Next Article in Journal
The Role of Sperm Membrane Potential and Ion Channels in Regulating Sperm Function
Next Article in Special Issue
Microcephaly Gene Mcph1 Deficiency Induces p19ARF-Dependent Cell Cycle Arrest and Senescence
Previous Article in Journal
Assessment of Interleukin-15 (IL-15) Concentration in Children with Idiopathic Nephrotic Syndrome
Previous Article in Special Issue
The Dual Roles of Triiodothyronine in Regulating the Morphology of Hair Cells and Supporting Cells during Critical Periods of Mouse Cochlear Development
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Early Steps towards Hearing: Placodes and Sensory Development

1
LBN, Laboratory of Bioengineering and Nanoscience, University of Montpellier, 34193 Montpellier, France
2
Department of Biology, CLAS, University of Iowa, Iowa City, IA 52242, USA
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2023, 24(8), 6994; https://doi.org/10.3390/ijms24086994
Submission received: 9 March 2023 / Revised: 3 April 2023 / Accepted: 6 April 2023 / Published: 10 April 2023
(This article belongs to the Special Issue Molecular Research in Neurodevelopmental Disorders)

Abstract

:
Sensorineural hearing loss is the most prevalent sensory deficit in humans. Most cases of hearing loss are due to the degeneration of key structures of the sensory pathway in the cochlea, such as the sensory hair cells, the primary auditory neurons, and their synaptic connection to the hair cells. Different cell-based strategies to replace damaged inner ear neurosensory tissue aiming at the restoration of regeneration or functional recovery are currently the subject of intensive research. Most of these cell-based treatment approaches require experimental in vitro models that rely on a fine understanding of the earliest morphogenetic steps that underlie the in vivo development of the inner ear since its initial induction from a common otic–epibranchial territory. This knowledge will be applied to various proposed experimental cell replacement strategies to either address the feasibility or identify novel therapeutic options for sensorineural hearing loss. In this review, we describe how ear and epibranchial placode development can be recapitulated by focusing on the cellular transformations that occur as the inner ear is converted from a thickening of the surface ectoderm next to the hindbrain known as the otic placode to an otocyst embedded in the head mesenchyme. Finally, we will highlight otic and epibranchial placode development and morphogenetic events towards progenitors of the inner ear and their neurosensory cell derivatives.

1. Introduction

Ear and epibranchial placodes are originating from the ectoderm to generate neurons and sensory cells, including the underlying gene regulatory networks, and signaling pathways to make sensory neurons, hair cells and taste buds. Much early development has been characterized in animal models, mostly in mice, but it is incompletey described for humans, including CHARGE [1,2] and BOR [3,4,5] syndrome defects.
Hearing starts from the placode that will be transformed into the otocyst or otic placode. The prosensory domain of the otic placode gives rise to the vestibular system as well as the cochlea and the spiral ganglion neurons (SGNs) that will connect with the cochlear hair cells [6,7]. Once the connection between the SGNs and the cochlear nuclei is established, the central auditory system develops to reach out the inferior colliculi and the auditory cortex [8,9]. In contrast to the gain of hair cells, SGN, cochlear nuclei and the auditory system, we also have a decline of sensory and neuron losses that will reduce the sound input [10,11]. Hearing deficits will likely affect close to one billion people [12] who are living with Alzheimer’s disease [13,14,15]. Likewise, the associated epibranchial placode that gives rise to distinct neuron development [16,17] can lose the sensory cells of taste in aging people [18,19,20]. Human placode development and restoration has remained a serious medical condition caused by the dysfunction of placode-derived tissues. The formation of new hair cells, vestibular ganglion neurons (VGN), SGNs, epibranchial neurons, taste buds, brainstem and cortex are a common problem in humans for which we need a solution to generate new sensory receptors and neurons.
In the context of the inner ear, cell-based strategies to replace damaged neurosensory cells by implantation of different cell types into the inner ear represent a challenge which is in continuous progress [21,22]. Transplantation of stem cell-derived otic neuronal and epithelial progenitor cells into the cochlear nerves or into the cochleae has been demonstrated in animal models of SNHL. Successful engraftment and integration have been observed into the target sites, although with a lower survival rate, and displayed molecular features of early neurosensory differentiation [23,24,25,26].
The results from these studies are a proof-of-principle, that transplantation of partially differentiated otic progenitors may be a useful for cell-based cell therapy therapeutic strategy to treat SNHL.
Despite significant progress in inner ear transplantation, methodologies will need to be refined to generate a homogenous population of differentiated otic progenitor cells in 2D and 3D culture systems and their optimal molecular characterization before in vivo implantation.
Stem cell-derived inner ear organoids are also in continuous progress for the generation of neurosensory-like cells. They harbor fairly ordered tissues closer to those in the in vivo developing ear than can be achieved in the 2D monolayer culture system, allowing for increased recapitulation of some developmental events. In these stem cell-derived 3D cultures, otic vesicles form autonomously within the aggregates, following otic–epibranchial progenitor domain formation [27,28]. In a similar manner, initial cell culture studies with isolated otic placodes have shown that once induction and invagination is complete, sensory cell and neuronal differentiation is autonomous [29]. However, in the current otic differentiation models, major limitations of the 3D organoid cultures are related to the variable reproducibility of the system [30,31], and the requirements to optimize protocols specific for a given pluripotent stem cell line [27,32] have substantially limited the studies using 3D inner ear organoids.
Therefore, a deep comprehension of the earliest requirements for pre-placodal ectoderm, epibranchial placode and sensory organ development is of utmost clarification to refine human 3D otic cell differentiation protocols that will contribute to the next generation of cell-based therapeutic approaches to restore sensorineural hearing loss.
In this review, we describe how the neural plate forms to generate the ear and epibranchial placode. We will also highlight the otic and epibranchial placode development towards progenitor cells and their neurosensory cell derivatives including hair cells and taste buds.

2. Early Ear Placodal Development: From Unspecified Cells to Neural Plate and Pre-Placodal Ectoderm

With the earliest transformation from a common ectoderm into two major placodes, the otic and the epibranchial placodes share for the initial formation, but it also has a slightly different gene signaling cascade. Neural induction involves genes known to be important for the earliest steps to general body patterning that will generate the neural plate [33]. Initially, BMP, Wnt, Activin and Nodal are expressed throughout the ectoderm [34,35]. The transition from ectodermal tissue to neural ectoderm requires the downregulation of BMP by Noggin, Chordin and Folistatin (Figure 1) and the downregulation of Wnt by Dkk and Cerberus [36]. Fgfs also contribute to this process [34].
Following initial neural induction, at least three genes, Geminin (Gmnn), Zic and Foxd4 (in frogs), are expressed that further define and stabilize the neural ectoderm by inhibiting BMP and Wnt signaling as well as upregulating other pro-neural genes [34,35,37,38] (Figure 1). Overexpression of Gmnn leads to expression of Zic1 and, while there is no direct interaction of these two genes, they were shown to have overlapping and cooperative roles in neuronal progenitor formation and maintenance [37]. Similarly, Foxd4 promotes the proliferation of neuronal precursors [35] and regulates expression of the Gmnn and Zic genes [34,39].
The neural plate is induced following the inhibition of the epidermally expressed BMPs and Wnts along with the upregulation of the Gmnn, Foxd4 and Zic genes [34,37,40]. The expression of Sox genes results in the transformation of these precursor cells into neural plate stem cells [34]. The SoxB1 family (Sox1, Sox2, Sox3), together with Gmnn, Foxd4 and Zic2, are essential for the continued proliferation of undifferentiated neural stem cells [34]. The explosive proliferation of these neural stem cells and the coordination with convergent–extension leads to the folding of the neural plate into the neural tube (Figure 1; Refs. [40,41]). The following transition from a neural stem cell into a neuronal progenitor cell requires the expression of Sox11 [34]. For a neuronal progenitor cell to exit the cell cycle and proceed to differentiate, the initial upregulation of Gmnn, Foxd4, Zics and Sox genes must be downregulated [37,38,42,43]. Following the downregulation of the various proliferative genes, bHLH genes involved in neuronal differentiation in the CNS become expressed in dorso-ventral expression patterns within the developing neural tube to regulate fates of neurons within various subdomains [44,45].
Figure 1. Comparison of the critical steps identified in distinct steps of neural induction of pre-placode. Initially, Nodal, Activin, BMP and Wnt are expressed throughout the ectoderm. Gmnn, Zic and Foxd4 become expressed, upstream from Irx, and upregulate Sox expression. An early expression of Shh drives Gli expression ventral and is counteracted by the roof plate/choroid plexus. Lmx1a/b drives the dorsal expression of Wnts and BMPs, including Gdf7. Additional downstream genes of Fgf3/10 and Foxi3 are needed for the earliest pre-placode definition. Downstream and largely independent include genes such as Eya1 and Pax2/8, among others. Modified after [37,38,44,46,47,48,49].
Figure 1. Comparison of the critical steps identified in distinct steps of neural induction of pre-placode. Initially, Nodal, Activin, BMP and Wnt are expressed throughout the ectoderm. Gmnn, Zic and Foxd4 become expressed, upstream from Irx, and upregulate Sox expression. An early expression of Shh drives Gli expression ventral and is counteracted by the roof plate/choroid plexus. Lmx1a/b drives the dorsal expression of Wnts and BMPs, including Gdf7. Additional downstream genes of Fgf3/10 and Foxi3 are needed for the earliest pre-placode definition. Downstream and largely independent include genes such as Eya1 and Pax2/8, among others. Modified after [37,38,44,46,47,48,49].
Ijms 24 06994 g001
After the neural plate fuses to form the neural tube (Figure 1), the newly formed dorsal roof plate begins to express BMPs and Wnts while the ventral floorplate and underlying notochord expresses Shh [50,51]. Although many genes play a role in patterning the neural tube, BMPs, Wnts and Shh play a primary role [50]. Expression of Wnts and BMPs in the dorsal neural tube is driven by the dorsal expression of Lmx1a/b [44,46,52]. Lmx1a/b expression itself is in the roof plate and is regulated by Zic1, Zic3 and Zic4 [53]. While the timing of Wnt and BMP expression during neural tube formation varies between mice and frogs [38,54], these two genes, as well as Fgfs and retinoids, are critical in defining the dorsal part of the brainstem and spinal cord [55]. Loss of Wnt or BMP signaling negatively affects dorsal progenitor cells [46].
Determination of dorso-ventral identity within the brainstem is governed by the expression of various homeodomain and bHLH transcription factors in restricted areas along the dorso-ventral axis [45,55]. The dorsal brainstem is subdivided into eight domains of neuronal progenitor populations [55]. The bHLH gene, Atoh1, is expressed in the dorsal-most domain (dA1) throughout the brainstem and spinal cord [55,56]. Atoh1-positive cells generate the rhombic lip of the hindbrain, superficial migratory neuron streams and cerebellar granule cells, that contribute to auditory, vestibular, solitary tract and proprioceptive networks [57]. As expected from the expression pattern, loss of Atoh1 eliminates most cochlear nuclei neurons [58]. Additional genes are expressed in the Olig3 domain including Neurog1/2 (dA2), Ascl1 (dA3-dB1) and Ptf1a (dA4-dB1) [45,59]. For example, the dorsal-most subdomain (dA1), that also expresses Atoh1, expresses Pax3, the next subdomain (dA2) expresses Pax3 and Pax7 and the third subdomain (dA3) expresses Pax3, Pax6 and Pax7 [55]. Additional genes are uniquely expressed in progenitor subdomains, such as Barhl1 in dA1 progenitors, Lhx in dA2, Tbx in dA3, Foxd3/Foxp2 in dA2 and dA4 and Phox2b in dB2 [55]. Genes, such as Pou4f1, are expressed in multiple subdomains (dA1-dA3, dB3) [55]. Interestingly, Atoh1, which defines the dorsal-most progenitor population (dA1), is also expressed in a more ventral subpopulation (dB2) during their maturation [45,55,60]. Phox2b is expressed in rhombomere 2–6 of the hindbrain (dB2), but not in rhombomere 7 or the spinal cord [55]. This unique hindbrain domain (dB2) also later expresses Atoh1. A second unique domain in the hindbrain (dA4) expresses Ptf1a, Foxd3 and FoxP2, among other genes [55]. Loss of Ptf1a results in the loss of dA4 and dB1 neurons in combination with the expansion of dA3 and dB3 neurons, leading to the eventual misspecification of somatosensory and viscerosensory nuclei neurons in the hindbrain [61,62]. A combination of bHLH genes Ascl1, Neurog2 and Olig3 defines dA3, in the brainstem, which extends through the spinal cord (rostral to caudal) at a location which includes the developing solitary tract [55]. In situ hybridization showed co-expression of Rnx (Tlx3) and Phox2b in the neurons of the solitary tract; in the absence of Neurog1, non-taste ganglia are lost. Projection neurons of nST are entirely absent in both Tlx3 (Rnx) and Phox2b knockouts, indicating that these neurons are dependent on these factors for their development [63].
Transition from neuronal progenitor to neuronal differentiation requires the interaction of pro-neural bHLH genes and additional bHLH genes, the Class I Hes/Hey genes [42] and the Class V ID genes [64,65,66]. Proliferation of neurosensory precursors is driven by Hes, ID, Sox2 and Myc genes, and the transition to differentiated cells is controlled by the balance of Notch signaling molecules and the pro-neural bHLH genes. Oscillation of a pro-neural bHLH gene and a repressive one, such as Ascl1 and Hes1, through cross regulation of each other controls the timing of neurogenesis [42]. This oscillation maintains proliferation of neuronal progenitors, whereas the eventual loss of Hes1 expression and subsequent sustained expression of a pro-neural bHLH gene leads to neuronal differentiation [42]. The expression of downstream bHLH genes, such as Neurod1, adds to these complex interactions. For instance, Atoh1 expression extends along the roof plate to the cerebellum, parallel to the slightly more ventral expression of Neurog1/2, and the loss of dorsal neurons in Atoh1 null mice [57,58] results in a reduced cerebellum and auditory nuclei [52]. Neurod1 negatively regulates Atoh1 expression during cerebellum, and gut proliferation and manipulating Neurod1 expression that may help to counteract Medulloblastoma [67,68,69,70]. Atoh1 shows a much higher level of expression in the auditory nuclei and counteracts with Neurod1, indicating a differential regulation of expression in auditory nuclei [71].
In summary, an overview is provided of the neural placode that develops to generate the pre-placodal induction (Figure 1) and develops the hindbrain region that defines the different longitudinal expression of bHLH and other gene expression.

3. Patterning of the Pre-Placode Epithelium and Restriction of Pan-Placodal Development

Once the neural plate is fused and generates the roof plate, neural crest and placodes develop from immediately adjacent ectodermal tissue [33,72]. Neural crest and placodal neurons give rise to the formation of all or part of the various sensory systems. The work of Northcutt and Gans provided a novel perspective of the organization of the neural crest and placodes [73]. Development of neural crest and placodes is highly conserved across vertebrate species [33,74].
Cranial placodes (Figure 1 and Figure 2) develop from a pan-placodal region at the neural border zone and give rise to the anterior pituitary gland, the olfactory epithelium, the lens of the eye and the large neurons in the trigeminal, facial, glossopharyngeal (petrosal) and vagus (nodose) nerves. This is in addition to the inner ear that contains the vestibular and auditory epithelia and sensory neurons [33,74]. The pan-placodal region broadly expresses Eya1, Six1 and Six4 [4,75]. During development, the pan-placodal region begins to subdivide in the trigeminal placode, the ear placode and three epibranchial placodes [16,76]. Differential expression of transcription factors along the anteroposterior axis gives rise to anterior and posterior compartments, which are further subdivided into individual placodes [77]. The posterior compartment gives rise to the otic and epibranchial placodes which expresses Pax2, Pax8, Sox2 and Sox3 [78].
While the vestibular system is highly conserved across vertebrates, the auditory system is much more variable [80,81]. Both sensory systems are housed within the inner ear. The otic placode uniquely depends on Foxi3 [48] and Fgf3/10 [49], but shares the requirement of Eya1/2/Six1/2 with many other sensory placodes [82,83]. Additionally, Foxg1 [84], Lmx1a/b [52], Pax2/8 [85], Fgf8 and Fgfr2b [86,87,88,89,90], Gbx2 [91], Gata3 [92,93] and Shh [94] are required for inner ear development (Figure 2). For instance, the ear never develops beyond the otocyst stage with the loss of both Pax2 and Pax8 [85] or Eya1 deletion [4]. During development, the inner ear is patterned anteroposteriorly, dorsoventrally and proximal-distally [95]. Shh specifies the ventral portion of the ear, and in its absence, ventral structures, such as the cochlea, are absent [94,96,97]. In contrast, BMP4 specifies dorsal vestibular structures [98,99]. Other genes, such as Foxg1 [84,100], Lmx1a/b [101,102], Otx1/2 [91,103] and n-Myc [104,105], affect specific sensory epithelia and/or structures of the inner ear. Notch signaling and retinoids are also involved in inner ear patterning [42,106,107].
Inner ear neurogenesis requires the induction of Eya1 by Sox2, which in turn downregulates Sox2 expression by Neurog1 [108] that provides a negative feedback. Neurog1 [109,110] upregulates Neurod1 [8,111,112] and several other bHLH genes [113,114]. Pou4f1 is also involved in neuronal development and in addition, for the proper pathfinding of inner ear afferent neurons [115,116]. Additional genes, such as Npr1, Prickle1, Fzd3/6, Ephrin’s and Vangl2, have been implicated in inner ear afferent central and/or peripheral pathfinding [117,118,119,120,121,122]. Neural crest induction occurs at the lateral edge of the neural plate [123]. Following closure of the neural tube, neural crest cells migrate throughout the embryo, contributing to a wide range of tissues including neurons, craniofacial skeleton, smooth muscles and melanocytes [72,124]. BMP, Wnt, Fgf and Notch signaling are all essential for the formation of neural crest cells [124]. Upon induction of the neural crest, many crest-specific genes are expressed, including Sox5, Sox8, Sox9 and Sox10 [72,123]. While there are specific species differences in the onset and sequence of expression, these Sox genes are important for the specification, migration and differentiation of neural crest cells [123,125]. Additional genes that are upregulated include Snail, Slug, Pax3/7, Hairy2, Msx1/2, Dlx5 and Gbx2 [72,124].
In summary, the earliest induction of distinct derivatives for the formation is presented of pre-placodal to pan-placodal development.

4. Development of Distinct Placodal Derivatives of the Otic Placode and the Epibranchial Placode

A unique feature is forming by sinking in the placode to generate an otic cup (Figure 2; about E9 in mice, d24 in humans) that will pinch off to form an otic cyst (about E9.5 in mice, d26 in humans). At E10 in mice (d28 in humans), the otocyst is separate from the overlaying ectoderm. The next step is to develop an endolymphatic duct that elongates. Caveating of the otocyst will develop the three semicircular canals, the utricle and saccule and elongates to become the cochlear duct (Figure 3, Ref. [105]). In contrast, only neurons are generated from epibranchial placodes [16,126]. Specifically, the facial (VII), glossopharyngeal (IX) and vagal (X) neurons develop from placodal and neural crest cells [127]. These neurons project centrally to specific brainstem nuclei (solitary tract, VII, IX, X; Figure 3) and peripherally to sensory cells (taste buds).
Induction of the epibranchial placodes depend critically on the expression of Eya1/2/Six1/4. These genes are upstream of Pax2, which help define the caudal (Pax2/8, VII, IX, X) placodes [75,126]. Additional genes are important in the development of the placodes contributing to the different cranial ganglia. For instance, Neurog1 is necessary for trigeminal sensory ganglia [129], whereas Neurog2 is necessary for sensory ganglia that develop from epibranchial placodes (VII, IX, X) [130]. Downstream of Neurog1/2 are Neurod1, Isl1 and Pou4f1 genes, which promote the differentiation of placodally derived sensory neurons [111,115,129,130,131]. In addition, the epibranchial placodally derived neurons (VII, IX, X) depend upon Phox2b expression [17], which is downstream of Eya1/Six1 [82]. Additional genes such as Notch, Hes, Rbpj, Fgf8, PDGF and Wnts are important for proliferation and differentiation of neurons [132].
In summary, we provide an overview of the placodal development that will incite future interactions to make different sensory neurons of the ear and epibranchial neurons first followed by the hair cells and taste buds that are forming later.

5. Hair Cells of the Inner Ear Have a Shared Developmental Program of Neurons

Vestibular hair cells generate at least two types, type I and type II (Figure 4 [76,133,134]). A unique development is defined for the mammalian cochlear hair cells, the inner hair cells (IHCs) and outer hair cells (OHCs). Upstream is the expression of Eya1 and Brg1 that are needed for the initiation of hair cell progenitors. Downstream is the expression of Sox2 to interact with specific bHLH genes to initiate cochlear hair cells [108]. The bHLH gene Atoh1 (Figure 4) is needed for vestibular and cochlear hair cell formation beyond undifferentiated progenitor hair cells [6,135]. Downstream are Pou4f3, Gfi1 and Barhl1 gene expression that are needed for hair cell maintenance [136].
An interaction is existing between the bHLH gene Neurog1 and Foxg1 that causes a reduced expression of Atoh1: The absence of Neurog1 or Foxg1 results into a short and wider cochlear set of hair cells that leads to more rows of OHCs (instead of three rows of OHCs; [84]). A misexpression of Neurog1 instead of Atoh1 results in IHCs and OHCs that are not functional and are converted as inner pillar cells into gaps between near normal IHCs [137]. An interaction also shows a shorter cochlea and converts OHCs into IHCs after deletion of Neurod1, a bHLH gene. Proliferation of more hair cells depend on n-Myc for normal differentiation [105]. Most recently, an effect of follistatin (Fst) is found in the apex [138]. Combined, the cochlear apex and base are differentially affected by Lmx1a, n-Myc, Fst, Neurog1 and Foxg1, among others.
Downstream of OHCs depend on Insm1 and Ikzf2 that are needed for OHC development; OHC development interacts with Neurod1 that also regulates OHCs instead converted to IHC [139]. In contrast, loss of Tbx2 converts IHC into OHC-like hair cells [140]. Furthermore, IHC depends on Fgf8 and Srrm3/4, and cochlea will lose nearly all IHCs in the absence of these two selectively expressed genes (Figure 4; Ref. [141]).
Figure 4. Hair cells and taste buds are independently derived. Progenitors of hair cells depend on Eya1, Sox2 and Atoh1 that are needed for their development. Downstream are Pou4f3 and Gfi1 that are needed to maintain hair cells. Tbx1,2,3 interacts with Neurog1, and Foxg1 regulates the number and distribution of cochlear to make it shorter and has increased the number of hair cells. A loss of all cochlear hair cells is revealed in Pax2, Gata3 and Lmx1a/b mice. Two types of vestibular hair cells exist that have a mixed distribution of type I and type II HCs. In contrast in the cochlea, we have a single row of IHCs and three rows of OHCs. Tbx2, Srrm3/4 and Fgf’s are needed for differentiation and viability of IHC. Insm1, Ikzf2 and Fgf20 are needed to differentiate OHCs or requires for forming three rows of OHCs. A very different sequence of genes is needed in taste buds. Starting is the upregulation Krt8/14 that is prior to Shh. Downstream is Sox2 that is required for taste bud differentiation. Overlapping are other genes some of which seem to differentiate into distinct taste sensory input. Compile with permission from Refs. [4,8,16,52,84,110,140,141,142,143,144,145].
Figure 4. Hair cells and taste buds are independently derived. Progenitors of hair cells depend on Eya1, Sox2 and Atoh1 that are needed for their development. Downstream are Pou4f3 and Gfi1 that are needed to maintain hair cells. Tbx1,2,3 interacts with Neurog1, and Foxg1 regulates the number and distribution of cochlear to make it shorter and has increased the number of hair cells. A loss of all cochlear hair cells is revealed in Pax2, Gata3 and Lmx1a/b mice. Two types of vestibular hair cells exist that have a mixed distribution of type I and type II HCs. In contrast in the cochlea, we have a single row of IHCs and three rows of OHCs. Tbx2, Srrm3/4 and Fgf’s are needed for differentiation and viability of IHC. Insm1, Ikzf2 and Fgf20 are needed to differentiate OHCs or requires for forming three rows of OHCs. A very different sequence of genes is needed in taste buds. Starting is the upregulation Krt8/14 that is prior to Shh. Downstream is Sox2 that is required for taste bud differentiation. Overlapping are other genes some of which seem to differentiate into distinct taste sensory input. Compile with permission from Refs. [4,8,16,52,84,110,140,141,142,143,144,145].
Ijms 24 06994 g004
Several genes are needed for the cochlear development which results in the absence of all cochlear hair cells. In the absence of Gata3, Lmx1a/b and Pax2 (Figure 4), the cochlea do not develop any hair cell that likely interacts with Atoh1 for hair cell formation [6,52,85,93]. In addition, we have a delayed loss of all hair cells that initially develop after a loss of Bdnf/Ntf3 double deletions [146]. Moreover, Bdnf deletion will result in the loss of apical OHCs, while the loss of Ntf3 causes the loss of OHCs in the basal cochlear turn. Likewise, it is dependent and will degenerate in Cdc42 for IHCs first, followed by OHCs in a base to apex progression [147,148,149]. In addition, MANF deletion results in the loss of OHCs in the base of the cochlea [150].
There is a delay between SGNs and cochlear HCs: SGNs form first in the base (E10.5) to apex (E12.5). In contrast, cochlear HCs start to from in the apex around E12.5 and progresses to the base on E14.5 [110]. Atoh1 is needed for all HCs but has a different progression starting in the base at E14.5 and progresses to the apex around E18.5 [135]. Innervation is reaching IHCs starting at E15 that expand to reach OHC innervation by Type II fibers at E18 (Figure 4; Ref. [151]).
In summary, at least two types of cochlear HCs develop that have a unique input–output connection [7,152] that are likely unrelated to the types I and II of vestibular hair cells [6,134].

6. A Unique Set of Genes Is Needed for Taste Bud Development

Like the taste ganglia, taste bud development is orchestrated by a specific sequence of gene expression and trophic interactions with nerve fibers. In the tongue, taste buds are in the papillae, which develop from placodes that arise prior to innervation and taste bud formation [153,154]. The initial signals that orchestrate taste bud development and establish their patterns on the tongue arise from the tongue epithelium [154,155]. Interestingly, signal patterning of the location of taste buds differs depending on the tongue region (i.e., fungiform papillae vs. circumvallate papillae), likely because these tissues arise from different branchial arches. Before any other gene is expressed, the earliest expression is Krt8 followed by Krt14 [143]. Specifically, fungiform papillae (Figure 4) are patterned during development by sonic hedgehog (Shh) and Wnt signaling pathways [156,157,158,159], whereas the circumvallate papilla is regulated by fibroblast growth factor 10 (FGF10) and its receptors Spry1-2 [88,160]. The cells within the developing taste epithelial placode that express Shh differentiate into taste buds during development [161]. Prior to differentiation, these Shh+ placodal cells become innervated, and this innervation is required to maintain Sox2 expression [162]. Both Sox2 expression and innervation are required for continued taste bud development [145,163]. A differentiation requires Lgr4, Six1, Hes6 and Foxa2 expression [143,164], among others. The loss of taste bud innervation following the knockout of either the neurotrophin, BDNF, or its receptor, TrkB, results in a loss of taste buds over time [165,166,167]. The factors that are produced by the neurons to support continued taste bud development are unclear; however, Shh and R-spondin are likely possibilities [168,169,170]. Although the taste system matures, some developmental processes continue into adulthood, including taste bud cell differentiation. One unique feature of taste buds is that cells are replaced every 8–20 days, depending on the specific taste bud types [171,172,173]. The stem cells that give rise to adult taste buds express Lgr6 in the fungiform papillae, but express Lgr5 within the circumvallate papillae [174,175]. Lgr5/6 cells give rise to all taste bud cell types [175]. The absence of Neurog2 leads to reduced Sox2 expression [145,176] and disrupts taste bud formation beyond a limited differentiation of small, single taste buds [163]. Taste bud cells that transduce bitter express Eya1 suggesting that it may be involved in the differentiation of this cell type [144]. Similarly, the differentiation of type II cells is dependent on the transcription factor, Pou2f3, while type III cells depend on Ascl1 for their differentiation [16,144,177,178,179].
In summary, the development of taste buds is controlled by a series of gene expression events. Once the axons of peripheral neurons reach the taste epithelium, these two cell types become interdependent. Thus, the formation of the peripheral taste system represents the interaction between early and late genes that regulate cell fate and the trophic interactions that occur between taste buds and nerve fibers.

7. Development of Neuronal Genes Are Needed for Ears and Epibranchial Placodes

A set of unique genes is characterized so that it requires the vestibular, spiral and epibranchial ganglion neurons [6,16].
Vestibular ganglion neurons (VGNs) and spiral ganglion neurons (SGNs) are independently derived from a common origin, Eya1, Sox2 and Neurog1 (Figure 5). Eya1 and Sox2 define SGN precursor populations, whereas Neurog1 is needed to initiate the proliferation and differentiation of SGNs [4]. Without Eya1, Sox2 or Neurog1, no sensory neurons would develop [4,108]. Downstream there is a segregation of genes that are Tlx3-positive for vestibular neurons but are negative for SGN [142]. Furthermore, Sall3 is in part positive for certain VGNs and negative for other VGNs [180].
The diversity of two types of SGNs has been defined: Neurod1, bHLHe22, Pou4f1, Runx1, Prox1 and Gata3 determines the type Ic (Figure 5). The expressions of Tshz2, Grhl1, Rxrg, Tie4 and Id1 are needed for type Ia (Runx1), Ib (Gata3) and type II (Etv4, Prrx2) SGNs [116]. There is the upregulation of Pou4f1, Lypd1 and Mgat4c (Type 1c), Lypd1, Pou4f1 and Calb1/2 (Type Ib), Calb1, Pcdg20 and Prph1 (Type Ia) and Plk5, Prph1 and Th [Type II; [116,151]]. In addition, the Nhlh1/2 and Isl1 are interacting with Neurod1 to regulate SGNs [181]. VGN formation [180] requires at least three populations that will develop into the bouton terminal, the calyx terminals, and the mixed terminals (calyx and boutons; [182]).
Several selective genes can selectively induce losses of distinct SGNs. Conditional deletion of Gata3, Lmx1a/b, Dicer, Shh or Pax2 results in the complete loss of SGNs (Figure 5), while many VGNs develop despite the loss of SGNs [52,85,93,183,184,185]. Furthermore, the cochlea is reduced into a sac without forming a cochlear duct in several mutants [52,85,93].
Figure 5. Progenitors of mammals are dependent by initial Eya1, Sox2 and Neurog1 that are needed for several genes to differentiate neurons of either the otocyst or the epibranchial placodes. In the ear, downstream are a large set of genes, among them are Neurod1, Isl1 and Pou4f1. A split is dependent on Tlx3 that forms VGNs, whereas SGNs are not dependent on Tlx3. After that, the upregulation of Sall3 is needed to make two different kinds of VGNs but has a third population that is unclear as to its origin. SGNs split into four populations: Ia, Ib, Ic and II. A unique set of genes and their expression needs to be verified to develop each type of SGN, notably Pax2, Gata3 and Lmx1a/b. For the epibranchial neurons, unique common genes are deviate in the epibranchial neurons by switching the Neurog1 into Neurog2 and have early expressions of Phox2b, Isl1, Neurod1 and Foxg1. Image is compiled with permission from Refs. [4,17,108,116,142,180,186].
Figure 5. Progenitors of mammals are dependent by initial Eya1, Sox2 and Neurog1 that are needed for several genes to differentiate neurons of either the otocyst or the epibranchial placodes. In the ear, downstream are a large set of genes, among them are Neurod1, Isl1 and Pou4f1. A split is dependent on Tlx3 that forms VGNs, whereas SGNs are not dependent on Tlx3. After that, the upregulation of Sall3 is needed to make two different kinds of VGNs but has a third population that is unclear as to its origin. SGNs split into four populations: Ia, Ib, Ic and II. A unique set of genes and their expression needs to be verified to develop each type of SGN, notably Pax2, Gata3 and Lmx1a/b. For the epibranchial neurons, unique common genes are deviate in the epibranchial neurons by switching the Neurog1 into Neurog2 and have early expressions of Phox2b, Isl1, Neurod1 and Foxg1. Image is compiled with permission from Refs. [4,17,108,116,142,180,186].
Ijms 24 06994 g005
A unique set of Tbx1, 2 and 3 deletions interact with Neurog1 to downturn VGNs and SGNs [142]. Tbx1 acts as a selector gene which controls neuronal fate in the otocyst. Ablation of Tbx2 leads to cochlear reduction, but it is unclear how many SGNs form. Likewise, absence of Tbx3 results in vestibular malformations that require the distribution of VGNs. Combined deletion of both Tbx2/3DKO results in incomplete and largely absent cochlear and vestibular structures in newborn mice for which we have no differentiation of VGNs and SGNs.
Downstream are neurotrophins that are needed for sensory neuron development and maturation [187]. The loss of all SGNs is documented for TrkB and TrkC that signal for Bdnf (TrkB) and Ntf3 (TrkC). There is an incomplete deletion in the basal cochlear turn that is dependent on Ntf3 and TrkC, while the loss of Bdnf and TrkB resulted in the reduction and loss of the apex SGNs: Bdnf depends on 95% of VGNs while only about 5% are lost in the SGN, mainly in the apex. In contrast, Ntf3 (NT3) is nearly lost of 95% of SGNs, has lost all basal turn in SGNs, but has only an additional loss of about 5% of VGNs [188].
Proliferation starts from E9–14 for VGNs and from E10.5–12.5 for SGNs. The VGNs delaminate and migrate to form a ganglion outside of the ear while the SGNs stay inside the cochlea. Developing neurons reach out first from VGNs at E10; whereas, the first SGNs fibers reach the auditory nuclei at E12.5. Deletion of certain genes [Neurod1, Isl1; Refs. [112,181]] as well as loss of Schwann cells after Sox10 or ErbB2 deletion show a migration of SGNs that mixed with the VGNs [125].
In summary, a set of genes is needed for the formation of the SGNs and depends on their development which diversifies into four SGN types.
Like we detailed for the origin of the otocyst, gustatory ganglion formation depends on the sequential expression of specific genes, resulting in epibranchial placode formation, delamination, migration and cellular differentiation [16]. Initial placode development depends on the expression of the transcription factor Foxi3, which is essential for the development of the ear and epibranchial placodes (Figure 5). In the absence of Foxi3 expression, the ectoderm fails to thicken, and all placodes fail to form [48]. Six1/2/4 and Eya1/2 are essential for the specific formation of epibranchial placodes [33,72,126] downstream of Foxi3. The Sox2, which is ubiquitously expressed in the epibranchial placodes, is required for neuron development [108]. Epibranchial placodes are patterned into rostral and caudal domains by Notch signaling [189], which is regulated by Eya1 [190]. From the Sox2+ precursor pool, placode cells divide into a non-neural population and neuroblasts defined by the transcription factor Neurog2 in mammals [190]. Neurog2 is critical for neuronal development [163,191], and Neurog1 plays a similar role in chickens [128,192]. The activation of Notch signaling discourages neuronal fate [190] in favor of a non-neuronal cell fate. Downstream of Neurod1, Isl1, Pou4f1 and Phox2b interact to regulate neuronal migration and differentiation [128,163,193]. Although all neuroblasts express the pro-neural transcription factor Isl1, Phox2b is required specifically for a visceral sensory neuron (taste) fate [194]. Foxg1, Phox2a and Phox2b are expressed in the placodes during delamination [128], followed by the set of genes Coe1, Drg11 and Dcx, which are only activated after the migrating cells have left the placode [128,192]. Once the final position is reached (facial, glossopharyngeal, vagus ganglion), the neurons grow a single process that branches to innervate the targeted taste bud cells and sends the proximal innervation to reach distinct areas of the hindbrain to reach out the solitary tract [16,127,195].
In summary, we provide an overview of otic and epibranchial neuron formation from the earliest expression of precursor cells to the differentiation of vestibular, spiral and epibranchial neurons.

8. Concluding Remarks

A detailed characterization of the human embryonic and fetal inner ear development is important. Specifically, induction and the early morphogenesis of the inner ear require signaling cues deployed in both a spatially and temporally restricted pattern. Mimicking these signals under a chemically defined 3D growth system has been shown in many exciting studies to differentiate stem cells into “inner ear organoids” containing sensory epithelia and neurons. However, it should not be concluded that these neurons that differentiated in inner ear organoids only belong to the otic placode lineage. For instance, many of these neurons might resemble cells in the hindbrain or epibranchial placode lineage. Thus, a careful characterization of stem cell-derived neurons in organoid and 3D cell culture systems, principally regarding their specific lineage identities, is still unsatisfactory.
The last several years, developments, although promising for the derivation of human inner ear neurosensory cells from stem cells, emphasize the importance to further improve and build on our deep understanding of inner ear induction and early organogenesis. This knowledge can be effectively exploited to faithfully recapitulate in vitro the crucial developmental steps leading to otic progenitors and their neurosensory derivatives, which can be used to develop therapies for the human inner ear.

Author Contributions

A.Z. initially drafted the paper; B.F. helped with layout of the paper and drafting some of the paper; A.Z. and B.F. finalized the paper editing. All authors have read and agreed to the published version of the manuscript.

Funding

This research has been financially supported by EU-FP7 under the health topic, grant number 603029 and by la Fondation pour l’Audition (Paris) to A.Z. and NIA R01 AG060504 to B.F.

Institutional Review Board Statement

Ethical review and approval were waived for this study due to providing a review, not primary data.

Informed Consent Statement

Written informed consent has been obtained from the patient(s) to publish this paper, if applicable.

Acknowledgments

We sincerely apologize for the articles that could not be referenced due to space limitations. We thank several collaborators, in particular Karen L Elliott, Ebenezer N. Yamoah, and Jen Kersigo.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Blake, K.D.; Prasad, C. CHARGE syndrome. Orphanet J. Rare Dis. 2006, 1, 1–8. [Google Scholar] [CrossRef] [Green Version]
  2. Jongmans, M.; Admiraal, R.; Van Der Donk, K.; Vissers, L.; Baas, A.; Kapusta, L.; van Hagen, J.M.; Donnai, D.; De Ravel, T.; Veltman, J. CHARGE syndrome: The phenotypic spectrum of mutations in the CHD7 gene. J. Med. Genet. 2006, 43, 306–314. [Google Scholar] [CrossRef] [Green Version]
  3. Xu, P.-X.; Adams, J.; Peters, H.; Brown, M.C.; Heaney, S.; Maas, R. Eya1-deficient mice lack ears and kidneys and show abnormal apoptosis of organ primordia. Nat. Genet. 1999, 23, 113–117. [Google Scholar] [CrossRef]
  4. Xu, J.; Li, J.; Zhang, T.; Jiang, H.; Ramakrishnan, A.; Fritzsch, B.; Shen, L.; Xu, P.X. Chromatin remodelers and lineage-specific factors interact to target enhancers to establish proneurosensory fate within otic ectoderm. Proc. Natl. Acad. Sci. USA 2021, 118, 1–12. [Google Scholar] [CrossRef]
  5. Wang, R.Y.; Earl, D.L.; Ruder, R.O.; Graham, J.M., Jr. Syndromic ear anomalies and renal ultrasounds. Pediatrics 2001, 108, e32. [Google Scholar] [CrossRef] [Green Version]
  6. Elliott, K.L.; Pavlínková, G.; Chizhikov, V.V.; Yamoah, E.N.; Fritzsch, B. Development in the Mammalian Auditory System Depends on Transcription Factors. Int. J. Mol. Sci. 2021, 22, 4189. [Google Scholar] [CrossRef]
  7. Duncan, J.S.; Sheltz-Kempf, S.N.; Elliott, K.L. Morphological and Molecular Ontogeny of the Auditory System. In Evolution of Neurosensory Cells and Systems; CRC Press: Boca Raton, FL, USA, 2022; pp. 175–200. [Google Scholar]
  8. Filova, I.; Bohuslavova, R.; Tavakoli, M.; Yamoah, E.N.; Fritzsch, B.; Pavlinkova, G. Early Deletion of Neurod1 Alters Neuronal Lineage Potential and Diminishes Neurogenesis in the Inner Ear. Front. Cell Dev. Biol. 2022, 10, 845461. [Google Scholar] [CrossRef]
  9. Budinger, E.; Kanold, P.O. Auditory cortex circuits. In The Mammalian Auditory Pathways; Springer: Berlin/Heidelberg, Germany, 2018; pp. 199–233. [Google Scholar]
  10. Elliott, K.L.; Fritzsch, B.; Yamoah, E.N.; Zine, A. Age-Related Hearing Loss: Sensory and Neural Etiology and Their Interdependence. Front. Aging Neurosci. 2022, 14, 814528. [Google Scholar] [CrossRef]
  11. Michalski, N.; Petit, C. Central auditory deficits associated with genetic forms of peripheral deafness. Hum. Genet. 2022, 141, 335–345. [Google Scholar] [CrossRef]
  12. Haile, L.M.; Kamenov, K.; Briant, P.S.; Orji, A.U.; Steinmetz, J.D.; Abdoli, A.; Abdollahi, M.; Abu-Gharbieh, E.; Afshin, A.; Ahmed, H. Hearing loss prevalence and years lived with disability, 1990–2019: Findings from the Global Burden of Disease Study 2019. Lancet 2021, 397, 996–1009. [Google Scholar] [CrossRef]
  13. Wang, H.-F.; Zhang, W.; Rolls, E.T.; Li, Y.; Wang, L.; Ma, Y.-H.; Kang, J.; Feng, J.; Yu, J.-T.; Cheng, W. Hearing impairment is associated with cognitive decline, brain atrophy and tau pathology. Ebiomedicine 2022, 86, 104336. [Google Scholar] [CrossRef]
  14. Billig, A.J.; Lad, M.; Sedley, W.; Griffiths, T.D. The hearing hippocampus. Prog. Neurobiol. 2022, 218, 102326. [Google Scholar] [CrossRef]
  15. Uchida, Y.; Nishita, Y.; Otsuka, R.; Sugiura, S.; Sone, M.; Yamasoba, T.; Kato, T.; Iwata, K.; Nakamura, A. Aging brain and hearing: A mini-review. Front. Aging Neurosci. 2022, 13, 991. [Google Scholar] [CrossRef]
  16. Roper, S.D.; Krimm, R.F.; Fritzsch, B. Taste Buds Explained: From Taste Sensing to Taste Processing in the Forebrain. In Evolution of Neurosensory Cells and Systems; CRC Press: Boca Raton, FL, USA, 2022; pp. 111–134. [Google Scholar]
  17. Alsina, B. Mechanisms of cell specification and differentiation in vertebrate cranial sensory systems. Curr. Opin. Cell Biol. 2020, 67, 79–85. [Google Scholar] [CrossRef]
  18. Menni, C.; Valdes, A.M.; Freidin, M.B.; Ganesh, S.; Moustafa, J.S.E.-S.; Visconti, A.; Hysi, P.; Bowyer, R.C.; Mangino, M.; Falchi, M. Loss of smell and taste in combination with other symptoms is a strong predictor of COVID-19 infection. MedRxiv 2020. [Google Scholar]
  19. Schiffman, S.S. Taste and smell losses in normal aging and disease. JAMA 1997, 278, 1357–1362. [Google Scholar] [CrossRef]
  20. Hannum, M.E.; Koch, R.J.; Ramirez, V.A.; Marks, S.S.; Toskala, A.K.; Herriman, R.D.; Lin, C.; Joseph, P.V.; Reed, D.R. Taste loss as a distinct symptom of COVID-19: A systematic review and meta-analysis. Chem. Senses 2022, 47, bjac001. [Google Scholar] [CrossRef]
  21. Zine, A.; Messat, Y.; Fritzsch, B. A human induced pluripotent stem cell-based modular platform to challenge sensorineural hearing loss. Stem Cells 2021, 39, 697–706. [Google Scholar] [CrossRef]
  22. Connolly, K.; Gonzalez-Cordero, A. Modelling inner ear development and disease using pluripotent stem cells—A pathway to new therapeutic strategies. Dis. Model Mech. 2022, 15, dmm049593. [Google Scholar] [CrossRef]
  23. Chen, W.; Jongkamonwiwat, N.; Abbas, L.; Eshtan, S.J.; Johnson, S.L.; Kuhn, S.; Milo, M.; Thurlow, J.K.; Andrews, P.W.; Marcotti, W.; et al. Restoration of auditory evoked responses by human ES-cell-derived otic progenitors. Nature 2012, 490, 278–282. [Google Scholar] [CrossRef] [Green Version]
  24. Takeda, H.; Hosoya, M.; Fujioka, M.; Saegusa, C.; Saeki, T.; Miwa, T.; Okano, H.; Minoda, R. Engraftment of Human Pluripotent Stem Cell-derived Progenitors in the Inner Ear of Prenatal Mice. Sci. Rep. 2018, 8, 1941. [Google Scholar] [CrossRef] [Green Version]
  25. Ishikawa, M.; Ohnishi, H.; Skerleva, D.; Sakamoto, T.; Yamamoto, N.; Hotta, A.; Ito, J.; Nakagawa, T. Transplantation of neurons derived from human iPS cells cultured on collagen matrix into guinea-pig cochleae. J. Tissue Eng. Regen. Med. 2017, 11, 1766–1778. [Google Scholar] [CrossRef] [Green Version]
  26. Flink, E.B. Magnesium deficiency: Causes and effects. Hosp. Pract. 1987, 22, 116a–116i, 116o–116p. [Google Scholar] [CrossRef]
  27. Koehler, K.R.; Nie, J.; Longworth-Mills, E.; Liu, X.P.; Lee, J.; Holt, J.R.; Hashino, E. Generation of inner ear organoids containing functional hair cells from human pluripotent stem cells. Nat. Biotechnol. 2017, 35, 583–589. [Google Scholar] [CrossRef] [Green Version]
  28. Koehler, K.R.; Mikosz, A.M.; Molosh, A.I.; Patel, D.; Hashino, E. Generation of inner ear sensory epithelia from pluripotent stem cells in 3D culture. Nature 2013, 500, 217–221. [Google Scholar] [CrossRef] [Green Version]
  29. Freter, S.; Muta, Y.; Mak, S.S.; Rinkwitz, S.; Ladher, R.K. Progressive restriction of otic fate: The role of FGF and Wnt in resolving inner ear potential. Development 2008, 135, 3415–3424. [Google Scholar] [CrossRef] [Green Version]
  30. Schaefer, S.A.; Higashi, A.Y.; Loomis, B.; Schrepfer, T.; Wan, G.; Corfas, G.; Dressler, G.R.; Duncan, R.K. From Otic Induction to Hair Cell Production: Pax2(EGFP) Cell Line Illuminates Key Stages of Development in Mouse Inner Ear Organoid Model. Stem. Cells Dev. 2018, 27, 237–251. [Google Scholar] [CrossRef]
  31. Carpena, N.T.; Chang, S.Y.; Choi, J.E.; Jung, J.Y.; Lee, M.Y. Wnt Modulation Enhances Otic Differentiation by Facilitating the Enucleation Process but Develops Unnecessary Cardiac Structures. Int. J. Mol. Sci. 2021, 22, 10306. [Google Scholar] [CrossRef]
  32. Liu, X.P.; Koehler, K.R.; Mikosz, A.M.; Hashino, E.; Holt, J.R. Functional development of mechanosensitive hair cells in stem cell-derived organoids parallels native vestibular hair cells. Nat. Commun. 2016, 7, 11508. [Google Scholar] [CrossRef] [Green Version]
  33. Moody, S.A.; LaMantia, A.-S. Transcriptional regulation of cranial sensory placode development. In Current Topics in Developmental Biology; Elsevier: Amsterdam, The Netherlands, 2015; Volume 111, pp. 301–350. [Google Scholar]
  34. Lee, H.-K.; Lee, H.-S.; Moody, S.A. Neural transcription factors: From embryos to neural stem cells. Mol. Cells 2014, 37, 705. [Google Scholar] [CrossRef] [Green Version]
  35. Moody, S.A.; Klein, S.L.; Karpinski, B.A.; Maynard, T.M.; LaMantia, A.-S. On becoming neural: What the embryo can tell us about differentiating neural stem cells. Am. J. Stem Cells 2013, 2, 74. [Google Scholar]
  36. Rogers, C.D.; Moody, S.A.; Casey, E.S. Neural induction and factors that stabilize a neural fate. Birth Defects Res. Part C: Embryo Today: Rev. 2009, 87, 249–262. [Google Scholar] [CrossRef] [Green Version]
  37. Sankar, S.; Yellajoshyula, D.; Zhang, B.; Teets, B.; Rockweiler, N.; Kroll, K.L. Gene regulatory networks in neural cell fate acquisition from genome-wide chromatin association of Geminin and Zic1. Sci. Rep. 2016, 6, 37412. [Google Scholar] [CrossRef] [Green Version]
  38. Aruga, J.; Hatayama, M. Comparative genomics of the Zic family genes. In Zic Family; Springer: Berlin/Heidelberg, Germany, 2018; pp. 3–26. [Google Scholar]
  39. Yan, B.; Neilson, K.M.; Moody, S.A. foxD5 plays a critical upstream role in regulating neural ectodermal fate and the onset of neural differentiation. Dev. Biol. 2009, 329, 80–95. [Google Scholar] [CrossRef] [Green Version]
  40. Sadler, T. Embryology of neural tube development. Am. J. Med. Genet. C Semin. Med. Genet. 2005, 135C, 2–8. [Google Scholar]
  41. Schoenwolf, G.C.; Alvarez, I.S. Roles of neuroepithelial cell rearrangement and division in shaping of the avian neural plate. Development 1989, 106, 427–439. [Google Scholar] [CrossRef]
  42. Kageyama, R.; Shimojo, H.; Ohtsuka, T. Dynamic control of neural stem cells by bHLH factors. Neurosci. Res. 2019, 138, 12–18. [Google Scholar] [CrossRef]
  43. Reiprich, S.; Wegner, M. From CNS stem cells to neurons and glia: Sox for everyone. Cell Tissue Res. 2015, 359, 111–124. [Google Scholar] [CrossRef]
  44. Glover, J.C.; Elliott, K.L.; Erives, A.; Chizhikov, V.V.; Fritzsch, B. Wilhelm His’ lasting insights into hindbrain and cranial ganglia development and evolution. Dev. Biol. 2018, 444, S14–S24. [Google Scholar] [CrossRef]
  45. van der Heijden, M.E.; Zoghbi, H.Y. Development of the brainstem respiratory circuit. Wiley Interdiscip. Rev. Dev. Biol. 2020, 9, e366. [Google Scholar]
  46. Mishima, Y.; Lindgren, A.G.; Chizhikov, V.V.; Johnson, R.L.; Millen, K.J. Overlapping function of Lmx1a and Lmx1b in anterior hindbrain roof plate formation and cerebellar growth. J. Neurosci. 2009, 29, 11377–11384. [Google Scholar] [CrossRef] [Green Version]
  47. Lee, K.J.; Dietrich, P.; Jessell, T.M. Genetic ablation reveals that the roof plate is essential for dorsal interneuron specification. Nature 2000, 403, 734–740. [Google Scholar] [CrossRef]
  48. Birol, O.; Ohyama, T.; Edlund, R.K.; Drakou, K.; Georgiades, P.; Groves, A.K. The mouse Foxi3 transcription factor is necessary for the development of posterior placodes. Dev. Biol. 2016, 409, 139–151. [Google Scholar] [CrossRef] [Green Version]
  49. Urness, L.D.; Paxton, C.N.; Wang, X.; Schoenwolf, G.C.; Mansour, S.L. FGF signaling regulates otic placode induction and refinement by controlling both ectodermal target genes and hindbrain Wnt8a. Dev. Biol. 2010, 340, 595–604. [Google Scholar] [CrossRef] [Green Version]
  50. Wilson, L.; Maden, M. The mechanisms of dorsoventral patterning in the vertebrate neural tube. Dev. Biol. 2005, 282, 1–13. [Google Scholar] [CrossRef] [Green Version]
  51. Ulloa, F.; Martí, E. Wnt won the war: Antagonistic role of Wnt over Shh controls dorso-ventral patterning of the vertebrate neural tube. Dev. Dyn. Off. Publ. Am. Assoc. Anat. 2010, 239, 69–76. [Google Scholar] [CrossRef]
  52. Chizhikov, V.V.; Iskusnykh, I.Y.; Fattakhov, N.; Fritzsch, B. Lmx1a and Lmx1b are Redundantly Required for the Development of Multiple Components of the Mammalian Auditory System. Neuroscience 2021, 452, 247–264. [Google Scholar] [CrossRef]
  53. Winata, C.L.; Korzh, V. Zebrafish Zic genes mediate developmental signaling. In Zic Family; Springer: Berlin/Heidelberg, Germany, 2018; pp. 157–177. [Google Scholar]
  54. Barratt, K.S.; Arkell, R.M. ZIC2 in Holoprosencephaly. In Zic Family; Springer: Berlin/Heidelberg, Germany, 2018; pp. 269–299. [Google Scholar]
  55. Hernandez-Miranda, L.R.; Müller, T.; Birchmeier, C. The dorsal spinal cord and hindbrain: From developmental mechanisms to functional circuits. Dev. Biol. 2017, 432, 34–42. [Google Scholar] [CrossRef] [Green Version]
  56. Bermingham, N.A.; Hassan, B.A.; Wang, V.Y.; Fernandez, M.; Banfi, S.; Bellen, H.J.; Fritzsch, B.; Zoghbi, H.Y. Proprioceptor pathway development is dependent on Math1. Neuron 2001, 30, 411–422. [Google Scholar] [CrossRef] [Green Version]
  57. Wang, V.Y.; Rose, M.F.; Zoghbi, H.Y. Math1 expression redefines the rhombic lip derivatives and reveals novel lineages within the brainstem and cerebellum. Neuron 2005, 48, 31–43. [Google Scholar] [CrossRef] [Green Version]
  58. Fritzsch, B.; Pauley, S.; Feng, F.; Matei, V.; Nichols, D. The molecular and developmental basis of the evolution of the vertebrate auditory system. Int. J. Comp. Psychol. 2006, 19, 1–25. [Google Scholar] [CrossRef]
  59. Storm, R.; Cholewa-Waclaw, J.; Reuter, K.; Bröhl, D.; Sieber, M.; Treier, M.; Müller, T.; Birchmeier, C. The bHLH transcription factor Olig3 marks the dorsal neuroepithelium of the hindbrain and is essential for the development of brainstem nuclei. Development 2009, 136, 295–305. [Google Scholar] [CrossRef] [Green Version]
  60. Glover, J.C.; Fritzsch, B. Molecular mechanisms governing development of the hindbrain choroid plexus and auditory projection: A validation of the seminal observations of Wilhelm His. IBRO Neurosci. Rep. 2022, 13, 306–313. [Google Scholar] [CrossRef] [PubMed]
  61. Iskusnykh, I.Y.; Steshina, E.Y.; Chizhikov, V.V. Loss of Ptf1a leads to a widespread cell-fate misspecification in the brainstem, affecting the development of somatosensory and viscerosensory nuclei. J. Neurosci. 2016, 36, 2691–2710. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  62. Fujiyama, T.; Yamada, M.; Terao, M.; Terashima, T.; Hioki, H.; Inoue, Y.U.; Inoue, T.; Masuyama, N.; Obata, K.; Yanagawa, Y. Inhibitory and excitatory subtypes of cochlear nucleus neurons are defined by distinct bHLH transcription factors, Ptf1a and Atoh1. Development 2009, 136, 2049–2058. [Google Scholar] [CrossRef] [Green Version]
  63. Qian, Y.; Fritzsch, B.; Shirasawa, S.; Chen, C.-L.; Choi, Y.; Ma, Q. Formation of brainstem (nor) adrenergic centers and first-order relay visceral sensory neurons is dependent on homeodomain protein Rnx/Tlx3. Genes Dev. 2001, 15, 2533–2545. [Google Scholar] [CrossRef] [Green Version]
  64. Sakamoto, S.; Tateya, T.; Omori, K.; Kageyama, R. Id genes are required for morphogenesis and cellular patterning in the developing mammalian cochlea. Dev. Biol. 2020, 460, 164–175. [Google Scholar] [CrossRef] [PubMed]
  65. Fritzsch, B.; Jahan, I.; Pan, N.; Elliott, K.L. Evolving gene regulatory networks into cellular networks guiding adaptive behavior: An outline how single cells could have evolved into a centralized neurosensory system. Cell Tissue Res. 2015, 359, 295–313. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  66. Jahan, I.; Pan, N.; Elliott, K.L.; Fritzsch, B. The quest for restoring hearing: Understanding ear development more completely. Bioessays 2015, 37, 1016–1027. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  67. Li, H.J.; Ray, S.K.; Pan, N.; Haigh, J.; Fritzsch, B.; Leiter, A.B. Intestinal Neurod1 expression impairs paneth cell differentiation and promotes enteroendocrine lineage specification. Sci. Rep. 2019, 9, 19489. [Google Scholar] [CrossRef] [Green Version]
  68. Kersigo, J.; Gu, L.; Xu, L.; Pan, N.; Vijayakuma, S.; Jones, T.; Shibata, S.B.; Fritzsch, B.; Hansen, M.R. Effects of Neurod1 Expression on Mouse and Human Schwannoma Cells. Laryngoscope 2020, 137, E259–E270. [Google Scholar] [CrossRef]
  69. Cheng, Y.; Liao, S.; Xu, G.; Hu, J.; Guo, D.; Du, F.; Contreras, A.; Cai, K.Q.; Peri, S.; Wang, Y. NeuroD1 Dictates Tumor Cell Differentiation in Medulloblastoma. Cell Rep. 2020, 31, 107782. [Google Scholar] [CrossRef] [PubMed]
  70. Pan, N.; Jahan, I.; Lee, J.E.; Fritzsch, B. Defects in the cerebella of conditional Neurod1 null mice correlate with effective Tg (Atoh1-cre) recombination and granule cell requirements for Neurod1 for differentiation. Cell Tissue Res. 2009, 337, 407–428. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  71. Elliott, K.L.; Pavlinkova, G.; Chizhikov, V.V.; Yamoah, E.N.; Fritzsch, B. Neurog1, Neurod1, and Atoh1 are essential for spiral ganglia, cochlear nuclei, and cochlear hair cell development. Fac. Rev. 2021, 10, 47. [Google Scholar] [CrossRef]
  72. Thawani, A.; Groves, A.K. Building the Border: Development of the Chordate Neural Plate Border Region and Its Derivatives. Front. Physiol. 2020, 11, 608880. [Google Scholar] [CrossRef] [PubMed]
  73. Northcutt, R.G.; Gans, C. The genesis of neural crest and epidermal placodes: A reinterpretation of vertebrate origins. Q. Rev. Biol. 1983, 58, 1–28. [Google Scholar] [CrossRef]
  74. Schlosser, G. Development of Sensory and Neurosecretory Cell Types: Vertebrate Cranial Placodes; CRC Press: Boca Raton, FL, USA, 2021; Volume 1. [Google Scholar]
  75. Schlosser, G. Chapter Four—Making Senses: Development of Vertebrate Cranial Placodes. In International Review of Cell and Molecular Biology; Kwang, J., Ed.; Academic Press: Cambridge, MA, USA, 2010; Volume 283, pp. 129–234. [Google Scholar]
  76. Elliott, K.L.; Straka, H. Assembly and Functional Organization of the Vestibular System. In Evolution of Neurosensory Cells and Systems; CRC Press: Boca Raton, FL, USA, 2022; pp. 135–174. [Google Scholar]
  77. Schlosser, G.; Ahrens, K. Molecular anatomy of placode development in Xenopus laevis. Dev. Biol. 2004, 271, 439–466. [Google Scholar] [CrossRef] [PubMed]
  78. Schlosser, G. Vertebrate cranial placodes as evolutionary innovations—The ancestor’s tale. In Current Topics in Developmental Biology; Elsevier: Amsterdam, The Netherlands, 2015; Volume 111, pp. 235–300. [Google Scholar]
  79. Kopecky, B.; Fritzsch, B. Regeneration of Hair Cells: Making Sense of All the Noise. Pharmaceuticals 2011, 4, 848–879. [Google Scholar] [CrossRef]
  80. Fritzsch, B.; Pan, N.; Jahan, I.; Duncan, J.S.; Kopecky, B.J.; Elliott, K.L.; Kersigo, J.; Yang, T. Evolution and development of the tetrapod auditory system: An organ of Corti-centric perspective. Evol. Dev. 2013, 15, 63–79. [Google Scholar] [CrossRef] [Green Version]
  81. Fritzsch, B. An Integrated Perspective of Evolution and Development: From Genes to Function to Ear, Lateral Line and Electroreception. Diversity 2021, 13, 364. [Google Scholar] [CrossRef]
  82. Zou, D.; Silvius, D.; Fritzsch, B.; Xu, P.-X. Eya1 and Six1 are essential for early steps of sensory neurogenesis in mammalian cranial placodes. Development 2004, 131, 5561–5572. [Google Scholar] [CrossRef] [Green Version]
  83. Li, J.; Zhang, T.; Ramakrishnan, A.; Fritzsch, B.; Xu, J.; Wong, E.Y.; Loh, Y.-H.E.; Ding, J.; Shen, L.; Xu, P.-X. Dynamic changes in cis-regulatory occupancy by Six1 and its cooperative interactions with distinct cofactors drive lineage-specific gene expression programs during progressive differentiation of the auditory sensory epithelium. Nucleic Acids Res. 2020, 48, 2880–2896. [Google Scholar] [CrossRef] [PubMed]
  84. Pauley, S.; Lai, E.; Fritzsch, B. Foxg1 is required for morphogenesis and histogenesis of the mammalian inner ear. Dev. Dyn. Off. Publ. Am. Assoc. Anat. 2006, 235, 2470–2482. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  85. Bouchard, M.; de Caprona, D.; Busslinger, M.; Xu, P.; Fritzsch, B. Pax2 and Pax8 cooperate in mouse inner ear morphogenesis and innervation. BMC Dev. Biol. 2010, 10, 89. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  86. Wright, T.J.; Mansour, S.L. Fgf3 and Fgf10 are required for mouse otic placode induction. Development 2003, 130, 3379–3390. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  87. Pirvola, U.; Spencer-Dene, B.; Xing-Qun, L.; Kettunen, P.; Thesleff, I.; Fritzsch, B.; Dickson, C.; Ylikoski, J. FGF/FGFR-2 (IIIb) signaling is essential for inner ear morphogenesis. J. Neurosci. 2000, 20, 6125–6134. [Google Scholar] [CrossRef] [Green Version]
  88. Pauley, S.; Wright, T.J.; Pirvola, U.; Ornitz, D.; Beisel, K.; Fritzsch, B. Expression and function of FGF10 in mammalian inner ear development. Dev. Dyn. Off. Publ. Am. Assoc. Anat. 2003, 227, 203–215. [Google Scholar] [CrossRef] [Green Version]
  89. Ladher, R.K.; O’Neill, P.; Begbie, J. From shared lineage to distinct functions: The development of the inner ear and epibranchial placodes. Development 2010, 137, 1777–1785. [Google Scholar] [CrossRef] [Green Version]
  90. Urness, L.D.; Bleyl, S.B.; Wright, T.J.; Moon, A.M.; Mansour, S.L. Redundant and dosage sensitive requirements for Fgf3 and Fgf10 in cardiovascular development. Dev. Biol. 2011, 356, 383–397. [Google Scholar] [CrossRef] [Green Version]
  91. Steventon, B.; Mayor, R.; Streit, A. Mutual repression between Gbx2 and Otx2 in sensory placodes reveals a general mechanism for ectodermal patterning. Dev. Biol. 2012, 367, 55–65. [Google Scholar] [CrossRef] [Green Version]
  92. Karis, A.; Pata, I.; van Doorninck, J.H.; Grosveld, F.; de Zeeuw, C.I.; de Caprona, D.; Fritzsch, B. Transcription factor GATA-3 alters pathway selection of olivocochlear neurons and affects morphogenesis of the ear. J. Comp. Neurol. 2001, 429, 615–630. [Google Scholar] [CrossRef]
  93. Duncan, J.S.; Fritzsch, B. Continued expression of GATA3 is necessary for cochlear neurosensory development. PLoS ONE 2013, 8, e62046. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  94. Riccomagno, M.M.; Martinu, L.; Mulheisen, M.; Wu, D.K.; Epstein, D.J. Specification of the mammalian cochlea is dependent on Sonic hedgehog. Genes Dev. 2002, 16, 2365–2378. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  95. Groves, A.K.; Fekete, D.M. Shaping sound in space: The regulation of inner ear patterning. Development 2012, 139, 245–257. [Google Scholar] [CrossRef] [Green Version]
  96. Wu, D.K.; Kelley, M.W. Molecular mechanisms of inner ear development. Cold Spring Harb. Perspect Biol. 2012, 4, a008409. [Google Scholar] [CrossRef]
  97. Bok, J.; Bronner-Fraser, M.; Wu, D.K. Role of the hindbrain in dorsoventral but not anteroposterior axial specification of the inner ear. Development 2005, 132, 2115–2124. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  98. Chang, W.; Lin, Z.; Kulessa, H.; Hebert, J.; Hogan, B.L.; Wu, D.K. Bmp4 is essential for the formation of the vestibular apparatus that detects angular head movements. PLoS Genet. 2008, 4, e1000050. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  99. Ohyama, T.; Basch, M.L.; Mishina, Y.; Lyons, K.M.; Segil, N.; Groves, A.K. BMP signaling is necessary for patterning the sensory and nonsensory regions of the developing mammalian cochlea. J. Neurosci. 2010, 30, 15044–15051. [Google Scholar] [CrossRef] [Green Version]
  100. Hwang, C.H.; Simeone, A.; Lai, E.; Wu, D.K. Foxg1 is required for proper separation and formation of sensory cristae during inner ear development. Dev. Dyn. Off. Publ. Am. Assoc. Anat. 2009, 238, 2725–2734. [Google Scholar]
  101. Nichols, D.H.; Bouma, J.E.; Kopecky, B.J.; Jahan, I.; Beisel, K.W.; He, D.Z.; Liu, H.; Fritzsch, B. Interaction with ectopic cochlear crista sensory epithelium disrupts basal cochlear sensory epithelium development in Lmx1a mutant mice. Cell Tissue Res. 2020, 380, 435–448. [Google Scholar] [CrossRef]
  102. Huang, Y.; Hill, J.; Yatteau, A.; Wong, L.; Jiang, T.; Petrovic, J.; Gan, L.; Dong, L.; Wu, D.K. Reciprocal negative regulation between Lmx1a and Lmo4 is required for inner ear formation. J. Neurosci. 2018, 38, 5429–5440. [Google Scholar] [CrossRef]
  103. Fritzsch, B.; Signore, M.; Simeone, A. Otx1 null mutant mice show partial segregation of sensory epithelia comparable to lamprey ears. Dev. Genes Evol. 2001, 211, 388–396. [Google Scholar] [CrossRef] [PubMed]
  104. Vendrell, V.; López-Hernández, I.; Alonso, M.B.D.; Feijoo-Redondo, A.; Abello, G.; Gálvez, H.; Giráldez, F.; Lamonerie, T.; Schimmang, T. Otx2 is a target of N-myc and acts as a suppressor of sensory development in the mammalian cochlea. Development 2015, 142, 2792–2800. [Google Scholar] [CrossRef] [Green Version]
  105. Kopecky, B.; Santi, P.; Johnson, S.; Schmitz, H.; Fritzsch, B. Conditional deletion of N-Myc disrupts neurosensory and non-sensory development of the ear. Dev. Dyn. 2011, 240, 1373–1390. [Google Scholar] [CrossRef] [Green Version]
  106. Gnedeva, K.; Wang, X.; McGovern, M.M.; Barton, M.; Tao, L.; Trecek, T.; Monroe, T.O.; Llamas, J.; Makmura, W.; Martin, J.F.; et al. Organ of Corti size is governed by Yap/Tead-mediated progenitor self-renewal. Proc. Natl. Acad. Sci. USA 2020, 117, 13552–13561. [Google Scholar] [CrossRef] [PubMed]
  107. Ono, K.; Sandell, L.L.; Trainor, P.A.; Wu, D.K. Retinoic acid synthesis and autoregulation mediate zonal patterning of vestibular organs and inner ear morphogenesis. Development 2020, 147, dev192070. [Google Scholar] [CrossRef] [PubMed]
  108. Dvorakova, M.; Macova, I.; Bohuslavova, R.; Anderova, M.; Fritzsch, B.; Pavlinkova, G. Early ear neuronal development, but not olfactory or lens development, can proceed without SOX2. Dev. Biol. 2020, 457, 43–56. [Google Scholar] [CrossRef] [PubMed]
  109. Ma, Q.; Anderson, D.J.; Fritzsch, B. Neurogenin 1 null mutant ears develop fewer, morphologically normal hair cells in smaller sensory epithelia devoid of innervation. J. Assoc. Res. Otolaryngol. 2000, 1, 129–143. [Google Scholar] [CrossRef] [Green Version]
  110. Matei, V.; Pauley, S.; Kaing, S.; Rowitch, D.; Beisel, K.; Morris, K.; Feng, F.; Jones, K.; Lee, J.; Fritzsch, B. Smaller inner ear sensory epithelia in Neurog1 null mice are related to earlier hair cell cycle exit. Dev. Dyn. 2005, 234, 633–650. [Google Scholar] [CrossRef] [Green Version]
  111. Kim, W.-Y.; Fritzsch, B.; Serls, A.; Bakel, L.A.; Huang, E.J.; Reichardt, L.F.; Barth, D.S.; Lee, J.E. NeuroD-null mice are deaf due to a severe loss of the inner ear sensory neurons during development. Development 2001, 128, 417–426. [Google Scholar] [CrossRef]
  112. Macova, I.; Pysanenko, K.; Chumak, T.; Dvorakova, M.; Bohuslavova, R.; Syka, J.; Fritzsch, B.; Pavlinkova, G. Neurod1 is essential for the primary tonotopic organization and related auditory information processing in the midbrain. J. Neurosci. 2019, 39, 984–1004. [Google Scholar] [CrossRef] [Green Version]
  113. Krüger, M.; Schmid, T.; Krüger, S.; Bober, E.; Braun, T. Functional redundancy of NSCL-1 and NeuroD during development of the petrosal and vestibulocochlear ganglia. Eur. J. Neurosci. 2006, 24, 1581–1590. [Google Scholar] [CrossRef] [PubMed]
  114. Jahan, I.; Pan, N.; Kersigo, J.; Fritzsch, B. Neurod1 suppresses hair cell differentiation in ear ganglia and regulates hair cell subtype development in the cochlea. PLoS ONE 2010, 5, e11661. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  115. Huang, E.J.; Liu, W.; Fritzsch, B.; Bianchi, L.M.; Reichardt, L.F.; Xiang, M. Brn3a is a transcriptional regulator of soma size, target field innervation and axon pathfinding of inner ear sensory neurons. Development 2001, 128, 2421–2432. [Google Scholar] [CrossRef]
  116. Petitpré, C.; Faure, L.; Uhl, P.; Fontanet, P.; Filova, I.; Pavlinkova, G.; Adameyko, I.; Hadjab, S.; Lallemend, F. Single-cell RNA-sequencing analysis of the developing mouse inner ear identifies molecular logic of auditory neuron diversification. Nat. Commun. 2022, 13, 1–15. [Google Scholar] [CrossRef]
  117. Salehi, P.; Ge, M.X.; Gundimeda, U.; Michelle Baum, L.; Lael Cantu, H.; Lavinsky, J.; Tao, L.; Myint, A.; Cruz, C.; Wang, J.; et al. Role of Neuropilin-1/Semaphorin-3A signaling in the functional and morphological integrity of the cochlea. PLoS Genet. 2017, 13, e1007048. [Google Scholar] [CrossRef] [Green Version]
  118. Yang, T.; Kersigo, J.; Jahan, I.; Pan, N.; Fritzsch, B. The molecular basis of making spiral ganglion neurons and connecting them to hair cells of the organ of Corti. Hear. Res. 2011, 278, 21–33. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  119. Ghimire, S.R.; Deans, M.R. Frizzled3 and Frizzled6 Cooperate with Vangl2 to Direct Cochlear Innervation by type II Spiral Ganglion Neurons. J. Neurosci. 2019, 39, 8013–8023. [Google Scholar] [CrossRef] [Green Version]
  120. Duncan, J.S.; Fritzsch, B.; Houston, D.W.; Ketchum, E.M.; Kersigo, J.; Deans, M.R.; Elliott, K.L. Topologically correct central projections of tetrapod inner ear afferents require Fzd3. Sci. Rep. 2019, 9, 10298. [Google Scholar] [CrossRef] [Green Version]
  121. Schmidt, H.; Fritzsch, B. Npr2 null mutants show initial overshooting followed by reduction of spiral ganglion axon projections combined with near-normal cochleotopic projection. Cell Tissue Res. 2019, 378, 15–32. [Google Scholar] [CrossRef]
  122. Hoshino, N.; Altarshan, Y.; Alzein, A.; Fernando, A.M.; Nguyen, H.T.; Majewski, E.F.; Chen, V.C.; Rochlin, M.W.; Yu, W.M. Ephrin-A3 is required for tonotopic map precision and auditory functions in the mouse auditory brainstem. J. Comp. Neurol. 2021, 529, 3633–3654. [Google Scholar] [CrossRef]
  123. Hong, C.-S.; Saint-Jeannet, J.-P. Sox proteins and neural crest development. Semin. Cell Dev. Biol. 2005, 16, 694–703. [Google Scholar] [CrossRef] [PubMed]
  124. Rogers, C.; Jayasena, C.; Nie, S.; Bronner, M.E. Neural crest specification: Tissues, signals, and transcription factors. Wiley Interdiscip. Rev. Dev. Biol. 2012, 1, 52–68. [Google Scholar] [CrossRef] [PubMed]
  125. Mao, Y.; Reiprich, S.; Wegner, M.; Fritzsch, B. Targeted deletion of Sox10 by Wnt1-cre defects neuronal migration and projection in the mouse inner ear. PLoS ONE 2014, 9, e94580. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  126. Nomdedeu-Sancho, G.; Alsina, B. Wiring the senses: Factors that regulate peripheral axon pathfinding in sensory systems. Dev. Dyn. 2022, 252, 81–103. [Google Scholar] [CrossRef] [PubMed]
  127. O’Neill, P.; Mak, S.-S.; Fritzsch, B.; Ladher, R.K.; Baker, C.V. The amniote paratympanic organ develops from a previously undiscovered sensory placode. Nat. Commun. 2012, 3, 1041. [Google Scholar] [CrossRef] [Green Version]
  128. Smith, A.C.; Fleenor, S.J.; Begbie, J. Changes in gene expression and cell shape characterise stages of epibranchial placode-derived neuron maturation in the chick. J. Anat. 2015, 227, 89–102. [Google Scholar] [CrossRef] [Green Version]
  129. Ma, Q.; Chen, Z.; del Barco Barrantes, I.; de la Pompa, J.L.; Anderson, D.J. neurogenin1 is essential for the determination of neuronal precursors for proximal cranial sensory ganglia. Neuron 1998, 20, 469–482. [Google Scholar] [CrossRef] [Green Version]
  130. Fode, C.; Gradwohl, G.; Morin, X.; Dierich, A.; LeMeur, M.; Goridis, C.; Guillemot, F. The bHLH protein NEUROGENIN 2 is a determination factor for epibranchial placode–derived sensory neurons. Neuron 1998, 20, 483–494. [Google Scholar] [CrossRef] [Green Version]
  131. Thiery, A.; Buzzi, A.L.; Streit, A. Cell fate decisions during the development of the peripheral nervous system in the vertebrate head. Curr. Top. Dev. Biol. 2020, 139, 127–167. [Google Scholar]
  132. Riddiford, N.; Schlosser, G. Dissecting the pre-placodal transcriptome to reveal presumptive direct targets of Six1 and Eya1 in cranial placodes. Elife 2016, 5, e17666. [Google Scholar] [CrossRef]
  133. Lysakowski, A. Anatomy and microstrucctural organization of vestibular hair cells. In The Senses; Fritzsch, B., Straka, H., Eds.; Elsevier: Amsterdam, The Netherlands, 2021; Volume 6, pp. 173–184. [Google Scholar]
  134. Burns, J.C.; Stone, J.S. Development and regeneration of vestibular hair cells in mammals. Semin. Cell Dev. Biol. 2017, 65, 96–105. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  135. Fritzsch, B.; Matei, V.; Nichols, D.; Bermingham, N.; Jones, K.; Beisel, K.; Wang, V. Atoh1 null mice show directed afferent fiber growth to undifferentiated ear sensory epithelia followed by incomplete fiber retention. Dev. Dyn. Off. Publ. Am. Assoc. Anat. 2005, 233, 570–583. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  136. Matern, M.S.; Milon, B.; Lipford, E.L.; McMurray, M.; Ogawa, Y.; Tkaczuk, A.; Song, Y.; Elkon, R.; Hertzano, R.J.D. GFI1 functions to repress neuronal gene expression in the developing inner ear hair cells. Development 2020, 147, dev186015. [Google Scholar] [CrossRef]
  137. Jahan, I.; Pan, N.; Kersigo, J.; Fritzsch, B. Neurog1 can partially substitute for Atoh1 function in hair cell differentiation and maintenance during organ of Corti development. Development 2015, 142, 2810–2821. [Google Scholar] [PubMed] [Green Version]
  138. Koo, H.Y.; Kim, M.-A.; Min, H.; Hwang, J.Y.; Prajapati-DiNubila, M.; Kim, K.S.; Matzuk, M.M.; Park, J.W.; Doetzlhofer, A.; Kim, U.-K. Follistatin regulates the specification of the apical cochlea responsible for low-frequency hearing in mammals. Proc. Natl. Acad. Sci. 2023, 120, e2213099120. [Google Scholar] [CrossRef] [PubMed]
  139. Lorenzen, S.M.; Duggan, A.; Osipovich, A.B.; Magnuson, M.A.; García-Añoveros, J. Insm1 promotes neurogenic proliferation in delaminated otic progenitors. Mech. Dev. 2015, 138 Pt 3, 233–245. [Google Scholar] [CrossRef]
  140. García-Añoveros, J.; Clancy, J.C.; Foo, C.Z.; García-Gómez, I.; Zhou, Y.; Homma, K.; Cheatham, M.A.; Duggan, A. Tbx2 is a master regulator of inner versus outer hair cell differentiation. Nature 2022, 605, 298–303. [Google Scholar] [CrossRef]
  141. Nakano, Y.; Wiechert, S.; Fritzsch, B.; Bánfi, B. Inhibition of a transcriptional repressor rescues hearing in a splicing factor–deficient mouse. Life Sci. Alliance 2020, 3, e202000841. [Google Scholar] [CrossRef]
  142. Kaiser, M.; Wojahn, I.; Rudat, C.; Lüdtke, T.H.; Christoffels, V.M.; Moon, A.; Kispert, A.; Trowe, M.O. Regulation of otocyst patterning by Tbx2 and Tbx3 is required for inner ear morphogenesis in the mouse. Development 2021, 148, dev195651. [Google Scholar] [CrossRef]
  143. Golden, E.J.; Larson, E.D.; Shechtman, L.A.; Trahan, G.D.; Gaillard, D.; Fellin, T.J.; Scott, J.K.; Jones, K.L.; Barlow, L.A. Onset of taste bud cell renewal starts at birth and coincides with a shift in SHH function. Elife 2021, 10, e64013. [Google Scholar] [CrossRef]
  144. Ohmoto, M.; Kitamoto, S.; Hirota, J. Expression of Eya1 in mouse taste buds. Cell Tissue Res 2020, 383, 979–986. [Google Scholar] [CrossRef] [PubMed]
  145. Okubo, T.; Pevny, L.H.; Hogan, B.L. Sox2 is required for development of taste bud sensory cells. Genes Dev. 2006, 20, 2654–2659. [Google Scholar] [CrossRef] [Green Version]
  146. Kersigo, J.; Fritzsch, B. Inner ear hair cells deteriorate in mice engineered to have no or diminished innervation. Front. Aging Neurosci. 2015, 7, 33. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  147. Ueyama, T.; Sakaguchi, H.; Nakamura, T.; Goto, A.; Morioka, S.; Shimizu, A.; Nakao, K.; Hishikawa, Y.; Ninoyu, Y.; Kassai, H. Maintenance of stereocilia and apical junctional complexes by Cdc42 in cochlear hair cells. J. Cell Sci. 2014, 127, 2040–2052. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  148. Kirjavainen, A.; Laos, M.; Anttonen, T.; Pirvola, U. The Rho GTPase Cdc42 regulates hair cell planar polarity and cellular patterning in the developing cochlea. Biol. Open 2015, 4, 516–526. [Google Scholar] [CrossRef] [Green Version]
  149. Dai, Y.-B.; Gao, X.; Liu, D.; Gong, J. The role of Rho GTPase family in cochlear hair cells and hearing. Neural Regen. Res. 2023, 18, 2167–2172. [Google Scholar]
  150. Herranen, A.; Ikäheimo, K.; Lankinen, T.; Pakarinen, E.; Fritzsch, B.; Saarma, M.; Lindahl, M.; Pirvola, U. Deficiency of the ER-stress-regulator MANF triggers progressive outer hair cell death and hearing loss. Cell Death Dis. 2020, 11, 1–12. [Google Scholar] [CrossRef] [Green Version]
  151. Elliott, K.L.; Kersigo, J.; Lee, J.H.; Jahan, I.; Pavlinkova, G.; Fritzsch, B.; Yamoah, E.N. Developmental Changes in Peripherin-eGFP Expression in Spiral Ganglion Neurons. Front. Cell. Neurosci. 2021, 15, 678113. [Google Scholar] [CrossRef]
  152. Elliott, K.L.; Fritzsch, B.; Duncan, J.S. Evolutionary and developmental biology provide insights into the regeneration of organ of Corti hair cells. Front. Cell. Neurosci. 2018, 12, 252. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  153. Paulson, R.B.; Alley, K.E.; Salata, L.J.; Whitmyer, C.C. A scanning electron-microscopic study of tongue development in the frog Rana pipiens. Arch. Oral. Biol. 1995, 40, 311–319. [Google Scholar] [CrossRef]
  154. Barlow, L.A. Progress and renewal in gustation: New insights into taste bud development. Development 2015, 142, 3620–3629. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  155. Barlow, L.A.; Northcutt, R.G. Taste buds develop autonomously from endoderm without induction by cephalic neural crest or paraxial mesoderm. Development 1997, 124, 949–957. [Google Scholar] [CrossRef] [PubMed]
  156. Liu, F.; Thirumangalathu, S.; Gallant, N.M.; Yang, S.H.; Stoick-Cooper, C.L.; Reddy, S.T.; Andl, T.; Taketo, M.M.; Dlugosz, A.A.; Moon, R.T.; et al. Wnt-beta-catenin signaling initiates taste papilla development. Nat. Genet. 2007, 39, 106–112. [Google Scholar] [CrossRef]
  157. Iwatsuki, K.; Liu, H.X.; Gronder, A.; Singer, M.A.; Lane, T.F.; Grosschedl, R.; Mistretta, C.M.; Margolskee, R.F. Wnt signaling interacts with Shh to regulate taste papilla development. Proc. Natl. Acad. Sci. USA 2007, 104, 2253–2258. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  158. Hall, J.M.; Bell, M.L.; Finger, T.E. Disruption of sonic hedgehog signaling alters growth and patterning of lingual taste papillae. Dev. Biol. 2003, 255, 263–277. [Google Scholar] [CrossRef] [Green Version]
  159. Mistretta, C.M.; Liu, H.X.; Gaffield, W.; MacCallum, D.K. Cyclopamine and jervine in embryonic rat tongue cultures demonstrate a role for Shh signaling in taste papilla development and patterning: Fungiform papillae double in number and form in novel locations in dorsal lingual epithelium. Dev. Biol. 2003, 254, 1–18. [Google Scholar] [CrossRef] [Green Version]
  160. Petersen, C.I.; Jheon, A.H.; Mostowfi, P.; Charles, C.; Ching, S.; Thirumangalathu, S.; Barlow, L.A.; Klein, O.D. FGF signaling regulates the number of posterior taste papillae by controlling progenitor field size. PLoS Genet. 2011, 7, e1002098. [Google Scholar] [CrossRef] [Green Version]
  161. Thirumangalathu, S.; Harlow, D.E.; Driskell, A.L.; Krimm, R.F.; Barlow, L.A. Fate mapping of mammalian embryonic taste bud progenitors. Development 2009, 136, 1519–1528. [Google Scholar] [CrossRef] [Green Version]
  162. Ito, A.; Nosrat, C.A. Gustatory papillae and taste bud development and maintenance in the absence of TrkB ligands BDNF and NT-4. Cell Tissue Res. 2009, 337, 349–359. [Google Scholar] [CrossRef]
  163. Fan, D.; Chettouh, Z.; Consalez, G.G.; Brunet, J.-F. Taste bud formation depends on taste nerves. Elife 2019, 8, e49226. [Google Scholar] [CrossRef]
  164. Thirumangalathu, S.; Barlow, L.A. β-Catenin signaling regulates temporally discrete phases of anterior taste bud development. Development 2015, 142, 4309–4317. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  165. Fritzsch, B.; Sarai, P.; Barbacid, M.; Silos-Santiago, I. Mice with a targeted disruption of the neurotrophin receptor trkB lose their gustatory ganglion cells early but do develop taste buds. Int. J. Dev. Neurosci. 1997, 15, 563–576. [Google Scholar] [CrossRef]
  166. Rios-Pilier, J.; Krimm, R.F. TrkB expression and dependence divides gustatory neurons into three subpopulations. Neural Dev. 2019, 14, 1–13. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  167. Mistretta, C.M.; Goosens, K.A.; Farinas, I.; Reichardt, L.F. Alterations in size, number, and morphology of gustatory papillae and taste buds in BDNF null mutant mice demonstrate neural dependence of developing taste organs. J. Comp. Neurol. 1999, 409, 13–24. [Google Scholar] [CrossRef]
  168. Castillo-Azofeifa, D.; Losacco, J.T.; Salcedo, E.; Golden, E.J.; Finger, T.E.; Barlow, L.A. Sonic hedgehog from both nerves and epithelium is a key trophic factor for taste bud maintenance. Development 2017, 144, 3054–3065. [Google Scholar] [CrossRef] [Green Version]
  169. Lu, W.-J.; Mann, R.K.; Nguyen, A.; Bi, T.; Silverstein, M.; Tang, J.Y.; Chen, X.; Beachy, P.A. Neuronal delivery of Hedgehog directs spatial patterning of taste organ regeneration. Proc. Natl. Acad. Sci. USA 2018, 115, E200–E209. [Google Scholar] [CrossRef] [Green Version]
  170. Lin, X.; Lu, C.; Ohmoto, M.; Choma, K.; Margolskee, R.F.; Matsumoto, I.; Jiang, P. R-spondin substitutes for neuronal input for taste cell regeneration in adult mice. Proc. Natl. Acad. Sci. USA 2021, 118, e2001833118. [Google Scholar] [CrossRef]
  171. Beidler, L.M.; Smallman, R.L. Renewal of cells within taste buds. J. Cell Biol. 1965, 27, 263–272. [Google Scholar] [CrossRef]
  172. Perea-Martinez, I.; Nagai, T.; Chaudhari, N. Functional cell types in taste buds have distinct longevities. PLoS ONE 2013, 8, e53399. [Google Scholar] [CrossRef] [Green Version]
  173. Okubo, T.; Clark, C.; Hogan, B.L. Cell lineage mapping of taste bud cells and keratinocytes in the mouse tongue and soft palate. Stem. Cells 2009, 27, 442–450. [Google Scholar] [CrossRef] [Green Version]
  174. Takeda, N.; Jain, R.; Li, D.; Li, L.; Lu, M.M.; Epstein, J.A. Lgr5 Identifies Progenitor Cells Capable of Taste Bud Regeneration after Injury. PLoS ONE 2013, 8, e66314. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  175. Ren, W.; Lewandowski, B.C.; Watson, J.; Aihara, E.; Iwatsuki, K.; Bachmanov, A.A.; Margolskee, R.F.; Jiang, P. Single Lgr5-or Lgr6-expressing taste stem/progenitor cells generate taste bud cells ex vivo. Proc. Natl. Acad. Sci. USA 2014, 111, 16401–16406. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  176. Ohmoto, M.; Ren, W.; Nishiguchi, Y.; Hirota, J.; Jiang, P.; Matsumoto, I. Genetic Lineage Tracing in Taste Tissues Using Sox2-CreERT2 Strain. Chem. Senses 2017, 42, 547–552. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  177. Kito-Shingaki, A.; Seta, Y.; Toyono, T.; Kataoka, S.; Kakinoki, Y.; Yanagawa, Y.; Toyoshima, K. Expression of GAD67 and Dlx5 in the taste buds of mice genetically lacking Mash1. Chem. Senses 2014, 39, 403–414. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  178. Seta, Y.; Oda, M.; Kataoka, S.; Toyono, T.; Toyoshima, K. Mash1 is required for the differentiation of AADC-positive type III cells in mouse taste buds. Dev. Dyn. 2011, 240, 775–784. [Google Scholar] [CrossRef] [PubMed]
  179. Matsumoto, I.; Ohmoto, M.; Narukawa, M.; Yoshihara, Y.; Abe, K. Skn-1a (Pou2f3) specifies taste receptor cell lineage. Nat. Neurosci. 2011, 14, 685–687. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  180. Sun, Y.; Wang, L.; Zhu, T.; Wu, B.; Wang, G.; Luo, Z.; Li, C.; Wei, W.; Liu, Z. Single-cell transcriptomic landscapes of the otic neuronal lineage at multiple early embryonic ages. Cell Rep. 2022, 38, 110542. [Google Scholar] [CrossRef]
  181. Filova, I.; Pysanenko, K.; Tavakoli, M.; Vochyanova, S.; Dvorakova, M.; Bohuslavova, R.; Smolik, O.; Fabriciova, V.; Hrabalova, P.; Benesova, S. ISL1 is necessary for auditory neuron development and contributes toward tonotopic organization. Proc. Natl. Acad. Sci. USA 2022, 119, e2207433119. [Google Scholar] [CrossRef]
  182. Kalluri, R. Similarities in the Biophysical Properties of Spiral-Ganglion and Vestibular-Ganglion Neurons in Neonatal Rats. Front. Neurosci. 2021, 15, 710275. [Google Scholar] [CrossRef]
  183. Kersigo, J.; D’Angelo, A.; Gray, B.D.; Soukup, G.A.; Fritzsch, B. The role of sensory organs and the forebrain for the development of the craniofacial shape as revealed by Foxg1-cre-mediated microRNA loss. Genesis 2011, 49, 326–341. [Google Scholar] [CrossRef] [Green Version]
  184. Muthu, V.; Rohacek, A.M.; Yao, Y.; Rakowiecki, S.M.; Brown, A.S.; Zhao, Y.-T.; Meyers, J.; Won, K.-J.; Ramdas, S.; Brown, C.D.; et al. Genomic architecture of Shh-dependent cochlear morphogenesis. Development 2019, 146, dev181339. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  185. Riccomagno, M.M.; Takada, S.; Epstein, D.J. Wnt-dependent regulation of inner ear morphogenesis is balanced by the opposing and supporting roles of Shh. Genes Dev. 2005, 19, 1612–1623. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  186. Sanders, T.R.; Kelley, M.W. Specification of neuronal subtypes in the spiral ganglion begins prior to birth in the mouse. Proc. Natl. Acad. Sci. USA 2022, 119, e2203935119. [Google Scholar] [CrossRef] [PubMed]
  187. Fritzsch, B.; Kersigo, J.; Yang, T.; Jahan, I.; Pan, N. Neurotrophic factor function during ear development: Expression changes define critical phases for neuronal viability. In The Primary Auditory Neurons of the Mammalian Cochlea; Springer: Berlin/Heidelberg, Germany, 2016; pp. 49–84. [Google Scholar]
  188. Elliott, K.L.; Kersigo, J.; Lee, J.H.; Yamoah, E.N.; Fritzsch, B. Sustained Loss of Bdnf Affects Peripheral but Not Central Vestibular Targets. Front. Neurol. 2021, 12, 768456. [Google Scholar] [CrossRef] [PubMed]
  189. Wang, L.; Xie, J.; Zhang, H.; Tsang, L.H.; Tsang, S.L.; Braune, E.B.; Lendahl, U.; Sham, M.H. Notch signalling regulates epibranchial placode patterning and segregation. Development 2020, 147, dev183665. [Google Scholar] [CrossRef] [Green Version]
  190. Zhang, H.; Wang, L.; Wong, E.Y.M.; Tsang, S.L.; Xu, P.X.; Lendahl, U.; Sham, M.H. An Eya1-Notch axis specifies bipotential epibranchial differentiation in mammalian craniofacial morphogenesis. Elife 2017, 6, e30126. [Google Scholar] [CrossRef]
  191. 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] [PubMed]
  192. Freter, S.; Muta, Y.; O’Neill, P.; Vassilev, V.S.; Kuraku, S.; Ladher, R.K. Pax2 modulates proliferation during specification of the otic and epibranchial placodes. Dev. Dyn. 2012, 241, 1716–1728. [Google Scholar] [CrossRef]
  193. Dykes, I.M.; Tempest, L.; Lee, S.I.; Turner, E.E. Brn3a and Islet1 act epistatically to regulate the gene expression program of sensory differentiation. J. Neurosci. 2011, 31, 9789–9799. [Google Scholar] [CrossRef] [Green Version]
  194. D’Autreaux, F.; Coppola, E.; Hirsch, M.R.; Birchmeier, C.; Brunet, J.F. Homeoprotein Phox2b commands a somatic-to-visceral switch in cranial sensory pathways. Proc. Natl. Acad. Sci. USA 2011, 108, 20018–20023. [Google Scholar] [CrossRef] [Green Version]
  195. Fritzsch, B.; Elliott, K.L.; Yamoah, E.N. Neurosensory development of the four brainstem-projecting sensory systems and their integration in the telencephalon. Front. Neural Circuits 2022, 16, 913480. [Google Scholar] [CrossRef] [PubMed]
Figure 2. Otic placode (OP) and epibranchial placode (EP) induction. Fgf3 and Wnt1/3a/8a from the hindbrain downstream of Lmx1a/b as well as Fgf8 and Fgf10 from the surrounding somite’s combined with Foxi3 are believed to be sufficient to induce the otic and epibranchial placode. The subsequent mechanism is not fully elucidated but may involve Jagged1 and Notch1 inhibition of Foxi2. Upregulation of Dlx5/6, Eya1, Foxg1, Gata3, Gbx2, Hes2, Hmx2/3, Lmx1a/b, Pax2/8, Six1, Sox9, Spry1 and others marks the otic and epibranchial placode and is essential for later otic vesicle specification and morphogenesis. Foxi2 delineates the ectoderm. Adapted with permission from Refs. [48,49,72,79].
Figure 2. Otic placode (OP) and epibranchial placode (EP) induction. Fgf3 and Wnt1/3a/8a from the hindbrain downstream of Lmx1a/b as well as Fgf8 and Fgf10 from the surrounding somite’s combined with Foxi3 are believed to be sufficient to induce the otic and epibranchial placode. The subsequent mechanism is not fully elucidated but may involve Jagged1 and Notch1 inhibition of Foxi2. Upregulation of Dlx5/6, Eya1, Foxg1, Gata3, Gbx2, Hes2, Hmx2/3, Lmx1a/b, Pax2/8, Six1, Sox9, Spry1 and others marks the otic and epibranchial placode and is essential for later otic vesicle specification and morphogenesis. Foxi2 delineates the ectoderm. Adapted with permission from Refs. [48,49,72,79].
Ijms 24 06994 g002
Figure 3. Generating the otic vesicle and the epibranchial placode. (A) The otic vesicle is depicted in a mouse using 3D reconstruction of the ear. An early formation segregates from the endolymphatic duct (white). Separate color is defined as the canal cristae (blue) next to the utricle (white) that interconnects the saccule (lilac). Note the progression of the length of the cochlear duct can about E16.5 (yellow). (B) The neuronal differentiation of epibranchial placodes starts with Eya/Six next to the otic vesicle (OV) followed by Pax2 and Sox2, which initiate transformation of the geniculate (G), petrosal (P) and nodose (N). As neuroblasts migrate from the ectoderm to deeper locations, additional factors are expressed in sequence, Neurog2 (in mice, Neurog1 in chicken) followed by Neurod1, Isl1, Foxg1, Pou4f1 and Phox2b. Abbreviations: OP, olfactory placode; OV, otic vesicle, T, trigeminal neurons; VA, vestibular and auditory neurons. Reprinted with permission from Refs [17,105,128].
Figure 3. Generating the otic vesicle and the epibranchial placode. (A) The otic vesicle is depicted in a mouse using 3D reconstruction of the ear. An early formation segregates from the endolymphatic duct (white). Separate color is defined as the canal cristae (blue) next to the utricle (white) that interconnects the saccule (lilac). Note the progression of the length of the cochlear duct can about E16.5 (yellow). (B) The neuronal differentiation of epibranchial placodes starts with Eya/Six next to the otic vesicle (OV) followed by Pax2 and Sox2, which initiate transformation of the geniculate (G), petrosal (P) and nodose (N). As neuroblasts migrate from the ectoderm to deeper locations, additional factors are expressed in sequence, Neurog2 (in mice, Neurog1 in chicken) followed by Neurod1, Isl1, Foxg1, Pou4f1 and Phox2b. Abbreviations: OP, olfactory placode; OV, otic vesicle, T, trigeminal neurons; VA, vestibular and auditory neurons. Reprinted with permission from Refs [17,105,128].
Ijms 24 06994 g003
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Zine, A.; Fritzsch, B. Early Steps towards Hearing: Placodes and Sensory Development. Int. J. Mol. Sci. 2023, 24, 6994. https://doi.org/10.3390/ijms24086994

AMA Style

Zine A, Fritzsch B. Early Steps towards Hearing: Placodes and Sensory Development. International Journal of Molecular Sciences. 2023; 24(8):6994. https://doi.org/10.3390/ijms24086994

Chicago/Turabian Style

Zine, Azel, and Bernd Fritzsch. 2023. "Early Steps towards Hearing: Placodes and Sensory Development" International Journal of Molecular Sciences 24, no. 8: 6994. https://doi.org/10.3390/ijms24086994

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