Osumi, N. Dmrta1 regulates proneural gene expression downstream of Pax6 in the mammalian telencephalon. Genes Cells 2013, 18, 636–649. [Google Scholar] [CrossRef]
  78. Kikkawa, T.; Osumi, N. Multiple Functions of the Dmrt Genes in the Development of the Central Nervous System. Front. Neurosci. 2021, 15, 789583. [Google Scholar] [CrossRef]
  79. Chenn, A.; McConnell, S.K. Cleavage orientation and the asymmetric inheritance of Notch1 immunoreactivity in mammalian neurogenesis. Cell 1995, 82, 631–641. [Google Scholar] [CrossRef] [Green Version]
  80. Kageyama, R.; Ochi, S.; Sueda, R.; Shimojo, H. The significance of gene expression dynamics in neural stem cell regulation. Proc. Jpn. Acad. Ser. B Phys. Biol. Sci. 2020, 96, 351–363. [Google Scholar] [CrossRef]
  81. Watanabe, N.; Kageyama, R.; Ohtsuka, T. Hbp1 regulates the timing of neuronal differentiation during cortical development by controlling cell cycle progression. Development 2015, 142, 2278–2290. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  82. Manuel, M.; Price, D.J. Role of Pax6 in forebrain regionalization. Brain Res. Bull. 2005, 66, 387–393. [Google Scholar] [CrossRef] [PubMed]
  83. Englund, C.; Fink, A.; Lau, C.; Pham, D.; Daza, R.A.; Bulfone, A.; Kowalczyk, T.; Hevner, R.F. Pax6, Tbr2, and Tbr1 are expressed sequentially by radial glia, intermediate progenitor cells, and postmitotic neurons in developing neocortex. J. Neurosci. 2005, 25, 247–251. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  84. Hevner, R.F.; Hodge, R.D.; Daza, R.A.; Englund, C. Transcription factors in glutamatergic neurogenesis: Conserved programs in neocortex, cerebellum, and adult hippocampus. Neurosci. Res. 2006, 55, 223–233. [Google Scholar] [CrossRef] [PubMed]
  85. Imayoshi, I.; Sakamoto, M.; Ohtsuka, T.; Kageyama, R. Continuous neurogenesis in the adult brain. Dev. Growth Differ. 2009, 51, 379–386. [Google Scholar] [CrossRef]
  86. Urbán, N.; Guillemot, F. Neurogenesis in the embryonic and adult brain: Same regulators, different roles. Front. Cell Neurosci. 2014, 8, 396. [Google Scholar] [CrossRef] [Green Version]
  87. Kobayashi, T.; Kageyama, R. Lysosomes and signaling pathways for maintenance of quiescence in adult neural stem cells. FEBS J. 2021, 288, 3082–3093. [Google Scholar] [CrossRef]
  88. Maekawa, M.; Takashima, N.; Arai, Y.; Nomura, T.; Inokuchi, K.; Yuasa, S.; Osumi, N. Pax6 is required for production and maintenance of progenitor cells in postnatal hippocampal neurogenesis. Genes Cells 2005, 10, 1001–1014. [Google Scholar] [CrossRef]
  89. Nacher, J.; Varea, E.; Blasco-Ibañez, J.M.; Castillo-Gomez, E.; Crespo, C.; Martinez-Guijarro, F.J.; McEwen, B.S. Expression of the transcription factor Pax 6 in the adult rat dentate gyrus. J. Neurosci. Res. 2005, 81, 753–761. [Google Scholar] [CrossRef]
  90. Sugiyama, T.; Osumi, N.; Katsuyama, Y. The germinal matrices in the developing dentate gyrus are composed of neuronal progenitors at distinct differentiation stages. Dev. Dyn. 2013, 242, 1442–1453. [Google Scholar] [CrossRef]
  91. Sakurai, K.; Osumi, N. The neurogenesis-controlling factor, Pax6, inhibits proliferation and promotes maturation in murine astrocytes. J. Neurosci. 2008, 28, 4604–4612. [Google Scholar] [CrossRef] [PubMed]
  92. Ghosh, M.; Lane, M.; Krizman, E.; Sattler, R.; Rothstein, J.D.; Robinson, M.B. The transcription factor Pax6 contributes to the induction of GLT-1 expression in astrocytes through an interaction with a distal enhancer element. J. Neurochem. 2016, 136, 262–275. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  93. Tsunekawa, Y.; Britto, J.M.; Takahashi, M.; Polleux, F.; Tan, S.S.; Osumi, N. Cyclin D2 in the basal process of neural progenitors is linked to non-equivalent cell fates. Embo. J. 2012, 31, 1879–1892. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  94. Mi, D.; Carr, C.B.; Georgala, P.A.; Huang, Y.T.; Manuel, M.N.; Jeanes, E.; Niisato, E.; Sansom, S.N.; Livesey, F.J.; Theil, T.; et al. Pax6 exerts regional control of cortical progenitor proliferation via direct repression of Cdk6 and hypophosphorylation of pRb. Neuron 2013, 78, 269–284. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  95. Estivill-Torrus, G.; Pearson, H.; van Heyningen, V.; Price, D.J.; Rashbass, P. Pax6 is required to regulate the cell cycle and the rate of progression from symmetrical to asymmetrical division in mammalian cortical progenitors. Development 2002, 129, 455–466. [Google Scholar] [CrossRef]
  96. Fujita, S. Mitotic pattern and histogenesis of the central nervous system. Nature 1960, 185, 702–703. [Google Scholar] [CrossRef]
  97. Fujita, S. Kinetics of cellular proliferation. Exp. Cell Res. 1962, 28, 52–60. [Google Scholar] [CrossRef]
  98. Miyata, T.; Kawaguchi, A.; Okano, H.; Ogawa, M. Asymmetric inheritance of radial glial fibers by cortical neurons. Neuron 2001, 31, 727–741. [Google Scholar] [CrossRef] [Green Version]
  99. Shinohara, H.; Sakayori, N.; Takahashi, M.; Osumi, N. Ninein is essential for the maintenance of the cortical progenitor character by anchoring the centrosome to microtubules. Biol. Open 2013, 2, 739–749. [Google Scholar] [CrossRef] [Green Version]
  100. Namihira, M.; Kohyama, J.; Abematsu, M.; Nakashima, K. Epigenetic mechanisms regulating fate specification of neural stem cells. Philos. Trans. R. Soc. Lond. B Biol. Sci. 2008, 363, 2099–2109. [Google Scholar] [CrossRef] [Green Version]
  101. Hirabayashi, Y.; Gotoh, Y. Epigenetic control of neural precursor cell fate during development. Nat. Rev. Neurosci. 2010, 11, 377–388. [Google Scholar] [CrossRef] [PubMed]
  102. Tyssowski, K.; Kishi, Y.; Gotoh, Y. Chromatin regulators of neural development. Neuroscience 2014, 264, 4–16. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  103. Albert, M.; Huttner, W.B. Epigenetic and Transcriptional Pre-patterning-An Emerging Theme in Cortical Neurogenesis. Front. Neurosci. 2018, 12, 359. [Google Scholar] [CrossRef] [Green Version]
  104. Ninkovic, J.; Steiner-Mezzadri, A.; Jawerka, M.; Akinci, U.; Masserdotti, G.; Petricca, S.; Fischer, J.; von Holst, A.; Beckers, J.; Lie, C.D.; et al. The BAF complex interacts with Pax6 in adult neural progenitors to establish a neurogenic cross-regulatory transcriptional network. Cell Stem Cell 2013, 13, 403–418. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  105. Tuoc, T.C.; Boretius, S.; Sansom, S.N.; Pitulescu, M.E.; Frahm, J.; Livesey, F.J.; Stoykova, A. Chromatin regulation by BAF170 controls cerebral cortical size and thickness. Dev. Cell 2013, 25, 256–269. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  106. Tuoc, T.C.; Narayanan, R.; Stoykova, A. BAF chromatin remodeling complex: Cortical size regulation and beyond. Cell Cycle 2013, 12, 2953–2959. [Google Scholar] [CrossRef] [Green Version]
  107. Nechiporuk, T.; McGann, J.; Mullendorff, K.; Hsieh, J.; Wurst, W.; Floss, T.; Mandel, G. The REST remodeling complex protects genomic integrity during embryonic neurogenesis. Elife 2016, 5, e09584. [Google Scholar] [CrossRef]
  108. Kimura, R.; Inoue, Y.U.; Kikkawa, T.; Tatehana, M.; Morimoto, Y.; Inada, H.; Oki, S.; Inoue, T.; Osumi, N. Detection of REST expression in the testis using epitope-tag knock-in mice generated by genome editing. Dev. Dyn. 2022, 251, 525–535. [Google Scholar] [CrossRef]
  109. Sansom, S.N.; Griffiths, D.S.; Faedo, A.; Kleinjan, D.J.; Ruan, Y.; Smith, J.; van Heyningen, V.; Rubenstein, J.L.; Livesey, F.J. The level of the transcription factor Pax6 is essential for controlling the balance between neural stem cell self-renewal and neurogenesis. PLoS Genet. 2009, 5, e1000511. [Google Scholar] [CrossRef] [Green Version]
  110. Saffary, R.; Xie, Z. FMRP regulates the transition from radial glial cells to intermediate progenitor cells during neocortical development. J. Neurosci. 2011, 31, 1427–1439. [Google Scholar] [CrossRef] [Green Version]
  111. Pilaz, L.J.; Lennox, A.L.; Rouanet, J.P.; Silver, D.L. Dynamic mRNA Transport and Local Translation in Radial Glial Progenitors of the Developing Brain. Curr. Biol. 2016, 26, 3383–3392. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  112. Casingal, C.R.; Kikkawa, T.; Inada, H.; Sasaki, Y.; Osumi, N. Identification of FMRP target mRNAs in the developmental brain: FMRP might coordinate Ras/MAPK, Wnt/β-catenin, and mTOR signaling during corticogenesis. Mol. Brain 2020, 13, 167. [Google Scholar] [CrossRef] [PubMed]
  113. Bertrand, N.; Castro, D.S.; Guillemot, F. Proneural genes and the specification of neural cell types. Nat. Rev. Neurosci. 2002, 3, 517–530. [Google Scholar] [CrossRef] [PubMed]
  114. Guillemot, F.; Hassan, B.A. Beyond proneural: Emerging functions and regulations of proneural proteins. Curr. Opin. Neurobiol. 2017, 42, 93–101. [Google Scholar] [CrossRef]
  115. Baker, N.E.; Brown, N.L. All in the family: Proneural bHLH genes and neuronal diversity. Development 2018, 145, dev159426. [Google Scholar] [CrossRef] [Green Version]
  116. Scardigli, R.; Schuurmans, C.; Gradwohl, G.; Guillemot, F. Crossregulation between Neurogenin2 and pathways specifying neuronal identity in the spinal cord. Neuron 2001, 31, 203–217. [Google Scholar] [CrossRef] [Green Version]
  117. Scardigli, R.; Bäumer, N.; Gruss, P.; Guillemot, F.; Le Roux, I. Direct and concentration-dependent regulation of the proneural gene Neurogenin2 by Pax6. Development 2003, 130, 3269–3281. [Google Scholar] [CrossRef] [Green Version]
  118. Ochiai, W.; Nakatani, S.; Takahara, T.; Kainuma, M.; Masaoka, M.; Minobe, S.; Namihira, M.; Nakashima, K.; Sakakibara, A.; Ogawa, M.; et al. Periventricular notch activation and asymmetric Ngn2 and Tbr2 expression in pair-generated neocortical daughter cells. Mol. Cell Neurosci. 2009, 40, 225–233. [Google Scholar] [CrossRef]
  119. Farkas, L.M.; Haffner, C.; Giger, T.; Khaitovich, P.; Nowick, K.; Birchmeier, C.; Pääbo, S.; Huttner, W.B. Insulinoma-associated 1 has a panneurogenic role and promotes the generation and expansion of basal progenitors in the developing mouse neocortex. Neuron 2008, 60, 40–55. [Google Scholar] [CrossRef] [Green Version]
  120. Mihalas, A.B.; Elsen, G.E.; Bedogni, F.; Daza, R.A.M.; Ramos-Laguna, K.A.; Arnold, S.J.; Hevner, R.F. Intermediate Progenitor Cohorts Differentially Generate Cortical Layers and Require Tbr2 for Timely Acquisition of Neuronal Subtype Identity. Cell Rep. 2016, 16, 92–105. [Google Scholar] [CrossRef] [Green Version]
  121. Fre, S.; Bardin, A.; Robine, S.; Louvard, D. Notch signaling in intestinal homeostasis across species: The cases of Drosophila, Zebrafish and the mouse. Exp. Cell Res. 2011, 317, 2740–2747. [Google Scholar] [CrossRef] [PubMed]
  122. Artavanis-Tsakonas, S.; Rand, M.D.; Lake, R.J. Notch signaling: Cell fate control and signal integration in development. Science 1999, 284, 770–776. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  123. Shimojo, H.; Ohtsuka, T.; Kageyama, R. Oscillations in notch signaling regulate maintenance of neural progenitors. Neuron 2008, 58, 52–64. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  124. Kageyama, R.; Ohtsuka, T.; Shimojo, H.; Imayoshi, I. Dynamic regulation of Notch signaling in neural progenitor cells. Curr. Opin. Cell Biol. 2009, 21, 733–740. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  125. Ochi, S.; Imaizumi, Y.; Shimojo, H.; Miyachi, H.; Kageyama, R. Oscillatory expression of Hes1 regulates cell proliferation and neuronal differentiation in the embryonic brain. Development 2020, 147, dev182204. [Google Scholar] [CrossRef] [PubMed]
  126. Mase, S.; Shitamukai, A.; Wu, Q.; Morimoto, M.; Gridley, T.; Matsuzaki, F. Notch1 and Notch2 collaboratively maintain radial glial cells in mouse neurogenesis. Neurosci. Res. 2021, 170, 122–132. [Google Scholar] [CrossRef]
  127. Yamashita, W.; Takahashi, M.; Kikkawa, T.; Gotoh, H.; Osumi, N.; Ono, K.; Nomura, T. Conserved and divergent functions of Pax6 underlie species-specific neurogenic patterns in the developing amniote brain. Development 2018, 145, dev159764. [Google Scholar] [CrossRef] [Green Version]
  128. Zhang, X.; Huang, C.T.; Chen, J.; Pankratz, M.T.; Xi, J.; Li, J.; Yang, Y.; Lavaute, T.M.; Li, X.J.; Ayala, M.; et al. Pax6 is a human neuroectoderm cell fate determinant. Cell Stem Cell 2010, 7, 90–100. [Google Scholar] [CrossRef] [Green Version]
  129. Penisson, M.; Ladewig, J.; Belvindrah, R.; Francis, F. Genes and Mechanisms Involved in the Generation and Amplification of Basal Radial Glial Cells. Front. Cell Neurosci. 2019, 13, 381. [Google Scholar] [CrossRef]
  130. Wong, F.K.; Fei, J.F.; Mora-Bermúdez, F.; Taverna, E.; Haffner, C.; Fu, J.; Anastassiadis, K.; Stewart, A.F.; Huttner, W.B. Sustained Pax6 Expression Generates Primate-like Basal Radial Glia in Developing Mouse Neocortex. PLoS Biol. 2015, 13, e1002217. [Google Scholar] [CrossRef] [Green Version]
  131. Hu, B.; Wang, Q.; Wang, Y.A.; Hua, S.; Sauvé, C.G.; Ong, D.; Lan, Z.D.; Chang, Q.; Ho, Y.W.; Monasterio, M.M.; et al. Epigenetic Activation of WNT5A Drives Glioblastoma Stem Cell Differentiation and Invasive Growth. Cell 2016, 167, 1281–1295.e18. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  132. Hegge, B.; Sjøttem, E.; Mikkola, I. Generation of a PAX6 knockout glioblastoma cell line with changes in cell cycle distribution and sensitivity to oxidative stress. BMC Cancer 2018, 18, 496. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  133. Shafi, O.; Siddiqui, G. Tracing the origins of glioblastoma by investigating the role of gliogenic and related neurogenic genes/signaling pathways in GBM development: A systematic review. World J. Surg. Oncol. 2022, 20, 146. [Google Scholar] [CrossRef] [PubMed]
  134. Walcher, T.; Xie, Q.; Sun, J.; Irmler, M.; Beckers, J.; Öztürk, T.; Niessing, D.; Stoykova, A.; Cvekl, A.; Ninkovic, J.; et al. Functional dissection of the paired domain of Pax6 reveals molecular mechanisms of coordinating neurogenesis and proliferation. Development 2013, 140, 1123–1136. [Google Scholar] [CrossRef] [Green Version]
  135. Tsunekawa, Y.; Kikkawa, T.; Osumi, N. Asymmetric inheritance of Cyclin D2 maintains proliferative neural stem/progenitor cells: A critical event in brain development and evolution. Dev. Growth Differ. 2014, 56, 349–357. [Google Scholar] [CrossRef] [PubMed]
  136. Kim, E.J.; Leung, C.T.; Reed, R.R.; Johnson, J.E. In vivo analysis of Ascl1 defined progenitors reveals distinct developmental dynamics during adult neurogenesis and gliogenesis. J. Neurosci. 2007, 27, 12764–12774. [Google Scholar] [CrossRef] [Green Version]
  137. Roybon, L.; Hjalt, T.; Stott, S.; Guillemot, F.; Li, J.Y.; Brundin, P. Neurogenin2 directs granule neuroblast production and amplification while NeuroD1 specifies neuronal fate during hippocampal neurogenesis. PLoS ONE 2009, 4, e4779. [Google Scholar] [CrossRef] [Green Version]
  138. Hodge, R.D.; Hevner, R.F. Expression and actions of transcription factors in adult hippocampal neurogenesis. Dev. Neurobiol. 2011, 71, 680–689. [Google Scholar] [CrossRef] [Green Version]
  139. Hsieh, J. Orchestrating transcriptional control of adult neurogenesis. Genes Dev. 2012, 26, 1010–1021. [Google Scholar] [CrossRef] [Green Version]
  140. Sueda, R.; Imayoshi, I.; Harima, Y.; Kageyama, R. High Hes1 expression and resultant Ascl1 suppression regulate quiescent vs. active neural stem cells in the adult mouse brain. Genes Dev. 2019, 33, 511–523. [Google Scholar] [CrossRef] [Green Version]
  141. Urbán, N.; Blomfield, I.M.; Guillemot, F. Quiescence of Adult Mammalian Neural Stem Cells: A Highly Regulated Rest. Neuron 2019, 104, 834–848. [Google Scholar] [CrossRef] [PubMed]
  142. Urbach, A.; Witte, O.W. Divide or Commit-Revisiting the Role of Cell Cycle Regulators in Adult Hippocampal Neurogenesis. Front. Cell Dev. Biol. 2019, 7, 55. [Google Scholar] [CrossRef] [PubMed]
  143. Niklison-Chirou, M.V.; Agostini, M.; Amelio, I.; Melino, G. Regulation of Adult Neurogenesis in Mammalian Brain. Int. J. Mol. Sci. 2020, 21, 4869. [Google Scholar] [CrossRef]
  144. Denoth-Lippuner, A.; Jessberger, S. Formation and integration of new neurons in the adult hippocampus. Nat. Rev. Neurosci. 2021, 22, 223–236. [Google Scholar] [CrossRef] [PubMed]
  145. Gessler, M.; Simola, K.O.; Bruns, G.A. Cloning of breakpoints of a chromosome translocation identifies the AN2 locus. Science 1989, 244, 1575–1578. [Google Scholar] [CrossRef] [Green Version]
  146. Lyons, L.A.; Martha, A.; Mintz-Hittner, H.A.; Saunders, G.F.; Ferrell, R.E. Resolution of the two loci for autosomal dominant aniridia, AN1 and AN2, to a single locus on chromosome 11p13. Genomics 1992, 13, 925–930. [Google Scholar] [CrossRef]
  147. McKusick, V.A. Mendelian Inheritance in Man and its online version, OMIM. Am. J. Hum. Genet. 2007, 80, 588–604. [Google Scholar] [CrossRef] [Green Version]
  148. Amberger, J.; Bocchini, C.; Hamosh, A. A new face and new challenges for Online Mendelian Inheritance in Man (OMIM®). Hum. Mutat. 2011, 32, 564–567. [Google Scholar] [CrossRef]
  149. Call, K.M.; Glaser, T.; Ito, C.Y.; Buckler, A.J.; Pelletier, J.; Haber, D.A.; Rose, E.A.; Kral, A.; Yeger, H.; Lewis, W.H.; et al. Isolation and characterization of a zinc finger polypeptide gene at the human chromosome 11 Wilms’ tumor locus. Cell 1990, 60, 509–520. [Google Scholar] [CrossRef]
  150. Yamamoto, T.; Togawa, M.; Shimada, S.; Sangu, N.; Shimojima, K.; Okamoto, N. Narrowing of the responsible region for severe developmental delay and autistic behaviors in WAGR syndrome down to 1.6 Mb including PAX6, WT1, and PRRG4. Am. J. Med. Genet. A 2014, 164, 634–638. [Google Scholar] [CrossRef]
  151. Xu, S.; Han, J.C.; Morales, A.; Menzie, C.M.; Williams, K.; Fan, Y.S. Characterization of 11p14-p12 deletion in WAGR syndrome by array CGH for identifying genes contributing to mental retardation and autism. Cytogenet. Genome Res. 2008, 122, 181–187. [Google Scholar] [CrossRef] [PubMed]
  152. Davis, L.K.; Meyer, K.J.; Rudd, D.S.; Librant, A.L.; Epping, E.A.; Sheffield, V.C.; Wassink, T.H. Pax6 3′ deletion results in aniridia, autism and mental retardation. Hum. Genet. 2008, 123, 371–378. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  153. Grant, M.K.; Bobilev, A.M.; Branch, A.; Lauderdale, J.D. Structural and functional consequences of PAX6 mutations in the brain: Implications for aniridia. Brain Res. 2021, 1756, 147283. [Google Scholar] [CrossRef] [PubMed]
  154. Maekawa, M.; Iwayama, Y.; Nakamura, K.; Sato, M.; Toyota, T.; Ohnishi, T.; Yamada, K.; Miyachi, T.; Tsujii, M.; Hattori, E.; et al. A novel missense mutation (Leu46Val) of PAX6 found in an autistic patient. Neurosci. Lett. 2009, 462, 267–271. [Google Scholar] [CrossRef] [PubMed]
  155. Banerjee-Basu, S.; Packer, A. SFARI Gene: An evolving database for the autism research community. Dis. Model. Mech. 2010, 3, 133–135. [Google Scholar] [CrossRef] [Green Version]
  156. Abrahams, B.S.; Arking, D.E.; Campbell, D.B.; Mefford, H.C.; Morrow, E.M.; Weiss, L.A.; Menashe, I.; Wadkins, T.; Banerjee-Basu, S.; Packer, A. SFARI Gene 2.0: A community-driven knowledgebase for the autism spectrum disorders (ASDs). Mol. Autism 2013, 4, 36. [Google Scholar] [CrossRef] [Green Version]
  157. Szatmari, P.; Paterson, A.D.; Zwaigenbaum, L.; Roberts, W.; Brian, J.; Liu, X.Q.; Vincent, J.B.; Skaug, J.L.; Thompson, A.P.; Senman, L.; et al. Mapping autism risk loci using genetic linkage and chromosomal rearrangements. Nat. Genet. 2007, 39, 319–328. [Google Scholar] [CrossRef] [Green Version]
  158. Turner, T.N.; Sharma, K.; Oh, E.C.; Liu, Y.P.; Collins, R.L.; Sosa, M.X.; Auer, D.R.; Brand, H.; Sanders, S.J.; Moreno-De-Luca, D.; et al. Loss of δ-catenin function in severe autism. Nature 2015, 520, 51–56. [Google Scholar] [CrossRef] [Green Version]
  159. Duparc, R.H.; Boutemmine, D.; Champagne, M.P.; Tétreault, N.; Bernier, G. Pax6 is required for delta-catenin/neurojugin expression during retinal, cerebellar and cortical development in mice. Dev. Biol. 2006, 300, 647–655. [Google Scholar] [CrossRef] [Green Version]
  160. Hiraoka, K.; Sumiyoshi, A.; Nonaka, H.; Kikkawa, T.; Kawashima, R.; Osumi, N. Regional Volume Decreases in the Brain of Pax6 Heterozygous Mutant Rats: MRI Deformation-Based Morphometry. PLoS ONE 2016, 11, e0158153. [Google Scholar] [CrossRef]
  161. Duan, D.; Fu, Y.; Paxinos, G.; Watson, C. Spatiotemporal expression patterns of Pax6 in the brain of embryonic, newborn, and adult mice. Brain Struct. Funct. 2013, 218, 353–372. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  162. Paxinos, G.; Franklin, K.B.J. The Mouse Brain in Stereotaxic Coordinates, 2nd ed.; Academic Press: San Diego, CA, USA, 2001. [Google Scholar]
  163. Tuoc, T.C.; Radyushkin, K.; Tonchev, A.B.; Piñon, M.C.; Ashery-Padan, R.; Molnár, Z.; Davidoff, M.S.; Stoykova, A. Selective cortical layering abnormalities and behavioral deficits in cortex-specific Pax6 knock-out mice. J. Neurosci. 2009, 29, 8335–8349. [Google Scholar] [CrossRef] [PubMed]
  164. Maekawa, M.; Takashima, N.; Matsumata, M.; Ikegami, S.; Kontani, M.; Hara, Y.; Kawashima, H.; Owada, Y.; Kiso, Y.; Yoshikawa, T.; et al. Arachidonic acid drives postnatal neurogenesis and elicits a beneficial effect on prepulse inhibition, a biological trait of psychiatric illnesses. PLoS ONE 2009, 4, e5085. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  165. Umeda, T.; Takashima, N.; Nakagawa, R.; Maekawa, M.; Ikegami, S.; Yoshikawa, T.; Kobayashi, K.; Okanoya, K.; Inokuchi, K.; Osumi, N. Evaluation of Pax6 mutant rat as a model for autism. PLoS ONE 2010, 5, e15500. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  166. Yoshizaki, K.; Furuse, T.; Kimura, R.; Tucci, V.; Kaneda, H.; Wakana, S.; Osumi, N. Paternal Aging Affects Behavior in Pax6 Mutant Mice: A Gene/Environment Interaction in Understanding Neurodevelopmental Disorders. PLoS ONE 2016, 11, e0166665. [Google Scholar] [CrossRef]
  167. Thaung, C.; West, K.; Clark, B.J.; McKie, L.; Morgan, J.E.; Arnold, K.; Nolan, P.M.; Peters, J.; Hunter, A.J.; Brown, S.D.; et al. Novel ENU-induced eye mutations in the mouse: Models for human eye disease. Hum. Mol. Genet. 2002, 11, 755–767. [Google Scholar] [CrossRef]
  168. Chhabra, N.F.; Amarie, O.V.; Wu, M.; Amend, A.L.; Rubey, M.; Gradinger, D.; Irmler, M.; Beckers, J.; Rathkolb, B.; Wolf, E.; et al. PAX6 mutation alters circadian rhythm and β cell function in mice without affecting glucose tolerance. Commun. Biol. 2020, 3, 628. [Google Scholar] [CrossRef]
  169. Morgan, R. Pax6 is a direct, positively regulated target of the circadian gene Clock. Dev. Dyn. 2004, 230, 643–650. [Google Scholar] [CrossRef]
  170. Inada, H.; Numayama-Tsuruta, K.; Mochizuki, K.; Sasaki-Hoshino, M.; Osumi, N. Pax6-dependent regulation of the rat Fabp7 promoter activity. Genes Cells 2018, 23, 702–714. [Google Scholar] [CrossRef] [Green Version]
  171. Gerstner, J.R.; Vander Heyden, W.M.; Lavaute, T.M.; Landry, C.F. Profiles of novel diurnally regulated genes in mouse hypothalamus: Expression analysis of the cysteine and histidine-rich domain-containing, zinc-binding protein 1, the fatty acid-binding protein 7 and the GTPase, ras-like family member 11b. Neuroscience 2006, 139, 1435–1448. [Google Scholar] [CrossRef] [Green Version]
  172. Ueda, H.R.; Chen, W.; Adachi, A.; Wakamatsu, H.; Hayashi, S.; Takasugi, T.; Nagano, M.; Nakahama, K.; Suzuki, Y.; Sugano, S.; et al. A transcription factor response element for gene expression during circadian night. Nature 2002, 418, 534–539. [Google Scholar] [CrossRef] [PubMed]
  173. Missig, G.; McDougle, C.J.; Carlezon, W.A., Jr. Sleep as a translationally-relevant endpoint in studies of autism spectrum disorder (ASD). Neuropsychopharmacology 2020, 45, 90–103. [Google Scholar] [CrossRef] [PubMed]
  174. Berntsson, S.G.; Kristoffersson, A.; Daniilidou, M.; Dahl, N.; Ekström, C.; Semnic, R.; Markström, A.; Niemelä, V.; Partinen, M.; Hallböök, F.; et al. Aniridia with PAX6 mutations and narcolepsy. J. Sleep Res. 2020, 29, e12982. [Google Scholar] [CrossRef]
  175. Lu, T.; Aron, L.; Zullo, J.; Pan, Y.; Kim, H.; Chen, Y.; Yang, T.H.; Kim, H.M.; Drake, D.; Liu, X.S.; et al. REST and stress resistance in ageing and Alzheimer’s disease. Nature 2014, 507, 448–454. [Google Scholar] [CrossRef] [PubMed] [Green Version]
Figure 3. Neural stem cell (NSC) differentiation in the embryonic cortex. (A) Neuroepithelial cells undergo symmetrical cell division to increase the population of neural stem cells (NSCs, proliferation phase). When brain development progresses, neuroepithelial cells elongate their processes stretching from the apical surface of the ventricular zone (VZ) to the basal tip (basal endfoot) at the pia surface; these cells are now called radial glial (RG) cells. They undergo asymmetrical cell division and induce intermediate progenitor cells (IPCs), as well as neurons. Intermediate progenitors migrate to the subventricular zone and differentiate into neurons. Subsequently, these neurons migrate towards the basal side. After producing neurons, RG cells produce glia, i.e., astrocytes and oligodendrocytes. The graphical diagram has been redrawn from [80]. (B) Expression of Pax6, Tbr2, Tuj1 in the developing cortex at E11.5, E14.5, and E17.5: Pax6 (red), Tbr2 (blue), and Tuj1 (green). These proteins are expressed in the VZ, subventricular zone (SVZ), intermediate zone (IZ), and cortical plate (CP), respectively. Scale bar: 100 μm. (C) The transition of the relative volume of brain subdivisions in the developing cortex. The Pax6-positive VZ area gradually narrowed during development.
Figure 3. Neural stem cell (NSC) differentiation in the embryonic cortex. (A) Neuroepithelial cells undergo symmetrical cell division to increase the population of neural stem cells (NSCs, proliferation phase). When brain development progresses, neuroepithelial cells elongate their processes stretching from the apical surface of the ventricular zone (VZ) to the basal tip (basal endfoot) at the pia surface; these cells are now called radial glial (RG) cells. They undergo asymmetrical cell division and induce intermediate progenitor cells (IPCs), as well as neurons. Intermediate progenitors migrate to the subventricular zone and differentiate into neurons. Subsequently, these neurons migrate towards the basal side. After producing neurons, RG cells produce glia, i.e., astrocytes and oligodendrocytes. The graphical diagram has been redrawn from [80]. (B) Expression of Pax6, Tbr2, Tuj1 in the developing cortex at E11.5, E14.5, and E17.5: Pax6 (red), Tbr2 (blue), and Tuj1 (green). These proteins are expressed in the VZ, subventricular zone (SVZ), intermediate zone (IZ), and cortical plate (CP), respectively. Scale bar: 100 μm. (C) The transition of the relative volume of brain subdivisions in the developing cortex. The Pax6-positive VZ area gradually narrowed during development.
Ijms 23 06115 g003
Figure 6. Expression patterns of Pax6 in the mouse brain and the regional volume decrease in the Pax6 heterozygous mutant (rSey2/+) rat brain. (A) Expression patterns of Pax6 in the adult mouse brain have been redrawn from [25,88,160,161]. The nomenclature and subdivided brain regions are based on previous literature [162]. (B,C) The sex differences in regional volume decrease in the brain of the rSey2/+ rat compared to the WT using a deformation-based morphometry analysis of MRI data [160]. Pink and blue represent the clusters of regional volume decreases in the brain of female (B) and male (C) rSey2/+ rats compared to WT rats, respectively. The pink region is larger than the blue region. The graphical diagram has been redrawn from [160]. Abbreviations; AMG: amygdala, DG: dentate gyrus, HPC: hippocampus, ICx: isocortex, PrC: precommissural nucleus, SVZ: subventricular zone.
Figure 6. Expression patterns of Pax6 in the mouse brain and the regional volume decrease in the Pax6 heterozygous mutant (rSey2/+) rat brain. (A) Expression patterns of Pax6 in the adult mouse brain have been redrawn from [25,88,160,161]. The nomenclature and subdivided brain regions are based on previous literature [162]. (B,C) The sex differences in regional volume decrease in the brain of the rSey2/+ rat compared to the WT using a deformation-based morphometry analysis of MRI data [160]. Pink and blue represent the clusters of regional volume decreases in the brain of female (B) and male (C) rSey2/+ rats compared to WT rats, respectively. The pink region is larger than the blue region. The graphical diagram has been redrawn from [160]. Abbreviations; AMG: amygdala, DG: dentate gyrus, HPC: hippocampus, ICx: isocortex, PrC: precommissural nucleus, SVZ: subventricular zone.
Ijms 23 06115 g006
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Ochi, S.; Manabe, S.; Kikkawa, T.; Osumi, N. Thirty Years’ History since the Discovery of Pax6: From Central Nervous System Development to Neurodevelopmental Disorders. Int. J. Mol. Sci. 2022, 23, 6115. https://doi.org/10.3390/ijms23116115

AMA Style

Ochi S, Manabe S, Kikkawa T, Osumi N. Thirty Years’ History since the Discovery of Pax6: From Central Nervous System Development to Neurodevelopmental Disorders. International Journal of Molecular Sciences. 2022; 23(11):6115. https://doi.org/10.3390/ijms23116115

Chicago/Turabian Style

Ochi, Shohei, Shyu Manabe, Takako Kikkawa, and Noriko Osumi. 2022. "Thirty Years’ History since the Discovery of Pax6: From Central Nervous System Development to Neurodevelopmental Disorders" International Journal of Molecular Sciences 23, no. 11: 6115. https://doi.org/10.3390/ijms23116115

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop