Next Article in Journal
Exploring the Pathophysiology of ATP-Dependent Potassium Channels in Insulin Resistance
Next Article in Special Issue
Proteomic Analysis of Domestic Cat Blastocysts and Their Secretome Produced in an In Vitro Culture System without the Presence of the Zona Pellucida
Previous Article in Journal
Responsive Supramolecular Polymers for Diagnosis and Treatment
Previous Article in Special Issue
Gene Expression Profiling Reveals Fundamental Sex-Specific Differences in SIRT3-Mediated Redox and Metabolic Signaling in Mouse Embryonic Fibroblasts
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Calcium and Neural Stem Cell Proliferation

by
Dafne Astrid Díaz-Piña
1,2,
Nayeli Rivera-Ramírez
1,
Guadalupe García-López
1,
Néstor Fabián Díaz
1 and
Anayansi Molina-Hernández
1,*
1
Departamento de Fisiología y Desarrollo Celular, Instituto Nacional de Perinatología Isidro Espinosa de los Reyes, Montes Urales 800, Miguel Hidalgo, Ciudad de México 11000, Mexico
2
Facultad de Medicina, Circuito Exterior Universitario, Universidad Nacional Autónoma de México Universitario, Copilco Universidad, Coyoacán, Ciudad de México 04360, Mexico
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2024, 25(7), 4073; https://doi.org/10.3390/ijms25074073
Submission received: 8 February 2024 / Revised: 31 March 2024 / Accepted: 2 April 2024 / Published: 6 April 2024
(This article belongs to the Special Issue Embryonic Development and Differentiation)

Abstract

:
Intracellular calcium plays a pivotal role in central nervous system (CNS) development by regulating various processes such as cell proliferation, migration, differentiation, and maturation. However, understanding the involvement of calcium (Ca2+) in these processes during CNS development is challenging due to the dynamic nature of this cation and the evolving cell populations during development. While Ca2+ transient patterns have been observed in specific cell processes and molecules responsible for Ca2+ homeostasis have been identified in excitable and non-excitable cells, further research into Ca2+ dynamics and the underlying mechanisms in neural stem cells (NSCs) is required. This review focuses on molecules involved in Ca2+ entrance expressed in NSCs in vivo and in vitro, which are crucial for Ca2+ dynamics and signaling. It also discusses how these molecules might play a key role in balancing cell proliferation for self-renewal or promoting differentiation. These processes are finely regulated in a time-dependent manner throughout brain development, influenced by extrinsic and intrinsic factors that directly or indirectly modulate Ca2+ dynamics. Furthermore, this review addresses the potential implications of understanding Ca2+ dynamics in NSCs for treating neurological disorders. Despite significant progress in this field, unraveling the elements contributing to Ca2+ intracellular dynamics in cell proliferation remains a challenging puzzle that requires further investigation.

1. Introduction

In mammals, the central nervous system (CNS) originates from a simple neuroepithelium populated by neural stem cells (NSCs), known as neuroepithelial cells. Through symmetric proliferation, these cells give rise to a pseudo-stratified neuroepithelium, where bipolar NSCs called radial glia cells (RGCs) are observed. RGCs are multipotent stem cells capable of self-renewal and differentiation into neurons, astrocytes, and oligodendrocytes in a time-dependent manner through symmetric and asymmetric cell division [1,2,3].
NSCs, also referred to as neural progenitor cells (NPCs), must be distinguished from intermediate progenitor cells (IPCs), as the latter are committed to a neuronal phenotype. NSCs and IPCs constitute two proliferative zones in the ventricular and subventricular zones (VZ and SVZ), respectively. IPCs primarily generate pyramidal neurons of the cerebral cortex, while RGCs directly produce a small portion of neurons, astrocytes, and oligodendrocytes [4,5,6,7].
Both extrinsic and intrinsic factors regulate NSCs’ self-renewal and differentiation in time and space, thereby affecting cell proliferative patterns, partly determined based on cell cycle duration [8,9]. Extrinsic factors such as growth factors, neurotransmitters, and morphogens play significant roles in CNS development through selective membrane receptors. Receptor tyrosine kinases (RTKs), guanine nucleotide-binding protein-coupled receptors (G-protein-coupled receptors; GPCRs), and mechanoreceptors are among the receptors capable of eliciting intracellular calcium (Ca2+) transients in non-excitable cells. Consequently, the increase in intracellular Ca2+ in NSC is mediated downstream from receptor activation via two mechanisms: release from store organelles and entry through plasma membrane non-voltage-dependent channels.
The activity of Ca2+ as a second messenger relies on changes in the intracellular Ca2+ concentration ([Ca2+]i) at microdomains sensed by several Ca2+ binding proteins that regulate intracellular Ca2+ homeostasis by buffering and transporting this ion. These proteins include calmodulin (CaM), calcyclin, Ca2+-sensitive enzymes, transcription factors, and Ca2+-sensitive ion channels. Additionally, pumps and exchangers localized in mitochondria, endoplasmic reticulum (ER), and Golgi membranes contribute to Ca2+ homeostasis [10].
Due to the wide molecular repertoire involved in Ca2+ homeostasis, it is challenging to associate the molecules involved in Ca2+ dynamics and cell proliferation in the context of CNS development, as several phenomena coincide in time and space. However, proliferation is the first event during neural tube development, and alterations in this process can lead to CNS defects such as microcephaly, megalocephaly, and hemimegaloencephaly [11].
In the VZ, where NSCs reside in the cortical neuroepithelium of rat embryos, three patterns of spontaneous Ca2+ fluctuations have been reported: (1) isolated fluctuation in a single cell, (2) paired fluctuation in two adjacent cells, and (3) clustered or synchronized fluctuation in neighboring cells, each with particular kinetics and frequency patterns [12,13]. Interestingly, in 15- to 20-day-old rat embryos (E15–E20), most cells in the VZ exhibit the isolated Ca2+ fluctuation pattern, followed by pair fluctuation and cluster fluctuations [13]. However, these patterns have not been related to symmetric or asymmetric RGC proliferation or the expression of non-voltage-gated channels.
NSCs proliferate symmetrically to increase their pool or asymmetrically to give rise, directly or indirectly, in a time-dependent manner, to neurons, astrocytes, and oligodendrocytes (Figure 1). The potential fates of NSCs rely on their proliferative pattern [8]. Furthermore, symmetric cell division in the VZ is more frequent at early stages, and as development progresses, the asymmetric pattern starts to generate committed cells with or without proliferative capacity [14,15,16].
The proliferation and differentiation of NSC into neurons, astrocytes, and oligodendrocytes is a complex and highly regulated process. Certain factors exhibit dual roles depending on their concentration and cellular context. Among these factors are Notch, bone morphogenetic proteins (BMPs), brain-derived neurotrophic factor (BDNF), sonic hedgehog (SHH), Wingless (Wnt), and nerve growth factor (NGF).
Epidermal growth factor (EGF), fibroblast growth factor (FGF), and insulin-like growth factor (IGF) promote NSC proliferation. Neurogenic factors include neurotrophin-3 (NT-3), vascular endothelial growth factor (VEGF), and transforming growth factor-beta (TGF-β). Astrocyte differentiation is influenced by ciliary neurotrophic factor (CNTF), BMP2, BMP4, leukemia inhibitory factor (LIF), and interleukin-6 (IL-6). Oligodendrocyte differentiation, on the other hand, is influenced by platelet-derived growth factor (PDGF), IGF-I, neuregulin-1 (NRG-1), and FGF9 [1,3,8].

2. Calcium and Cell Cycle Regulation

Considerable information has been amassed concerning the role of Ca2+ in cell cycle regulation, albeit only a fraction of it originates directly from NSCs. While the mechanisms involved are likely redundant across cell types, certain peculiarities may exist between them. This section will discuss the role of Ca2+ in cell cycle regulation and the involvement of effector proteins at various stages of the cell cycle within the context of NSCs.
Cytosolic Ca2+ concentration ([Ca2+]c) regulates cell proliferation, with Ca2+ oscillations occurring throughout the cell cycle believed to regulate this process [17,18,19]. In RGC, [Ca2+]c increases via the binding of morphogens and growth factors to GPCRs and RTKs or through mechanoreceptor activation [19,20,21].
Changes in [Ca2+]c are associated with different phases of the cell cycle and the quiescent stage. In NSCs, Ca2+ signaling is essential for maintaining self-renewal or promoting cell differentiation by participating in various stages of the cell cycle, including the G1, S, G2, and M phases. Ca2+ binds to several proteins during this process, leading to alterations in the expression of cell cycle regulatory proteins [22,23]. During the G1 phase, the increase in [Ca2+]c can directly activate Ca2+-dependent kinases, such as Ca2+/CaM-dependent protein kinases (CaMKs) and protein kinases C (PKC), thereby increasing the expression of cell cycle regulatory proteins, including cyclins and cyclin-dependent kinases (CDKs). Ca2+ affects DNA replication during the S phase by modulating the activity of enzymes involved in DNA synthesis through Ca2+-dependent tyrosine kinase (PYK2), CaM, the activated protein kinase (MAPK), and ERK signaling pathways [24,25]. Furthermore, in G2, Ca2+-dependent kinases regulate cell cycle progression by controlling the activity of cyclin B and CDK1. In the M phase, Ca2+ signaling is essential for spindle formation, chromosome segregation, microtubule dynamics, and the activation of proteins such as mitotic kinases [26,27].
It has been suggested that in NSCs, changes in the duration of the G1 phase dictate self-renewal or differentiation, with a shorter G1 observed during symmetric division (self-renewal) and a longer G1 observed to be NSC divide in an asymmetric and symmetric non-proliferative fashion [28,29]. The role of Ca2+ in symmetric or asymmetric cell division remains unclear. However, using a cyclin-dependent kinase inhibitor to prolong the cell cycle appears sufficient to induce premature neurogenesis in mouse embryos [29]. This suggests that changes in [Ca2+]c during the G1 phase may influence the duration of this stage, potentially affected by factors such as receptor density and type, Ca2+ membrane channels, and levels of Ca2+-dependent kinases, all of which can impact cell cycle regulatory proteins. Interestingly, elevated levels of GPCRs associated with PKC activation and Ca2+ mobilization have been shown to enhance neuron differentiation, hinting at the involvement of this pathway in asymmetric NSC division [25,30,31,32].
Other studies have revealed that Ca2+ waves are propagated through connexins in the VZ of the embryonic brain to enhance RGC proliferation [23]. Since RGCs are situated close to the ventricular lumen, they are exposed to physical and chemical factors from both the surrounding tissue and the cerebrospinal fluid (CSF) in their basal and apical regions, respectively. The ventricular system originates from an open tube in the embryonic brain that gradually closes and rapidly expands due to the accumulation of CSF [33]. Consequently, spontaneous Ca2+ transients may be induced through mechanoreceptor activation via CSF circulation from a caudal to a rostral direction [21]. The positive hydrostatic pressure exerted by CSF stimulates the proliferation of neuroepithelial cells and RGCs through primary cilia, which are considered specialized organelles for Ca2+ signaling [34,35,36]. Interestingly, only approximately 56% of RGCs obtained from E16 mice exhibit Ca2+ transients in response to mechanical stimulation [37], suggesting heterogeneity among RGCs [38,39].
This observation also implies that 44% of RGCs do not exhibit mechanically stimulated Ca2+ transients, suggesting that intracellular Ca2+ transients in these populations may occur through ligand–receptor activation, leading to Ca2+ release from storage and its subsequent entry through membrane channels. In this regard, the signal transduction pathways of GPCRs or TKs are of interest, as several morphogens and growth factors act through them. Receptor activation leads to the hydrolysis of phosphatidylinositol 4,5-biphosphate (PIP2)-producing diacylglycerol (DAG) and inositol-triphosphate (IP3); IP3 predominantly binds to its receptor in the ER, resulting in Ca2+ release into the cytoplasm. Subsequently, as [Ca2+]c increases, Ca2+-sensitive proteins, such as CaM [19], play a pivotal role in transducing Ca2+ signals within the cell, controlling the transition from one phase of the cell cycle to the next by activating CaMKs, which phosphorylate target proteins involved in the cell cycle, such as cyclin-dependent kinases, essential for cell cycle progression. Furthermore, CaM can interact with growth factor receptors, such as the EGF receptor (EGFR), to modulate their activity and regulate cell proliferation [40]. In particular, CaM intervenes in the G1-S phase boundary, the transition from the G2 phase to mitosis (M), and the anaphase–telophase transition [37,41].
The involvement of CaM in the cell cycle is further supported by the use of antagonists, such as W13, W7, and trifluoperazine, or monoclonal antibodies, which promote cell cycle arrest and the inhibition of DNA synthesis [42,43,44]. Together with CaM, Ca2+ stimulates the expression of genes that lead to the activation of the cyclin-dependent kinases p33cdk2 and p34cdc2, which are necessary for the progression of G2 to M and control the signaling cascade that regulates the phosphorylation of the retinoblastoma protein, controlling a restriction point in G1-phase to enter the S-phase [45,46,47]. Furthermore, CaMKs and calcineurin can activate different signal pathways, such as those mediated by the nuclear factor of activated T-cells (NFAT), nuclear kappa-light-chain enhancer of activated B cells (NFκB), and stimulate transcription through the adenosine 3′,5′-cyclic monophosphate (cAMP) response element-binding protein (CREB) affecting cell proliferation and differentiation [48,49,50].
Calcyclin, a member of the S-100 family, also known as S100 Ca2+-Binding Protein A6 (S100A6), is expressed in various cell types, including neuron-like and glial-like cells in the pyramidal and molecular layers of the hippocampus at midgestation, in the entorhinal cortex throughout gestation, and in the fetal occipital cortex [51,52,53]. Its expression in glia-like cells, particularly in the entorhinal cortex and the occipital cortex in fetuses, suggests an important role in neural development.
In the adult brain, S100A6 is expressed in NSC and astrocyte precursors, indicating its potential involvement in generating astrocytes in the hippocampus [54]. Other studies reported its expression in epithelial cells and its overexpression in cancer epithelial cells, suggesting a role in cell proliferation [55,56,57,58]. Additionally, its effect on fibroblasts, where its deficiency prolongs the G0/G1 phase and leads to cell cycle withdrawal, further supports its involvement in the cell cycle [59]. These findings provide insights into the potential role of calcyclin in NSC neurogenesis and gliogenesis. However, further studies are needed to understand how this protein influences the NSC division pattern for self-renewal and differentiation.
It is important to mention that Ca2+ not only promotes or regulates the length of the cell cycle but also plays an essential role in the degradation of certain cyclins by activating the Ca2+-dependent protease calpain, thereby halting cell cycle progression [60].
As mentioned above, growth factors such as the EGF and FGF stimulate NSC proliferation through receptor activation, leading to increased cytosolic [Ca2+]c and CaMK activation. These kinases can phosphorylate and activate transcription factors involved in cell cycle progression and proliferation, such as myocyte enhancer factor 2 (MEF2) and NFAT, and modulate the activity of CDKs and their inhibitors, thus regulating NSC progression through different cell cycle phases [23,61,62,63].
In summary, Ca2+ entry into NSC is crucial for cell cycle regulation, and depending on the activated pathway, NSCs will either achieve self-renewal or differentiate into specific neural lineages. Therefore, changes in the transcriptional profiles of NSCs during CNS development will determine the effects of Ca2+ on these cells. Consequently, we will review the expression of membrane proteins such as GPCRs, transitory receptor potential channels (TRPs), RTKs, and connexins during development and their roles in NSC proliferation.

3. G-Protein-Coupled Receptors and Calcium Signaling

GPCRs constitute a large family of cell surface receptors that play a crucial role in transmitting signals from the extracellular environment to the inside of the cell through the hydrolysis of heteromeric guanine nucleotide-binding proteins (G-proteins). Upon the activation of GPCR, the liberation of the alpha subunit (αi, αs or αq) determines the downstream signaling pathway [64].
G-proteins are composed of the β/γ (Gβ/γ) dimmer and the α subunits, forming a trimeric protein complex. They are classified based on their α (Gα) subunit, which dictates their function upon receptor activation and promotes the disassembly of the trimetric protein. While the Gα subunits initiate the transduction signaling pathway downstream from receptor activation, the Gβ/γ subunit regulates the activity and stability of the Gα subunit, thereby contributing to GPCR-mediated signaling. The Gα subunits are further classified into Gαs, Gαi, Gαq11, and Gα12/13 subfamilies based on their primary sequence and function. Specifically, Gαs and Gαi stimulate and inhibit cAMP synthesis, respectively, while Gα12/13 regulates actin cytoskeletal remodeling by activating Rho GTPase, and Gαq11 activates phospholipase C, leading to intracellular Ca2+ release and entry [64].
Moreover, as mentioned previously, after IP3R binds to its receptors in the ER and Ca2+ depletion occurs, conformational changes in the Ca2+ sensor proteins, named stromal interaction molecules (STIM) in the ER membrane, are triggered. These changes activate membrane proteins such as Ca2+ channels, such as ORAI (1, 2, and 3) and TRPC1, facilitating store-operated Ca2+ entry (SOCE) and [65,66] (Figure 2). Consequently, the increased [Ca2+]c serves to refill the ER and activate Ca2+-sensitive proteins, including CaM, CaMKs, calcyclin, calcineurine, and Ca2+-dependent PKCs, all of which play critical roles in NSC proliferation. For instance, CaMKII promotes cell cycle progression, entry into the S phase, and neurogenesis while also phosphorylating PKC, predominantly affecting the G0/G1 and G2 phases [67,68].
Moreover, GPCRαq has been reported to influence NSC self-renewal and differentiation (Table 1). Its effect on NSC differentiation likely occurs through symmetric non-proliferative or asymmetric divisions.

4. Transitory Receptor Potential Channels and NSC Proliferation

The TRP ion channels comprise a large family of proteins characterized by six transmembrane domains and relatively long intracellular amino and carboxyl termini, encoded by 28 genes. These proteins form channels by arranging into tetramers (homotetramers or heterotetramers), where the S5 and S6 domains come together with the interconnecting loop to create the central pore [140,141]. Most TRPs are non-selective cation channels permeable to Na+, Ca2+, and Mg2+ and expressed in both excitable and non-excitable cells [142]. The mechanisms underlying the activation are poorly understood; however, they have been reported to be activated by voltage changes, receptor-operated activation, Ca2+-store depletion, ligands, and sensory responses to heat, cold, and pain [142,143].
The first member of the TRP superfamily was identified in Drosophila as a protein involved in phototransduction [144,145]. Since then, these proteins have been classified into seven subfamilies based on their homology to Drosophila’s photoreceptor: TRPC (canonical), TRPV (vanilloid), TRPM (melastatin), TRPP (polycystin), TRPML (mucolipin), TRPA (ankyrin), and TRPN (no mechanoreceptor potential C; nompC) [146,147]. TRP channels are implicated in several physiological processes, including cell proliferation, migration, survival, Ca2+ and Mg2+ homeostasis, neuronal growth, temperature sensation, and pain perception [142,148]. Here, we will discuss the TRP channels involved in the proliferation of NSCs.

4.1. TRPCs

TRPCs consist of seven members grouped into four subfamilies, all of which are activated downstream from GPCRs and intracellular Ca2+ release or DAG [147] (Figure 2). The first family comprises TRPC4 and TRPC5; the second includes only TRPC1; the third encompasses TRPC3, TRPC6, and TRPC7; and TRPC2 is the sole member of the fourth family; however, it is not expressed in humans as it is encoded by a pseudogene [149]. These channels are expressed early during embryo development and persist into adulthood, playing essential roles in neuronal development, including NSC proliferation, cerebellar granule cell survival, axon pathfinding, neuronal morphogenesis, and synaptogenesis [150].
Among the TRPC channels, TRPC1, TRPC3, TRPC6, and TRPC7 have been involved in cell proliferation [151,152,153,154,155,156]. Ca2+ mobilization through these channels occurs via store-operated and receptor-operated activation, following phospholipase C (PLC) activation downstream from Gαq/11PCRs and RTKs, which leads to DAG and IP3 production. DAG is required for the receptor-operated activation of at least TRPC3, TRPC6, and TRPC7, while IP3 causes Ca2+ release from the ER and the activation of TRPC1, a process triggered by store-operated channel activation. Furthermore, TPCR4 is activated via a receptor-operated mechanism by Gαi-PCRs (Figure 3) [157,158,159,160,161,162,163].
Strübing and colleagues (2003) reported greater expression of five TRPC proteins (TRPC1, 3, 4, 5, 6) in the embryonic compared with the adult brain. Furthermore, they proposed that TRPC1 plays an important role in forming heterotetramers composed of TRPC1/TRPC4/TRPC5 and TRPC1/TRPC3/TRPC6 [164]. Studies using smooth muscle and endothelial cells have shown that TRPC1 mediates Ca2+ influx induced by basic FGF (bFGF), thereby increasing cell proliferation [165,166]. Since bFGF plays an important role in cortical NSC proliferation in vitro and in vivo through FGF receptor-1 (FGFR-1) [167,168,169], it is likely that the mechanism reported in endothelial cells is also active in NSCs. Indeed, in the murine telencephalic neuroepithelium, TRPC1 and FGFR-1 are co-expressed in proliferating NSCs, and they have been coimmunoprecipitated from membrane extract preparations [164,169].
Furthermore, Ca2+ entrance and NSC proliferation induced by bFGF are reduced in TRPC1 knockdown NSCs [164,170], leading to cell cycle arrest in G0/G1 due to the overexpression of CaMK-II beta (Camk2b) and CDK inhibitor 2A (Cdkn2a) [171].
TRPC3 is also highly expressed in NSCs derived from mouse embryonic stem cells (mESC). Knocking out this channel in this cell type promotes impaired pluripotency, neural differentiation, and increased apoptosis, suggesting that TRPC3 activity participates in cell survival, the maintenance of pluripotency, and the transition of mESCs to NSCs through the disruption of the mitochondrial membrane potential in undifferentiated mESCs and mESCs undergoing neural differentiation. This reveals that TRPC3 is required for both early and late neural differentiation [153]. Moreover, it has recently been reported that the increased neuron differentiation of NSCs by ketamine is due to dramatic repression of the expression of TRPC3, partly by regulating the Glycogen synthase kinase 3β (GSK3β)/β-catenin pathway [172].
The above findings are relevant since medical practitioners and veterinarians use ketamine as an anesthetic during pregnancy or early after birth, which may affect brain development [173]. However, how ketamine regulates NSC differentiation and whether GSK3β participates in the ketamine-induced differentiation of NSCs is still unclear.
The participation of TRPC1 and TRPC3 in cell differentiation has also been observed after the knockdown of these channels in immortalized rat hippocampal cells (H19-7) derived from E17 Holtzman rat embryos, where cell differentiation is blocked. Another channel that is highly expressed in this cell type during proliferation and decreases dramatically upon differentiation is TRPC7 [173]. This suggests that TRPC1 and TRPC3 could be related to asymmetric division, while TRPC7 may be involved in the symmetric division of NSCs.

4.2. TRPVs

The TRPV family comprises six non-voltage-gated channel members named TRPV1 to TRPV6, which are further subdivided into the thermo-TRPV (TRPV1 to TRPV4) and the Ca2+-selective-TRPV (TRPV5 and TRPV6) subtypes [174]. In mammals, these proteins possess six ankyrins (Ank) repeat domains at the N-termini and a TRP-box in the C-termini. It has been proposed that the Ank repeat, TRPV1, and TRPV4 harbor an ATP-binding site, which might be important for channel activation and inactivation by CaM. Additionally, the TRP-box domain is believed to facilitate channel inactivation, gating, and protein tetramerization [175,176,177].
Mechanical stimulation, pH, and osmotic pressure changes are among the stimuli involved in its activation. The physiological functions of these channels vary depending on the TRPV subtype and the tissue in which they are expressed [178] (Figure 4). To our knowledge, TRPV1, TRPV5, and TRPV6 do not participate in NSC proliferation or CNS development, while TRPV2, TRPV3, and TRVP4 participate in these processes.
TRPV2 is a non-selective cation channel exhibiting Ca2+ permeability and highly expressed in the brain, lung, and spleen, where it contributes to Ca2+ homeostasis and macrophage activation [179]. It is activated by noxious heat (threshold of >52 °C), changes in osmolarity and membrane stretch, and growth factors, the latter causing PI-3K-dependent and -independent translocation to the plasma membrane [180,181].
Morgan and colleagues reported that ventral mesencephalic human NSCs exhibit high expressions of TRPV2 and TRPV3, which decline as differentiation progresses. Moreover, they suggested that these channels underline the mechanism for nearly all spontaneous Ca2+ activity in both proliferating and differentiating cells [182]. However, Santoni and collaborators found that the overexpression of TRPV2 in glioblastoma stem cells leads to the upregulation of GFAP and β-III tubulin levels, promotes differentiation both in vitro and in vivo, and reduces cell proliferation [183]. This suggests that the function of this channel may vary depending on the cellular type and context.
Although no effect of TRPV4 has been reported in NSCs, an interesting aspect of this channel is its association with the increased proliferation of oligodendrocyte progenitor cells (OPCs). The stimulation of rat OPCs with a selective TRPV4 agonist, GSK1016790A, induced a concentration-dependent increase in cell proliferation through Ca2+ influx via a PKC-dependent pathway. This effect was prevented by the TRPV4-antagonist, HC067047 [184].

4.3. TRPPs

TRPP is a family comprising two members: polycystin-1 (PC1) and polycystin-2 (PC2) [185]. The latter is closely related to TRPV1 and TRPV2, while PC1 is not considered to be part of the TRP family due to its molecular structure [186]. However, both PC1 and PC2 have been detected in the primary cilia of RGCs, where they appear to contribute to the planar cell polarity of late RGCs [187]. PC2 is situated in the plasma and ER membranes, where it interacts with TRPC1 and IP3R (Figure 5), thereby modulating intracellular Ca2+ signaling, for example, by prolonging the half-time decay of the IP3-induced Ca2+ transient [188,189,190].
Interestingly, these channels are expressed in mice at E12.5 in the developing cerebral cortex [191]. The knockdown of the expression of PC1 or PC2 in E13.5 primary cerebral cortical NSCs promotes an increase in proliferation and a decrease in neuron differentiation through a mechanism in which the Notch and STAT3 signaling are enhanced. The above is supported by the participation of this pathway in NSC self-renewal and blocking its differentiation [192]. Furthermore, reducing STAT3 expression leads to higher IPC production from NSCs [193]. Hence, it can be suggested that reducing PC1 and PC2 expression promotes symmetric cell division. During CNS development, these two channels may contribute to switching from symmetric to asymmetric NSC division and RGC transition to IPCs and, consequently, neuron differentiation.

5. Receptor Tyrosine Kinases

RTKs are cell surface receptors with intrinsic enzymatic activity regulated via ligand binding. Their molecular structure consists of an N-terminal extracellular domain for ligand recognition, a single transmembrane domain, and an intracellular kinase domain followed by a largely C-terminal tail region, which generates regulatory signals [194].
Ligand-activated RTKs stimulate many cellular processes, including protein synthesis, cell cycle progression, DNA synthesis, and cellular replication. Each RTK unleashes a specific set of responses through ligand binding that leads to the cross-phosphorylation of the RTK and the subsequent recruitment and activation of intracellular proteins, which transmit their signal to a series of targets, such as cyclins, PKC, phosphatidylinositol-3-kinase (PI-3K), GTPase, rat sarcoma virus/mitogen-activated protein kinase (Ras/MAPK), phospholipase A2 (PLA2), and transcription factors [194].
Growth factors such as PDGF, EGF, bFGF, and IGF-I promote an increase in [Ca2+]c [195,196,197,198]. In contrast with GPCR activation, RTK stimulation induces a long-lasting Ca2+ elevation that activates a specific pattern of genes, different from those regulated by the brief spikes caused by the release of Ca2+ from the RE and SOCE.
Several transduction pathways can be activated downstream from RTKs, and it is well established that these promote cell proliferation. However, the only pathway related to Ca2+ dynamics is the one associated with PLCγ activation, which promotes IP3 and DAG production (Figure 6). Recent studies suggest that PLCγ1 plays an important role in CNS development through ligand–receptor activation. However, PLCγ1 has been linked to neurite outgrowth, cell migration, axon guidance, and synaptic plasticity in vitro, and there is limited evidence linking it to NSC proliferation [199,200,201,202,203].
There is no doubt that RTK activation leads to NSC proliferation (Table 2). However, it is likely that ligand-binding activation may differentially affect cell division patterns in a spatiotemporal manner for proper CNS development. Interestingly, the PLCγ1 pathway regulates the dynamics of the actin cytoskeleton through the production of IP3 and DAG, which promote Ca2+ release from intracellular storage and Ca2+ influx, respectively [204,205].
The regulation of actin-cytoskeleton allows neuronal cells to undergo morphological changes during mitosis, cell polarity, neuronal processes extension, cell migration, synapse formation, and axon pathfinding. Interestingly, changes in the length of the basal processes of RGCs and the position of their cellular body occur during cell division, and the dynamics of actin cytoskeleton for cell proliferation are related to transducing signals from the cell surface to the nucleus, thereby regulating gene expression and cell cycle progression. Furthermore, actin filaments are essential for forming the contractile ring during cytokinesis and the regulation of cyclase-associated proteins that coordinate actin polymerization and depolymerization, which influence cell cycle progression and cell division patterns [9,206,207,208].
Despite differences in ligands, amino acid sequences, and structures, almost all RTKs can activate the same signaling pathways, which affect cell proliferation. Cellular function is likely to depend on receptor density, the coactivation of RTKs, and cross-talk with GPCRs, which selectively influence gene expression and cell function.
Table 2. RTKs that affect neural stem cell proliferation.
Table 2. RTKs that affect neural stem cell proliferation.
RTK Role Ligands/Agonists Antagonists
EGFR (ErbB)Increases NSC proliferation, migration, and survival in vivo and in vitro [209,210,211].
Promotes oligodendrocyte differentiation of NSCs in vivo [198,212].
The knockout mice show the atrophy of the anterior cerebral cortex in vivo [213]
EGF, TGF-α [209],
Amphiregulin, Betacellulin,
Heparin-binding EGF-like growth factor [214]
Epiregulin, Epigen [215], Neuroregulins [211]
Cetuximab, Panitumumab [214],
Trastuzumab, Pertuzumab, ABX-EGF, EMD-7200, h-R3, ICR-62, ZD1839 (Gefitinib Iressa), OSI-774 (Erlotinib), Lapatinib (GW572016, GW2016), EKB-569, AEE788, BMS-599626, AZD 9291, Dacomitinib, Afatinib, CO-1686, Neratinib, Canertinib, AC-480, AZD 8931, AST 1306 [216]
FGFR1Maintains the self-renewal of NSCs [217]. NSC proliferation and neurogenesis in the developing cerebral cortex [169]
Its deletion, together with FGFR2 and FGFR3, leads to Foxg1-positive precursors telencephalic cell death, resulting in the loss of the basal ganglia and cortex in vivo [218].
FGF-1, -2, -4, -6, -7, -8, -10, -16, -17, -18, -22 [219]Derazantinib (ARQ087, ASP5878, AZD4547), Infigratinib (BGJ398), Debio-1347, Dovitinib [220,221], Brivanib (BMS-582664), BMS-540215, E-3810 (AL3810), NP603, LY2874455, Fisogatinib (CH518328, Debio 1347, E7090), Rogaratinib (BAY1163877), Futibatinib (TAS-120), Pemigatinib (INCB054828), Erdafitinib (JNJ-42756493) [221]
PDGFRαThe knocking down or blocking antibodies of PDGFRα suppresses the proliferation of NSCs and increases the cell death rate [222].PDGF-A, PDGF-B, PDGF-AB, PDGF-C [223]Dasatinib, Masitinib [224],
Axitinib [225],
Sorafenib, Pazopanib, Cediranib [225,226]
Imatinib,
Sunitinib, Nilotinib [224,225,226]
IGF-1R CThe knocking down of the receptor reduces NSC proliferation, stunts brain growth, and decreases the neuronal number [227].IGF-II, Insulin, IGF-I (Somatomedin C) [228]NVP-ADW742, α-IR3, JB-1 [229]
NVP-AEW541 [230],
MAB391, OSI-906 [231]
TrkB (Ntrk2)Promotes NSC survival and proliferation [232] in cortical precursors in vivo and promotes proliferation and enhanced neurogenesis [233].BDNF [232,233], NT-4, NT-3 [234], L-783,281 [235]
Amitriptyline, 7,8-Dihydroxyflavone,
Deoxygedunin, Paecilomycine A [236]
Larotrectinib, Entrectinib,
Selitrectinib (LOXO-195), Altiratinib, DS-6051b, Lestaurtinib, Merestinib,
MGCD516, PLX7486, ONO-5390556, TPX-0005, Repotrectinib [237]
TrkC (Ntrk3)Increases neuron differentiation in vitro [238] and NSC proliferation in vitro and in vivo [233].NT-3 [234,238], L-783,281 [235]Larotrectinib, Entrectinib, Selitrectinib (LOXO-195), Repotrectinib, Altiratinib, Crizotinib, DS-6051b, Lestaurtinib, Merestinib, MGCD516, TSR-011, ONO-5390556, TPX-0005 [237]
Ntrk2: Neurotrophic Receptor Tyrosine Kinase 2; Ntrk3: Neurotrophic Receptor Tyrosine Kinase 3.

6. Gap Junctions

Connexins are proteins with four transmembrane domains named based on their predicted molecular weight. These molecules form hemichannels composed of six connexins which together form a connexon. The union between connexons in different cells forms a gap junction, allowing the direct cytoplasmic exchange in low-weight molecules and ions between adjacent cells [239]. In total, 20 and 21 types of connexins have been reported in mice and humans, respectively [240]. In the developing cerebral cortex of mice (E14–E18), Cx26, Cx36, Cx37, Cx43, and Cx45 are highly and differentially expressed [241].
In NSCs, connexins play several crucial roles, including cell–cell communication, cell synchronization, cell cycle progression, and niche maintenance. Connexins expressed during late corticogenesis in all cortical layers include Cx26 and Cx37, while those abundant in the VZ are Cx36, Cx43, and Cx45 [241]. Therefore, the latter are likely to play an important role in NSC proliferation. Moreover, their expression varies according to the phase of the cell cycle and developmental stage. The pharmacological blocking of connexins prevents DNA synthesis in NSCs [23,242,243,244,245]. Notably, reducing Cx43 translation using short hairpin technology decreases spontaneous Ca2+ activity and the rate of RGC proliferation, thereby reducing NSC and IP pools [23,246].
Spontaneous Ca2+ activity during early corticogenesis (E15) in rats propagates in clusters of RGCs within the VZ. As development progresses and neurons migrate, this Ca2+ pattern decreases [247]. Ca2+ waves occur through gap junctions, initiated by ATP binding to the GPCR subtype P2Y1, coupled to a Gαq/11 protein. This leads to the activation of PLCγ, IP3 synthesis, and intracellular Ca2+ release due to IP3R activation in the ER membrane [23,248,249].
As mentioned above, Cx43 is one of the most prominent connexins implicated in NSC proliferation (Figure 7). Cx43 forms gap junction channels that directly connect the cytoplasm of adjacent cells, essential for coordinating NSC proliferation within their niche. The pharmacological blocking of gap junctions or knockdown Cx43 stunts the nuclear migration of the S/G2-phase cells in the upper strata of the VZ, while the knockout of this connexin disorganized the VZ/SVZ and promoted deficiencies in neuronal migration in mice [245,250].
Furthermore, Cx43 facilitates the integration of environmental signals within the NSC niche by allowing NSCs to communicate with each other and responding to growth factors, neurotransmitters, and other cues influencing proliferation. Although Cx43 appears particularly important for NSC proliferation, other connexins, such as Cx30 and Cx45, may also contribute to cell–cell communication within NSC populations [251]. However, their specific roles in regulating NSC proliferation are less characterized than Cx43.
Figure 7. Radial glia cell cycle and conexin-43. (A) Scheme showing the level of expression of Cx43 during the cell cycle and gap junction. During the S Phase, Cx43 is highly expressed, and RGCs are coupled; as the cell cycle progresses, Cx43 decreases its expression, and RGCs are less coupled, while in the M phase, cells are uncoupled. (B) In the S phase, radial glia initiates Ca2+ waves by releasing ATP, which binds to P2Y1 receptors in the membrane of adjacent cells, inducing IP3 Ca2+ release from the endoplasmic reticulum (ER). Created using BioRender.com based on [23,242,252].
Figure 7. Radial glia cell cycle and conexin-43. (A) Scheme showing the level of expression of Cx43 during the cell cycle and gap junction. During the S Phase, Cx43 is highly expressed, and RGCs are coupled; as the cell cycle progresses, Cx43 decreases its expression, and RGCs are less coupled, while in the M phase, cells are uncoupled. (B) In the S phase, radial glia initiates Ca2+ waves by releasing ATP, which binds to P2Y1 receptors in the membrane of adjacent cells, inducing IP3 Ca2+ release from the endoplasmic reticulum (ER). Created using BioRender.com based on [23,242,252].
Ijms 25 04073 g007

7. Conclusions

It is evident that NSC proliferation and [Ca2+]c levels are closely related. Therefore, the regulation of cell proliferation and differentiation during brain development must be finely tuned to ensure correct cytoarchitecture and function development. As development progresses, molecules involved in Ca2+ dynamics exhibit changes in expression, impacting molecule–cell interactions and downstream signaling pathways in a temporally and spatially different manner. Furthermore, studying NSCs in vivo and in vitro presents challenges due to the diversity of NSC populations coexisting with progenitor-committed and differentiated cells as development progresses.
By providing an overview of the role of Ca2+ signaling in NSC proliferation and differentiation, we aim to deepen our understanding of the intricate relationship between Ca2+ signaling and its dynamics in NSCs, particularly focusing on proliferation and differentiation process. Through in-depth exploration of the role of RGCs and the impact of Ca2+ fluctuations on cell cycle regulation, we underscore the complex mechanisms governing CNS development. By synthesizing findings from several studies, we emphasize the significance of Ca2+ signaling pathways in orchestrating the fates of NCSs, with potential implications for understanding neurodevelopmental processes and devising therapeutic interventions for neurological disorders originating in fetal development. This review underscores the expanding knowledge in this field and the importance of further research to unravel the nuances of Ca2+-mediated signaling in NSC biology.

Author Contributions

D.A.D.-P., N.R.-R. and G.G.-L. manuscript draft and images; N.F.D. and A.M.-H. participated in bibliography selection, manuscript correction, final image selection, and manuscript editing. All authors actively participated in the writing of the draft and final manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

A.M-H’s research is funded by the “Instituto Nacional de Perinatología” (protocol number 2020-1-9).

Informed Consent Statement

Not applicable.

Data Availability Statement

No new data were created or analyzed in this study. Data sharing is not applicable to this article.

Conflicts of Interest

The authors declare no conflicts of interest in the writing of the manuscript.

References

  1. Noctor, S.C.; Martinez-Cerdeno, V.; Ivic, L.; Kriegstein, A.R. Cortical neurons arise in symmetric and asymmetric division zones and migrate through specific phases. Nat. Neurosci. 2004, 7, 136–144. [Google Scholar] [CrossRef] [PubMed]
  2. Noctor, S.C.; Martinez-Cerdeno, V.; Kriegstein, A.R. Distinct behaviors of neural stem and progenitor cells underlie cortical neurogenesis. J. Comp. Neurol. 2008, 508, 28–44. [Google Scholar] [CrossRef] [PubMed]
  3. Kriegstein, A.; Alvarez-Buylla, A. The glial nature of embryonic and adult neural stem cells. Annu. Rev. Neurosci. 2009, 32, 149–184. [Google Scholar] [CrossRef]
  4. Lodato, S.; Arlotta, P. Generating neuronal diversity in the mammalian cerebral cortex. Annu. Rev. Cell Dev. Biol. 2015, 31, 699–720. [Google Scholar] [CrossRef] [PubMed]
  5. Jiang, X.; Nardelli, J. Cellular and molecular introduction to brain development. Neurobiol. Dis. 2016, 92, 3–17. [Google Scholar] [CrossRef] [PubMed]
  6. Kriegstein, A.R.; Gotz, M. Radial glia diversity: A matter of cell fate. Glia 2003, 43, 37–43. [Google Scholar] [CrossRef]
  7. Tan, X.; Shi, S.H. Neocortical neurogenesis and neuronal migration. Wiley Interdiscip. Rev. Dev. Biol. 2013, 2, 443–459. [Google Scholar] [CrossRef]
  8. Gotz, M.; Huttner, W.B. The cell biology of neurogenesis. Nat. Rev. Mol. Cell Biol. 2005, 6, 777–788. [Google Scholar] [CrossRef]
  9. Huttner, W.B.; Kosodo, Y. Symmetric versus asymmetric cell division during neurogenesis in the developing vertebrate central nervous system. Curr. Opin. Cell Biol. 2005, 17, 648–657. [Google Scholar] [CrossRef]
  10. Berridge, M.J.; Bootman, M.D.; Roderick, H.L. Calcium signalling: Dynamics, homeostasis and remodelling. Nat. Rev. Mol. Cell Biol. 2003, 4, 517–529. [Google Scholar] [CrossRef]
  11. Leibovitz, Z.; Lerman-Sagie, T.; Haddad, L. Fetal Brain Development: Regulating Processes and Related Malformations. Life 2022, 12, 809. [Google Scholar] [CrossRef]
  12. Owens, D.F.; Flint, A.C.; Dammerman, R.S.; Kriegstein, A.R. Calcium dynamics of neocortical ventricular zone cells. Dev. Neurosci. 2000, 22, 25–33. [Google Scholar] [CrossRef] [PubMed]
  13. Owens, D.F.; Kriegstein, A.R. Patterns of intracellular calcium fluctuation in precursor cells of the neocortical ventricular zone. J. Neurosci. 1998, 18, 5374–5388. [Google Scholar] [CrossRef]
  14. Bayer, S.A.; Altman, J. Development of the endopiriform nucleus and the claustrum in the rat brain. Neuroscience 1991, 45, 391–412. [Google Scholar] [CrossRef] [PubMed]
  15. Takahashi, T.; Nowakowski, R.S.; Caviness, V.S., Jr. Mode of cell proliferation in the developing mouse neocortex. Proc. Natl. Acad. Sci. USA 1994, 91, 375–379. [Google Scholar] [CrossRef] [PubMed]
  16. Takahashi, T.; Nowakowski, R.S.; Caviness, V.S., Jr. The leaving or Q fraction of the murine cerebral proliferative epithelium: A general model of neocortical neuronogenesis. J. Neurosci. 1996, 16, 6183–6196. [Google Scholar] [CrossRef]
  17. Barrack, D.S.; Thul, R.; Owen, M.R. Modelling the coupling between intracellular calcium release and the cell cycle during cortical brain development. J. Theor. Biol. 2014, 347, 17–32. [Google Scholar] [CrossRef]
  18. Saran, S. Calcium levels during cell cycle correlate with cell fate of Dictyostelium discoideum. Cell Biol. Int. 1999, 23, 399–405. [Google Scholar] [CrossRef]
  19. Humeau, J.; Bravo-San Pedro, J.M.; Vitale, I.; Nunez, L.; Villalobos, C.; Kroemer, G.; Senovilla, L. Calcium signaling and cell cycle: Progression or death. Cell Calcium 2018, 70, 3–15. [Google Scholar] [CrossRef]
  20. MacKrill, J.J. Protein-protein interactions in intracellular Ca2+-release channel function. Biochem. J. 1999, 337 Pt 3, 345–361. [Google Scholar] [CrossRef]
  21. Guirao, B.; Meunier, A.; Mortaud, S.; Aguilar, A.; Corsi, J.M.; Strehl, L.; Hirota, Y.; Desoeuvre, A.; Boutin, C.; Han, Y.G.; et al. Coupling between hydrodynamic forces and planar cell polarity orients mammalian motile cilia. Nat. Cell Biol. 2010, 12, 341–350. [Google Scholar] [CrossRef] [PubMed]
  22. Kahl, C.R.; Means, A.R. Regulation of cell cycle progression by calcium/calmodulin-dependent pathways. Endocr. Rev. 2003, 24, 719–736. [Google Scholar] [CrossRef] [PubMed]
  23. Weissman, T.A.; Riquelme, P.A.; Ivic, L.; Flint, A.C.; Kriegstein, A.R. Calcium waves propagate through radial glial cells and modulate proliferation in the developing neocortex. Neuron 2004, 43, 647–661. [Google Scholar] [CrossRef] [PubMed]
  24. Zhao, W.Q.; Alkon, D.L.; Ma, W. c-Src protein tyrosine kinase activity is required for muscarinic receptor-mediated DNA synthesis and neurogenesis via ERK1/2 and c-AMP-responsive element-binding protein signaling in neural precursor cells. J. Neurosci. Res. 2003, 72, 334–342. [Google Scholar] [CrossRef] [PubMed]
  25. Resende, R.R.; Gomes, K.N.; Adhikari, A.; Britto, L.R.; Ulrich, H. Mechanism of acetylcholine-induced calcium signaling during neuronal differentiation of P19 embryonal carcinoma cells in vitro. Cell Calcium 2008, 43, 107–121. [Google Scholar] [CrossRef] [PubMed]
  26. Berridge, M.J.; Lipp, P.; Bootman, M.D. The versatility and universality of calcium signalling. Nat. Rev. Mol. Cell Biol. 2000, 1, 11–21. [Google Scholar] [CrossRef] [PubMed]
  27. Uhlen, P.; Fritz, N.; Smedler, E.; Malmersjo, S.; Kanatani, S. Calcium signaling in neocortical development. Dev. Neurobiol. 2015, 75, 360–368. [Google Scholar] [CrossRef] [PubMed]
  28. Takahashi, T.; Nowakowski, R.S.; Caviness, V.S., Jr. The cell cycle of the pseudostratified ventricular epithelium of the embryonic murine cerebral wall. J. Neurosci. 1995, 15, 6046–6057. [Google Scholar] [CrossRef] [PubMed]
  29. Calegari, F.; Huttner, W.B. An inhibition of cyclin-dependent kinases that lengthens, but does not arrest, neuroepithelial cell cycle induces premature neurogenesis. J. Cell Sci. 2003, 116, 4947–4955. [Google Scholar] [CrossRef]
  30. Molina-Hernandez, A.; Rodriguez-Martinez, G.; Escobedo-Avila, I.; Velasco, I. Histamine up-regulates fibroblast growth factor receptor 1 and increases FOXP2 neurons in cultured neural precursors by histamine type 1 receptor activation: Conceivable role of histamine in neurogenesis during cortical development in vivo. Neural Dev. 2013, 8, 4. [Google Scholar] [CrossRef]
  31. Marquez-Valadez, B.; Aquino-Miranda, G.; Quintero-Romero, M.O.; Papacostas-Quintanilla, H.; Bueno-Nava, A.; Lopez-Rubalcava, C.; Diaz, N.F.; Arias-Montano, J.A.; Molina-Hernandez, A. The Systemic Administration of the Histamine H1 Receptor Antagonist/Inverse Agonist Chlorpheniramine to Pregnant Rats Impairs the Development of Nigro-Striatal Dopaminergic Neurons. Front. Neurosci. 2019, 13, 360. [Google Scholar] [CrossRef] [PubMed]
  32. Molina-Hernandez, A.; Velasco, I. Histamine induces neural stem cell proliferation and neuronal differentiation by activation of distinct histamine receptors. J. Neurochem. 2008, 106, 706–717. [Google Scholar] [CrossRef] [PubMed]
  33. Chau, K.F.; Springel, M.W.; Broadbelt, K.G.; Park, H.Y.; Topal, S.; Lun, M.P.; Mullan, H.; Maynard, T.; Steen, H.; LaMantia, A.S.; et al. Progressive Differentiation and Instructive Capacities of Amniotic Fluid and Cerebrospinal Fluid Proteomes following Neural Tube Closure. Dev. Cell 2015, 35, 789–802. [Google Scholar] [CrossRef] [PubMed]
  34. Delling, M.; DeCaen, P.G.; Doerner, J.F.; Febvay, S.; Clapham, D.E. Primary cilia are specialized calcium signalling organelles. Nature 2013, 504, 311–314. [Google Scholar] [CrossRef]
  35. Desmond, M.E.; Levitan, M.L. Brain expansion in the chick embryo initiated by experimentally produced occlusion of the spinal neurocoel. Anat. Rec. 2002, 268, 147–159. [Google Scholar] [CrossRef]
  36. Park, M.G.; Jang, H.; Lee, S.H.; Lee, C.J. Flow Shear Stress Enhances the Proliferative Potential of Cultured Radial Glial Cells Possibly Via an Activation of Mechanosensitive Calcium Channel. Exp. Neurobiol. 2017, 26, 71–81. [Google Scholar] [CrossRef]
  37. Rasmussen, C.D.; Means, A.R. Calmodulin is involved in regulation of cell proliferation. EMBO J. 1987, 6, 3961–3968. [Google Scholar] [CrossRef]
  38. Malatesta, P.; Hartfuss, E.; Gotz, M. Isolation of radial glial cells by fluorescent-activated cell sorting reveals a neuronal lineage. Development 2000, 127, 5253–5263. [Google Scholar] [CrossRef]
  39. Eze, U.C.; Bhaduri, A.; Haeussler, M.; Nowakowski, T.J.; Kriegstein, A.R. Single-cell atlas of early human brain development highlights heterogeneity of human neuroepithelial cells and early radial glia. Nat. Neurosci. 2021, 24, 584–594. [Google Scholar] [CrossRef]
  40. Chin, D.; Means, A.R. Calmodulin: A prototypical calcium sensor. Trends Cell Biol. 2000, 10, 322–328. [Google Scholar] [CrossRef]
  41. Rasmussen, C.D.; Means, A.R. Calmodulin is required for cell-cycle progression during G1 and mitosis. EMBO J. 1989, 8, 73–82. [Google Scholar] [CrossRef] [PubMed]
  42. Chafouleas, J.G.; Bolton, W.E.; Hidaka, H.; Boyd, A.E., 3rd; Means, A.R. Calmodulin and the cell cycle: Involvement in regulation of cell-cycle progression. Cell 1982, 28, 41–50. [Google Scholar] [CrossRef] [PubMed]
  43. Chafouleas, J.G.; Lagace, L.; Bolton, W.E.; Boyd, A.E., 3rd; Means, A.R. Changes in calmodulin and its mRNA accompany reentry of quiescent (G0) cells into the cell cycle. Cell 1984, 36, 73–81. [Google Scholar] [CrossRef] [PubMed]
  44. Sasaki, Y.; Hidaka, H. Calmodulin and cell proliferation. Biochem. Biophys. Res. Commun. 1982, 104, 451–456. [Google Scholar] [CrossRef] [PubMed]
  45. Takuwa, N.; Zhou, W.; Kumada, M.; Takuwa, Y. Ca2+-dependent stimulation of retinoblastoma gene product phosphorylation and p34cdc2 kinase activation in serum-stimulated human fibroblasts. J. Biol. Chem. 1993, 268, 138–145. [Google Scholar] [CrossRef] [PubMed]
  46. Takuwa, N.; Zhou, W.; Takuwa, Y. Calcium, calmodulin and cell cycle progression. Cell. Signal. 1995, 7, 93–104. [Google Scholar] [CrossRef] [PubMed]
  47. Takuwa, N.; Zhou, W.; Kumada, M.; Takuwa, Y. Ca2+/calmodulin is involved in growth factor-induced retinoblastoma gene product phosphorylation in human vascular endothelial cells. FEBS Lett. 1992, 306, 173–175. [Google Scholar] [CrossRef] [PubMed]
  48. Hogan, P.G.; Chen, L.; Nardone, J.; Rao, A. Transcriptional regulation by calcium, calcineurin, and NFAT. Genes Dev. 2003, 17, 2205–2232. [Google Scholar] [CrossRef] [PubMed]
  49. Sun, P.; Enslen, H.; Myung, P.S.; Maurer, R.A. Differential activation of CREB by Ca2+/calmodulin-dependent protein kinases type II and type IV involves phosphorylation of a site that negatively regulates activity. Genes Dev. 1994, 8, 2527–2539. [Google Scholar] [CrossRef]
  50. Ma, H.; Groth, R.D.; Cohen, S.M.; Emery, J.F.; Li, B.; Hoedt, E.; Zhang, G.; Neubert, T.A.; Tsien, R.W. γCaMKII shuttles Ca2+/CaM to the nucleus to trigger CREB phosphorylation and gene expression. Cell 2014, 159, 281–294. [Google Scholar] [CrossRef]
  51. Tiu, S.C.; Chan, W.Y.; Heizmann, C.W.; Schafer, B.W.; Shu, S.Y.; Yew, D.T. Differential expression of S100B and S100A61 in the human fetal and aged cerebral cortex. Brain Res. Dev. Brain Res. 2000, 119, 159–168. [Google Scholar] [CrossRef] [PubMed]
  52. Yamashita, N.; Kosaka, K.; Ilg, E.C.; Schafer, B.W.; Heizmann, C.W.; Kosaka, T. Selective association of S100A6 (calcyclin)-immunoreactive astrocytes with the tangential migration pathway of subventricular zone cells in the rat. Brain Res. 1997, 778, 388–392. [Google Scholar] [CrossRef] [PubMed]
  53. Yamashita, N.; Ilg, E.C.; Schafer, B.W.; Heizmann, C.W.; Kosaka, T. Distribution of a specific calcium-binding protein of the S100 protein family, S100A6 (calcyclin), in subpopulations of neurons and glial cells of the adult rat nervous system. J. Comp. Neurol. 1999, 404, 235–257. [Google Scholar] [CrossRef]
  54. Yamada, J.; Jinno, S. S100A6 (calcyclin) is a novel marker of neural stem cells and astrocyte precursors in the subgranular zone of the adult mouse hippocampus. Hippocampus 2014, 24, 89–101. [Google Scholar] [CrossRef] [PubMed]
  55. Kuznicki, J.; Kordowska, J.; Puzianowska, M.; Wozniewicz, B.M. Calcyclin as a marker of human epithelial cells and fibroblasts. Exp. Cell Res. 1992, 200, 425–430. [Google Scholar] [CrossRef] [PubMed]
  56. Weterman, M.A.; Stoopen, G.M.; van Muijen, G.N.; Kuznicki, J.; Ruiter, D.J.; Bloemers, H.P. Expression of calcyclin in human melanoma cell lines correlates with metastatic behavior in nude mice. Cancer Res. 1992, 52, 1291–1296. [Google Scholar] [PubMed]
  57. Weterman, M.A.; van Muijen, G.N.; Bloemers, H.P.; Ruiter, D.J. Expression of calcyclin in human melanocytic lesions. Cancer Res. 1993, 53, 6061–6066. [Google Scholar] [PubMed]
  58. Hirschhorn, R.R.; Aller, P.; Yuan, Z.A.; Gibson, C.W.; Baserga, R. Cell-cycle-specific cDNAs from mammalian cells temperature sensitive for growth. Proc. Natl. Acad. Sci. USA 1984, 81, 6004–6008. [Google Scholar] [CrossRef] [PubMed]
  59. Slomnicki, L.P.; Lesniak, W. S100A6 (calcyclin) deficiency induces senescence-like changes in cell cycle, morphology and functional characteristics of mouse NIH 3T3 fibroblasts. J. Cell. Biochem. 2010, 109, 576–584. [Google Scholar] [CrossRef]
  60. Santella, L.; Ercolano, E.; Nusco, G.A. The cell cycle: A new entry in the field of Ca2+ signaling. Cell. Mol. Life Sci. 2005, 62, 2405–2413. [Google Scholar] [CrossRef]
  61. Li, Z.; McKercher, S.R.; Cui, J.; Nie, Z.; Soussou, W.; Roberts, A.J.; Sallmen, T.; Lipton, J.H.; Talantova, M.; Okamoto, S.; et al. Myocyte enhancer factor 2C as a neurogenic and antiapoptotic transcription factor in murine embryonic stem cells. J. Neurosci. 2008, 28, 6557–6568. [Google Scholar] [CrossRef]
  62. Urso, K.; Fernandez, A.; Velasco, P.; Cotrina, J.; de Andres, B.; Sanchez-Gomez, P.; Hernandez-Lain, A.; Hortelano, S.; Redondo, J.M.; Cano, E. NFATc3 controls tumour growth by regulating proliferation and migration of human astroglioma cells. Sci. Rep. 2019, 9, 9361. [Google Scholar] [CrossRef]
  63. Gonzalez, L.; Domingo-Muelas, A.; Duart-Abadia, P.; Nunez, M.; Mikolcevic, P.; Llonch, E.; Cubillos-Rojas, M.; Canovas, B.; Forrow, S.M.A.; Morante-Redolat, J.M.; et al. The atypical CDK activator RingoA/Spy1 regulates exit from quiescence in neural stem cells. iScience 2023, 26, 106202. [Google Scholar] [CrossRef]
  64. Lynch, J.R.; Wang, J.Y. G Protein-Coupled Receptor Signaling in Stem Cells and Cancer. Int. J. Mol. Sci. 2016, 17, 707. [Google Scholar] [CrossRef] [PubMed]
  65. Taylor, C.W. Store-operated Ca2+ entry: A STIMulating stOrai. Trends Biochem. Sci. 2006, 31, 597–601. [Google Scholar] [CrossRef]
  66. Berridge, M.J. Inositol trisphosphate and calcium signalling mechanisms. Biochim. Biophys. Acta 2009, 1793, 933–940. [Google Scholar] [CrossRef]
  67. Clapham, D.E. Calcium signaling. Cell 2007, 131, 1047–1058. [Google Scholar] [CrossRef]
  68. Black, A.R.; Black, J.D. Protein kinase C signaling and cell cycle regulation. Front. Immunol. 2012, 3, 423. [Google Scholar] [CrossRef]
  69. Xing, L.; Kalebic, N.; Namba, T.; Vaid, S.; Wimberger, P.; Huttner, W.B. Serotonin Receptor 2A Activation Promotes Evolutionarily Relevant Basal Progenitor Proliferation in the Developing Neocortex. Neuron 2020, 108, 1113–1129.e6. [Google Scholar] [CrossRef] [PubMed]
  70. Bortolotto, V.; Mancini, F.; Mangano, G.; Salem, R.; Xia, E.; Del Grosso, E.; Bianchi, M.; Canonico, P.L.; Polenzani, L.; Grilli, M. Proneurogenic Effects of Trazodone in Murine and Human Neural Progenitor Cells. ACS Chem. Neurosci. 2017, 8, 2027–2038. [Google Scholar] [CrossRef] [PubMed]
  71. Casey, A.B.; Cui, M.; Booth, R.G.; Canal, C.E. “Selective” serotonin 5-HT2A receptor antagonists. Biochem. Pharmacol. 2022, 200, 115028. [Google Scholar] [CrossRef]
  72. Nelson, D.L.; Lucaites, V.L.; Wainscott, D.B.; Glennon, R.A. Comparisons of hallucinogenic phenylisopropylamine binding affinities at cloned human 5-HT2A, 5-HT2B and 5-HT2C receptors. Naunyn Schmiedebergs Arch. Pharmacol. 1999, 359, 1–6. [Google Scholar] [CrossRef]
  73. Jensen, A.A.; McCorvy, J.D.; Leth-Petersen, S.; Bundgaard, C.; Liebscher, G.; Kenakin, T.P.; Brauner-Osborne, H.; Kehler, J.; Kristensen, J.L. Detailed Characterization of the In Vitro Pharmacological and Pharmacokinetic Properties of N-(2-Hydroxybenzyl)-2,5-dimethoxy-4-cyanophenylethylamine (25CN-NBOH), a Highly Selective and Brain-Penetrant 5-HT2A Receptor Agonist. J. Pharmacol. Exp. Ther. 2017, 361, 441–453. [Google Scholar] [CrossRef]
  74. Vamvakopoulou, I.A.; Narine, K.A.D.; Campbell, I.; Dyck, J.R.B.; Nutt, D.J. Mescaline: The forgotten psychedelic. Neuropharmacology 2023, 222, 109294. [Google Scholar] [CrossRef]
  75. Gardell, L.R.; Vanover, K.E.; Pounds, L.; Johnson, R.W.; Barido, R.; Anderson, G.T.; Veinbergs, I.; Dyssegaard, A.; Brunmark, P.; Tabatabaei, A.; et al. ACP-103, a 5-hydroxytryptamine 2A receptor inverse agonist, improves the antipsychotic efficacy and side-effect profile of haloperidol and risperidone in experimental models. J. Pharmacol. Exp. Ther. 2007, 322, 862–870. [Google Scholar] [CrossRef]
  76. Millan, M.J.; Schreiber, R.; Dekeyne, A.; Rivet, J.M.; Bervoets, K.; Mavridis, M.; Sebban, C.; Maurel-Remy, S.; Newman-Tancredi, A.; Spedding, M.; et al. S 16924 ((R)-2-[1-[2-(2,3-dihydro-benzo[1,4] dioxin-5-yloxy)-ethyl]-pyrrolidin-3yl]-1-(4-fluoro-phenyl)-ethanone), a novel, potential antipsychotic with marked serotonin (5-HT)1A agonist properties: II. Functional profile in comparison to clozapine and haloperidol. J. Pharmacol. Exp. Ther. 1998, 286, 1356–1373. [Google Scholar]
  77. Shapiro, D.A.; Kristiansen, K.; Kroeze, W.K.; Roth, B.L. Differential modes of agonist binding to 5-hydroxytryptamine2A serotonin receptors revealed by mutation and molecular modeling of conserved residues in transmembrane region 5. Mol. Pharmacol. 2000, 58, 877–886. [Google Scholar] [CrossRef]
  78. Fiorino, F.; Magli, E.; Kedzierska, E.; Ciano, A.; Corvino, A.; Severino, B.; Perissutti, E.; Frecentese, F.; Di Vaio, P.; Saccone, I.; et al. New 5-HT1A, 5HT2A and 5HT2C receptor ligands containing a picolinic nucleus: Synthesis, in vitro and in vivo pharmacological evaluation. Bioorg Med. Chem. 2017, 25, 5820–5837. [Google Scholar] [CrossRef]
  79. Canal, C.E.; Morgan, D.; Felsing, D.; Kondabolu, K.; Rowland, N.E.; Robertson, K.L.; Sakhuja, R.; Booth, R.G. A novel aminotetralin-type serotonin (5-HT) 2C receptor-specific agonist and 5-HT2A competitive antagonist/5-HT2B inverse agonist with preclinical efficacy for psychoses. J. Pharmacol. Exp. Ther. 2014, 349, 310–318. [Google Scholar] [CrossRef]
  80. Sinh, V.; Ootsuka, Y. Blockade of 5-HT2A receptors inhibits emotional hyperthermia in mice. J. Physiol. Sci. 2019, 69, 1097–1102. [Google Scholar] [CrossRef]
  81. Aly, S.A.; Hossain, M.; Bhuiyan, M.A.; Nakamura, T.; Nagatomo, T. Assessment of binding affinity to 5-hydroxytryptamine 2A (5-HT2A) receptor and inverse agonist activity of naftidrofuryl: Comparison with those of sarpogrelate. J. Pharmacol. Sci. 2009, 110, 445–450. [Google Scholar] [CrossRef]
  82. Schotte, A.; Janssen, P.F.; Gommeren, W.; Luyten, W.H.; Van Gompel, P.; Lesage, A.S.; De Loore, K.; Leysen, J.E. Risperidone compared with new and reference antipsychotic drugs: In vitro and in vivo receptor binding. Psychopharmacology 1996, 124, 57–73. [Google Scholar] [CrossRef]
  83. Meltzer, H.Y.; Massey, B.W. The role of serotonin receptors in the action of atypical antipsychotic drugs. Curr. Opin. Pharmacol. 2011, 11, 59–67. [Google Scholar] [CrossRef]
  84. Kennett, G.A.; Pittaway, K.; Blackburn, T.P. Evidence that 5-HT2c receptor antagonists are anxiolytic in the rat Geller-Seifter model of anxiety. Psychopharmacology 1994, 114, 90–96. [Google Scholar] [CrossRef]
  85. Martin, J.R.; Bos, M.; Jenck, F.; Moreau, J.; Mutel, V.; Sleight, A.J.; Wichmann, J.; Andrews, J.S.; Berendsen, H.H.; Broekkamp, C.L.; et al. 5-HT2C receptor agonists: Pharmacological characteristics and therapeutic potential. J. Pharmacol. Exp. Ther. 1998, 286, 913–924. [Google Scholar]
  86. Smith, B.M.; Smith, J.M.; Tsai, J.H.; Schultz, J.A.; Gilson, C.A.; Estrada, S.A.; Chen, R.R.; Park, D.M.; Prieto, E.B.; Gallardo, C.S.; et al. Discovery and structure-activity relationship of (1R)-8-chloro-2,3,4,5-tetrahydro-1-methyl-1H-3-benzazepine (Lorcaserin), a selective serotonin 5-HT2C receptor agonist for the treatment of obesity. J. Med. Chem. 2008, 51, 305–313. [Google Scholar] [CrossRef]
  87. do Carmo Silva, R.X.; do Nascimento, B.G.; Gomes, G.C.V.; da Silva, N.A.H.; Pinheiro, J.S.; da Silva Chaves, S.N.; Pimentel, A.F.N.; Costa, B.P.D.; Herculano, A.M.; Lima-Maximino, M.; et al. 5-HT2C agonists and antagonists block different components of behavioral responses to potential, distal, and proximal threat in zebrafish. Pharmacol. Biochem. Behav. 2021, 210, 173276. [Google Scholar] [CrossRef] [PubMed]
  88. Demireva, E.Y.; Suri, D.; Morelli, E.; Mahadevia, D.; Chuhma, N.; Teixeira, C.M.; Ziolkowski, A.; Hersh, M.; Fifer, J.; Bagchi, S.; et al. 5-HT2C receptor blockade reverses SSRI-associated basal ganglia dysfunction and potentiates therapeutic efficacy. Mol. Psychiatry 2020, 25, 3304–3321. [Google Scholar] [CrossRef]
  89. Millan, M.J. Serotonin 5-HT2C receptors as a target for the treatment of depressive and anxious states: Focus on novel therapeutic strategies. Therapie 2005, 60, 441–460. [Google Scholar] [CrossRef]
  90. Cervo, L.; Samanin, R. 5-HT1A receptor full and partial agonists and 5-HT2C (but not 5-HT3) receptor antagonists increase rates of punished responding in rats. Pharmacol. Biochem. Behav. 1995, 52, 671–676. [Google Scholar] [CrossRef]
  91. Williams, B.P.; Milligan, C.J.; Street, M.; Hornby, F.M.; Deuchars, J.; Buckley, N.J. Transcription of the M1 muscarinic receptor gene in neurons and neuronal progenitors of the embryonic rat forebrain. J. Neurochem. 2004, 88, 70–77. [Google Scholar] [CrossRef] [PubMed]
  92. Benninghoff, J.; Rauh, W.; Brantl, V.; Schloesser, R.J.; Moessner, R.; Moller, H.J.; Rujescu, D. Cholinergic impact on neuroplasticity drives muscarinic M1 receptor mediated differentiation into neurons. World J. Biol. Psychiatry 2013, 14, 241–246. [Google Scholar] [CrossRef] [PubMed]
  93. Fisher, A.; Heldman, E.; Gurwitz, D.; Haring, R.; Barak, D.; Meshulam, H.; Marciano, D.; Brandeis, R.; Pittel, Z.; Segal, M.; et al. Selective signaling via unique M1 muscarinic agonists. Ann. N. Y. Acad. Sci. 1993, 695, 300–303. [Google Scholar] [CrossRef] [PubMed]
  94. Bymaster, F.P.; Wong, D.T.; Mitch, C.H.; Ward, J.S.; Calligaro, D.O.; Schoepp, D.D.; Shannon, H.E.; Sheardown, M.J.; Olesen, P.H.; Suzdak, P.D.; et al. Neurochemical effects of the M1 muscarinic agonist xanomeline (LY246708/NNC11-0232). J. Pharmacol. Exp. Ther. 1994, 269, 282–289. [Google Scholar] [PubMed]
  95. Harries, M.H.; Samson, N.A.; Cilia, J.; Hunter, A.J. The profile of sabcomeline (SB-202026), a functionally selective M1 receptor partial agonist, in the marmoset. Br. J. Pharmacol. 1998, 124, 409–415. [Google Scholar] [CrossRef] [PubMed]
  96. Jacobson, M.A.; Kreatsoulas, C.; Pascarella, D.M.; O’Brien, J.A.; Sur, C. The M1 muscarinic receptor allosteric agonists AC-42 and 1-[1’-(2-methylbenzyl)-1,4’-bipiperidin-4-yl]-1,3-dihydro-2H-benzimidazol-2-one bind to a unique site distinct from the acetylcholine orthosteric site. Mol. Pharmacol. 2010, 78, 648–657. [Google Scholar] [CrossRef]
  97. Vanderheyden, P.; Gies, J.P.; Ebinger, G.; De Keyser, J.; Landry, Y.; Vauquelin, G. Human M1-, M2- and M3-muscarinic cholinergic receptors: Binding characteristics of agonists and antagonists. J. Neurol. Sci. 1990, 97, 67–80. [Google Scholar] [CrossRef] [PubMed]
  98. Langmead, C.J.; Austin, N.E.; Branch, C.L.; Brown, J.T.; Buchanan, K.A.; Davies, C.H.; Forbes, I.T.; Fry, V.A.; Hagan, J.J.; Herdon, H.J.; et al. Characterization of a CNS penetrant, selective M1 muscarinic receptor agonist, 77-LH-28-1. Br. J. Pharmacol. 2008, 154, 1104–1115. [Google Scholar] [CrossRef]
  99. Del Tacca, M.; Danesi, R.; Blandizzi, C.; Bernardini, M.C. A selective antimuscarinic agent: Pirenzepine. Review of its pharmacologic and clinical properties. Minerva Dietol. Gastroenterol. 1989, 35, 175–189. [Google Scholar] [PubMed]
  100. Eltze, M. Assessment of selectivity of muscarinic antagonists for M1 and M2 receptors in rabbit isolated vas deferens. Pharmacology 1988, 37 (Suppl. 1), 40–47. [Google Scholar] [CrossRef]
  101. Salin-Pascual, R.J.; Granados-Fuentes, D.; Galicia-Polo, L.; Nieves, E.; Gillin, J.C. Development of tolerance after repeated administration of a selective muscarinic M1 antagonist biperiden in healthy human volunteers. Biol. Psychiatry 1993, 33, 188–193. [Google Scholar] [CrossRef]
  102. Palma, A.; Chara, J.C.; Montilla, A.; Otxoa-de-Amezaga, A.; Ruiz-Jaen, F.; Planas, A.M.; Matute, C.; Perez-Samartin, A.; Domercq, M. Clemastine Induces an Impairment in Developmental Myelination. Front. Cell Dev. Biol. 2022, 10, 841548. [Google Scholar] [CrossRef]
  103. Wu, G.Y.; Han, X.H.; Zhuang, Q.X.; Zhang, J.; Yung, W.H.; Chan, Y.S.; Zhu, J.N.; Wang, J.J. Excitatory effect of histamine on rat spinal motoneurons by activation of both H1 and H2 receptors in vitro. J. Neurosci. Res. 2012, 90, 132–142. [Google Scholar] [CrossRef]
  104. Lacour, M.; Sterkers, O. Histamine and betahistine in the treatment of vertigo: Elucidation of mechanisms of action. CNS Drugs 2001, 15, 853–870. [Google Scholar] [CrossRef]
  105. Carman-Krzan, M.; Bavec, A.; Zorko, M.; Schunack, W. Molecular characterization of specific H1-receptor agonists histaprodifen and its Nα-substituted analogues on bovine aortic H1-receptors. Naunyn Schmiedebergs Arch. Pharmacol. 2003, 367, 538–546. [Google Scholar] [CrossRef]
  106. Monti, J.M.; Pellejero, T.; Jantos, H. Effects of H1- and H2-histamine receptor agonists and antagonists on sleep and wakefulness in the rat. J. Neural Transm. 1986, 66, 1–11. [Google Scholar] [CrossRef]
  107. Diaz, P.; Jones, D.G.; Kay, A.B. Histamine receptors on guinea-pig alveolar macrophages: Chemical specificity and the effects of H1- and H2-receptor agonists and antagonists. Clin. Exp. Immunol. 1979, 35, 462–469. [Google Scholar]
  108. Reinhardt, D.; Wiemann, H.M.; Schumann, H.J. Effects of the H1-antagonist promethazine and the H2-antagonist burimamide on chronotropic, inotropic and coronary vascular responses to histamine in isolated perfused guinea-pig hearts. Agents Actions 1976, 6, 683–688. [Google Scholar] [CrossRef]
  109. Moguilevsky, N.; Varsalona, F.; Noyer, M.; Gillard, M.; Guillaume, J.P.; Garcia, L.; Szpirer, C.; Szpirer, J.; Bollen, A. Stable expression of human H1-histamine-receptor cDNA in Chinese hamster ovary cells. Pharmacological characterisation of the protein, tissue distribution of messenger RNA and chromosomal localisation of the gene. Eur. J. Biochem. 1994, 224, 489–495. [Google Scholar] [CrossRef]
  110. Gillard, M.; Van Der Perren, C.; Moguilevsky, N.; Massingham, R.; Chatelain, P. Binding characteristics of cetirizine and levocetirizine to human H1 histamine receptors: Contribution of Lys191 and Thr194. Mol. Pharmacol. 2002, 61, 391–399. [Google Scholar] [CrossRef]
  111. Zhu, J.; Ma, R.; Li, G. Drug repurposing: Clemastine fumarate and neurodegeneration. Biomed. Pharmacother. 2023, 157, 113904. [Google Scholar] [CrossRef] [PubMed]
  112. Nakamura, T.; Hiraoka, K.; Harada, R.; Matsuzawa, T.; Ishikawa, Y.; Funaki, Y.; Yoshikawa, T.; Tashiro, M.; Yanai, K.; Okamura, N. Brain histamine H1 receptor occupancy after oral administration of desloratadine and loratadine. Pharmacol. Res. Perspect. 2019, 7, e00499. [Google Scholar] [CrossRef] [PubMed]
  113. Devillier, P.; Roche, N.; Faisy, C. Clinical pharmacokinetics and pharmacodynamics of desloratadine, fexofenadine and levocetirizine: A comparative review. Clin. Pharmacokinet. 2008, 47, 217–230. [Google Scholar] [CrossRef]
  114. Slater, J.W.; Zechnich, A.D.; Haxby, D.G. Second-generation antihistamines: A comparative review. Drugs 1999, 57, 31–47. [Google Scholar] [CrossRef] [PubMed]
  115. Hiramoto, T.; Ihara, Y.; Watanabe, Y. α-1 Adrenergic receptors stimulation induces the proliferation of neural progenitor cells in vitro. Neurosci. Lett. 2006, 408, 25–28. [Google Scholar] [CrossRef] [PubMed]
  116. Kim, A.K.; Souza-Formigoni, M.L. Alpha1-adrenergic drugs affect the development and expression of ethanol-induced behavioral sensitization. Behav. Brain Res. 2013, 256, 646–654. [Google Scholar] [CrossRef]
  117. Thiele, R.H.; Nemergut, E.C.; Lynch, C., 3rd. The physiologic implications of isolated alpha1 adrenergic stimulation. Anesth. Analg. 2011, 113, 284–296. [Google Scholar] [CrossRef] [PubMed]
  118. Insel, P.A. Structure and function of alpha-adrenergic receptors. Am. J. Med. 1989, 87, 12S–18S. [Google Scholar] [CrossRef] [PubMed]
  119. Bishop, M.J. Recent advances in the discovery of alpha1-adrenoceptor agonists. Current topics in medicinal chemistry. Curr. Top. Med. Chem. 2007, 7, 135–145. [Google Scholar] [CrossRef]
  120. Frishman, W.H.; Eisen, G.; Lapsker, J. Terazosin: A new long-acting α1-adrenergic antagonist for hypertension. Med. Clin. North Am. 1988, 72, 441–448. [Google Scholar] [CrossRef]
  121. Goldstein, I.I. Oral phentolamine: An alpha-1, alpha-2 adrenergic antagonist for the treatment of erectile dysfunction. Int. J. Impot. Res. 2000, 12, S75–S80. [Google Scholar] [CrossRef] [PubMed]
  122. Ito, K.; Ohtani, H.; Sawada, Y. Assessment of α1-adrenoceptor antagonists in benign prostatic hyperplasia based on the receptor occupancy theory. Br. J. Clin. Pharmacol. 2007, 63, 394–403. [Google Scholar] [CrossRef] [PubMed]
  123. Lowe, F.C. Role of the newer alpha, -adrenergic-receptor antagonists in the treatment of benign prostatic hyperplasia-related lower urinary tract symptoms. Clin. Ther. 2004, 26, 1701–1713. [Google Scholar] [CrossRef] [PubMed]
  124. Stanic, D.; Paratcha, G.; Ledda, F.; Herzog, H.; Kopin, A.S.; Hokfelt, T. Peptidergic influences on proliferation, migration, and placement of neural progenitors in the adult mouse forebrain. Proc. Natl. Acad. Sci. USA 2008, 105, 3610–3615. [Google Scholar] [CrossRef] [PubMed]
  125. Asin, K.E.; Bednarz, L.; Nikkel, A.L.; Gore, P.A., Jr.; Nadzan, A.M. A-71623, a selective CCK-A receptor agonist, suppresses food intake in the mouse, dog, and monkey. Pharmacol. Biochem. Behav. 1992, 42, 699–704. [Google Scholar] [CrossRef] [PubMed]
  126. Gouldson, P.; Legoux, P.; Carillon, C.; Delpech, B.; Le Fur, G.; Ferrara, P.; Shire, D. The agonist SR 146131 and the antagonist SR 27897 occupy different sites on the human CCK1 receptor. Eur. J. Pharmacol. 2000, 400, 185–194, Erratum in Eur. J. Pharmacol. 2000, 407, 333. [Google Scholar] [CrossRef] [PubMed]
  127. Szewczyk, J.R.; Laudeman, C. CCK1R agonists: A promising target for the pharmacological treatment of obesity. Curr. Top. Med. Chem. 2003, 3, 837–854. [Google Scholar] [CrossRef] [PubMed]
  128. Wang, H.H.; Portincasa, P.; Wang, D.Q. The cholecystokinin-1 receptor antagonist devazepide increases cholesterol cholelithogenesis in mice. Eur. J. Clin. Investig. 2016, 46, 158–169. [Google Scholar] [CrossRef]
  129. Makovec, F.; Bani, M.; Cereda, R.; Chiste, R.; Pacini, M.A.; Revel, L.; Rovati, L.A.; Rovati, L.C.; Setnikar, I. Pharmacological properties of lorglumide as a member of a new class of cholecystokinin antagonists. Arzneimittelforschung 1987, 37, 1265–1268. [Google Scholar]
  130. Bandyopadhyay, S.; Tfelt-Hansen, J.; Chattopadhyay, N. Diverse roles of extracellular calcium-sensing receptor in the central nervous system. J. Neurosci. Res. 2010, 88, 2073–2082. [Google Scholar] [CrossRef]
  131. Gregory, K.J.; Kufareva, I.; Keller, A.N.; Khajehali, E.; Mun, H.C.; Goolam, M.A.; Mason, R.S.; Capuano, B.; Conigrave, A.D.; Christopoulos, A.; et al. Dual Action Calcium-Sensing Receptor Modulator Unmasks Novel Mode-Switching Mechanism. ACS Pharmacol. Transl. Sci. 2018, 1, 96–109. [Google Scholar] [CrossRef]
  132. McLarnon, S.J.; Riccardi, D. Physiological and pharmacological agonists of the extracellular Ca2+-sensing receptor. Eur. J. Pharmacol. 2002, 447, 271–278. [Google Scholar] [CrossRef] [PubMed]
  133. Martin, K.J.; Bell, G.; Pickthorn, K.; Huang, S.; Vick, A.; Hodsman, P.; Peacock, M. Velcalcetide (AMG 416), a novel peptide agonist of the calcium-sensing receptor, reduces serum parathyroid hormone and FGF23 levels in healthy male subjects. Nephrol. Dial. Transplant. 2014, 29, 385–392. [Google Scholar] [CrossRef] [PubMed]
  134. Centeno, P.P.; Herberger, A.; Mun, H.C.; Tu, C.; Nemeth, E.F.; Chang, W.; Conigrave, A.D.; Ward, D.T. Phosphate acts directly on the calcium-sensing receptor to stimulate parathyroid hormone secretion. Nat. Commun. 2019, 10, 4693. [Google Scholar] [CrossRef] [PubMed]
  135. Filopanti, M.; Corbetta, S.; Barbieri, A.M.; Spada, A. Pharmacology of the calcium sensing receptor. Clin. Cases Miner. Bone Metab. 2013, 10, 162–165. [Google Scholar] [PubMed]
  136. Shcherbakova, I.; Huang, G.; Geoffroy, O.J.; Nair, S.K.; Swierczek, K.; Balandrin, M.F.; Fox, J.; Heaton, W.L.; Conklin, R.L. Design, new synthesis, and calcilytic activity of substituted 3H-pyrimidin-4-ones. Bioorg. Med. Chem. Lett. 2005, 15, 2537–2540. [Google Scholar] [CrossRef] [PubMed]
  137. Yang, W.; Wang, Y.; Roberge, J.Y.; Ma, Z.; Liu, Y.; Michael Lawrence, R.; Rotella, D.P.; Seethala, R.; Feyen, J.H.; Dickson, J.K., Jr. Discovery and structure-activity relationships of 2-benzylpyrrolidine-substituted aryloxypropanols as calcium-sensing receptor antagonists. Bioorg. Med. Chem. Lett. 2005, 15, 1225–1228. [Google Scholar] [CrossRef] [PubMed]
  138. Kessler, A.; Faure, H.; Roussanne, M.C.; Ferry, S.; Ruat, M.; Dauban, P.; Dodd, R.H. N1-Arylsulfonyl-N2-(1-(1-naphthyl)ethyl)-1,2-diaminocyclohexanes: A new class of calcilytic agents acting at the calcium-sensing receptor. Chembiochem 2004, 5, 1131–1136. [Google Scholar] [CrossRef]
  139. Gavai, A.V.; Vaz, R.J.; Mikkilineni, A.B.; Roberge, J.Y.; Liu, Y.; Lawrence, R.M.; Corte, J.R.; Yang, W.; Bednarz, M.; Dickson, J.K., Jr.; et al. Discovery of novel 1-arylmethyl pyrrolidin-2-yl ethanol amines as calcium-sensing receptor antagonists. Bioorg. Med. Chem. Lett. 2005, 15, 5478–5482. [Google Scholar] [CrossRef]
  140. Montell, C. The TRP superfamily of cation channels. Sci. STKE 2005, 2005, re3. [Google Scholar] [CrossRef]
  141. Gaudet, R. Divide and conquer: High resolution structural information on TRP channel fragments. J. Gen. Physiol. 2009, 133, 231–237. [Google Scholar] [CrossRef]
  142. Zhang, M.; Ma, Y.; Ye, X.; Zhang, N.; Pan, L.; Wang, B. TRP (transient receptor potential) ion channel family: Structures, biological functions and therapeutic interventions for diseases. Signal Transduct. Target. Ther. 2023, 8, 261. [Google Scholar] [CrossRef] [PubMed]
  143. Kalia, J.; Swartz, K.J. Exploring structure-function relationships between TRP and Kv channels. Sci. Rep. 2013, 3, 1523. [Google Scholar] [CrossRef] [PubMed]
  144. Minke, B.; Wu, C.; Pak, W.L. Induction of photoreceptor voltage noise in the dark in Drosophila mutant. Nature 1975, 258, 84–87. [Google Scholar] [CrossRef] [PubMed]
  145. Cosens, D.J.; Manning, A. Abnormal electroretinogram from a Drosophila mutant. Nature 1969, 224, 285–287. [Google Scholar] [CrossRef]
  146. Montell, C.; Birnbaumer, L.; Flockerzi, V.; Bindels, R.J.; Bruford, E.A.; Caterina, M.J.; Clapham, D.E.; Harteneck, C.; Heller, S.; Julius, D.; et al. A unified nomenclature for the superfamily of TRP cation channels. Mol. Cell 2002, 9, 229–231. [Google Scholar] [CrossRef] [PubMed]
  147. Clapham, D.E.; Montell, C.; Schultz, G.; Julius, D. International Union of Pharmacology. XLIII. Compendium of voltage-gated ion channels: Transient receptor potential channels. Pharmacol. Rev. 2003, 55, 591–596. [Google Scholar] [CrossRef] [PubMed]
  148. Cheng, W.; Sun, C.; Zheng, J. Heteromerization of TRP channel subunits: Extending functional diversity. Protein Cell 2010, 1, 802–810. [Google Scholar] [CrossRef] [PubMed]
  149. Vannier, B.; Peyton, M.; Boulay, G.; Brown, D.; Qin, N.; Jiang, M.; Zhu, X.; Birnbaumer, L. Mouse trp2, the homologue of the human trpc2 pseudogene, encodes mTrp2, a store depletion-activated capacitative Ca2+ entry channel. Proc. Natl. Acad. Sci. USA 1999, 96, 2060–2064. [Google Scholar] [CrossRef]
  150. Zeng, C.; Tian, F.; Xiao, B. TRPC Channels: Prominent Candidates of Underlying Mechanism in Neuropsychiatric Diseases. Mol. Neurobiol. 2016, 53, 631–647. [Google Scholar] [CrossRef]
  151. Hao, B.; Lu, Y.; Wang, Q.; Guo, W.; Cheung, K.H.; Yue, J. Role of STIM1 in survival and neural differentiation of mouse embryonic stem cells independent of Orai1-mediated Ca2+ entry. Stem Cell Res. 2014, 12, 452–466. [Google Scholar] [CrossRef] [PubMed]
  152. Hao, B.; Webb, S.E.; Miller, A.L.; Yue, J. The role of Ca2+ signaling on the self-renewal and neural differentiation of embryonic stem cells (ESCs). Cell Calcium 2016, 59, 67–74. [Google Scholar] [CrossRef] [PubMed]
  153. Hao, H.B.; Webb, S.E.; Yue, J.; Moreau, M.; Leclerc, C.; Miller, A.L. TRPC3 is required for the survival, pluripotency and neural differentiation of mouse embryonic stem cells (mESCs). Sci. China Life Sci. 2018, 61, 253–265. [Google Scholar] [CrossRef] [PubMed]
  154. Shin, H.Y.; Hong, Y.H.; Jang, S.S.; Chae, H.G.; Paek, S.L.; Moon, H.E.; Kim, D.G.; Kim, J.; Paek, S.H.; Kim, S.J. A role of canonical transient receptor potential 5 channel in neuronal differentiation from A2B5 neural progenitor cells. PLoS ONE 2010, 5, e10359. [Google Scholar] [CrossRef] [PubMed]
  155. Tai, Y.; Feng, S.; Du, W.; Wang, Y. Functional roles of TRPC channels in the developing brain. Pflugers Arch. 2009, 458, 283–289. [Google Scholar] [CrossRef] [PubMed]
  156. Tai, Y.; Jia, Y. TRPC Channels and Neuron Development, Plasticity, and Activities. Adv. Exp. Med. Biol. 2017, 976, 95–110. [Google Scholar] [CrossRef] [PubMed]
  157. Berg, A.P.; Sen, N.; Bayliss, D.A. TrpC3/C7 and Slo2.1 are molecular targets for metabotropic glutamate receptor signaling in rat striatal cholinergic interneurons. J. Neurosci. 2007, 27, 8845–8856. [Google Scholar] [CrossRef] [PubMed]
  158. Nakata, H.; Nakamura, S. Brain-derived neurotrophic factor regulates AMPA receptor trafficking to post-synaptic densities via IP3R and TRPC calcium signaling. FEBS Lett. 2007, 581, 2047–2054. [Google Scholar] [CrossRef]
  159. Lemonnier, L.; Trebak, M.; Putney, J.W., Jr. Complex regulation of the TRPC3, 6 and 7 channel subfamily by diacylglycerol and phosphatidylinositol-4,5-bisphosphate. Cell Calcium 2008, 43, 506–514. [Google Scholar] [CrossRef]
  160. Trebak, M.; Lemonnier, L.; DeHaven, W.I.; Wedel, B.J.; Bird, G.S.; Putney, J.W., Jr. Complex functions of phosphatidylinositol 4,5-bisphosphate in regulation of TRPC5 cation channels. Pflugers Arch. 2009, 457, 757–769. [Google Scholar] [CrossRef]
  161. Tang, J.; Lin, Y.; Zhang, Z.; Tikunova, S.; Birnbaumer, L.; Zhu, M.X. Identification of common binding sites for calmodulin and inositol 1,4,5-trisphosphate receptors on the carboxyl termini of trp channels. J. Biol. Chem. 2001, 276, 21303–21310. [Google Scholar] [CrossRef] [PubMed]
  162. Zhang, Z.; Tang, J.; Tikunova, S.; Johnson, J.D.; Chen, Z.; Qin, N.; Dietrich, A.; Stefani, E.; Birnbaumer, L.; Zhu, M.X. Activation of Trp3 by inositol 1,4,5-trisphosphate receptors through displacement of inhibitory calmodulin from a common binding domain. Proc. Natl. Acad. Sci. USA 2001, 98, 3168–3173. [Google Scholar] [CrossRef] [PubMed]
  163. Numaga, T.; Wakamori, M.; Mori, Y. Trpc7. In Transient Receptor Potential (TRP) Chanels; Handbook of Expermental Pharmacology Book Series; Springer: Berlin/Heidelberg, Germany, 2007; Volume 179, pp. 143–151. [Google Scholar] [CrossRef]
  164. Strubing, C.; Krapivinsky, G.; Krapivinsky, L.; Clapham, D.E. Formation of novel TRPC channels by complex subunit interactions in embryonic brain. J. Biol. Chem. 2003, 278, 39014–39019. [Google Scholar] [CrossRef] [PubMed]
  165. Golovina, V.A.; Platoshyn, O.; Bailey, C.L.; Wang, J.; Limsuwan, A.; Sweeney, M.; Rubin, L.J.; Yuan, J.X. Upregulated TRP and enhanced capacitative Ca2+ entry in human pulmonary artery myocytes during proliferation. Am. J. Physiol. Heart Circ. Physiol. 2001, 280, H746–H755. [Google Scholar] [CrossRef] [PubMed]
  166. Antoniotti, S.; Lovisolo, D.; Fiorio Pla, A.; Munaron, L. Expression and functional role of bTRPC1 channels in native endothelial cells. FEBS Lett. 2002, 510, 189–195. [Google Scholar] [CrossRef] [PubMed]
  167. Vaccarino, F.M.; Schwartz, M.L.; Raballo, R.; Nilsen, J.; Rhee, J.; Zhou, M.; Doetschman, T.; Coffin, J.D.; Wyland, J.J.; Hung, Y.T. Changes in cerebral cortex size are governed by fibroblast growth factor during embryogenesis. Nat. Neurosci. 1999, 2, 848. [Google Scholar] [CrossRef] [PubMed]
  168. Vaccarino, F.M.; Schwartz, M.L.; Raballo, R.; Rhee, J.; Lyn-Cook, R. Fibroblast growth factor signaling regulates growth and morphogenesis at multiple steps during brain development. Curr. Top. Dev. Biol. 1999, 46, 179–200. [Google Scholar] [CrossRef] [PubMed]
  169. Raballo, R.; Rhee, J.; Lyn-Cook, R.; Leckman, J.F.; Schwartz, M.L.; Vaccarino, F.M. Basic fibroblast growth factor (Fgf2) is necessary for cell proliferation and neurogenesis in the developing cerebral cortex. J. Neurosci. 2000, 20, 5012–5023. [Google Scholar] [CrossRef]
  170. Fiorio Pla, A.; Maric, D.; Brazer, S.C.; Giacobini, P.; Liu, X.; Chang, Y.H.; Ambudkar, I.S.; Barker, J.L. Canonical transient receptor potential 1 plays a role in basic fibroblast growth factor (bFGF)/FGF receptor-1-induced Ca2+ entry and embryonic rat neural stem cell proliferation. J. Neurosci. 2005, 25, 2687–2701. [Google Scholar] [CrossRef]
  171. Li, M.; Chen, C.; Zhou, Z.; Xu, S.; Yu, Z. A TRPC1-mediated increase in store-operated Ca2+ entry is required for the proliferation of adult hippocampal neural progenitor cells. Cell Calcium 2012, 51, 486–496. [Google Scholar] [CrossRef]
  172. She, Y.J.; Pan, J.; Peng, L.M.; Ma, L.; Guo, X.; Lei, D.X.; Wang, H.Z. Ketamine modulates neural stem cell differentiation by regulating TRPC3 expression through the GSK3β/β-catenin pathway. Neurotoxicology 2023, 94, 1–10. [Google Scholar] [CrossRef] [PubMed]
  173. Wu, X.; Zagranichnaya, T.K.; Gurda, G.T.; Eves, E.M.; Villereal, M.L. A TRPC1/TRPC3-mediated increase in store-operated calcium entry is required for differentiation of H19-7 hippocampal neuronal cells. J. Biol. Chem. 2004, 279, 43392–43402. [Google Scholar] [CrossRef] [PubMed]
  174. Seebohm, G.; Schreiber, J.A. Beyond Hot and Spicy: TRPV Channels and their Pharmacological Modulation. Cell Physiol. Biochem. 2021, 55, 108–130. [Google Scholar] [CrossRef] [PubMed]
  175. Clapham, D.E. TRP channels as cellular sensors. Nature 2003, 426, 517–524. [Google Scholar] [CrossRef] [PubMed]
  176. Garcia-Sanz, N.; Valente, P.; Gomis, A.; Fernandez-Carvajal, A.; Fernandez-Ballester, G.; Viana, F.; Belmonte, C.; Ferrer-Montiel, A. A role of the transient receptor potential domain of vanilloid receptor I in channel gating. J. Neurosci. 2007, 27, 11641–11650. [Google Scholar] [CrossRef] [PubMed]
  177. Lishko, P.V.; Procko, E.; Jin, X.; Phelps, C.B.; Gaudet, R. The ankyrin repeats of TRPV1 bind multiple ligands and modulate channel sensitivity. Neuron 2007, 54, 905–918. [Google Scholar] [CrossRef] [PubMed]
  178. Sasase, T.; Fatchiyah, F.; Ohta, T. Transient receptor potential vanilloid (TRPV) channels: Basal properties and physiological potential. Gen. Physiol. Biophys. 2022, 41, 165–190. [Google Scholar] [CrossRef] [PubMed]
  179. Caterina, M.J.; Rosen, T.A.; Tominaga, M.; Brake, A.J.; Julius, D. A capsaicin-receptor homologue with a high threshold for noxious heat. Nature 1999, 398, 436–441. [Google Scholar] [CrossRef]
  180. Kanzaki, M.; Zhang, Y.Q.; Mashima, H.; Li, L.; Shibata, H.; Kojima, I. Translocation of a calcium-permeable cation channel induced by insulin-like growth factor-I. Nat. Cell Biol. 1999, 1, 165–170. [Google Scholar] [CrossRef]
  181. Penna, A.; Juvin, V.; Chemin, J.; Compan, V.; Monet, M.; Rassendren, F.A. PI3-kinase promotes TRPV2 activity independently of channel translocation to the plasma membrane. Cell Calcium 2006, 39, 495–507. [Google Scholar] [CrossRef]
  182. Morgan, P.J.; Hubner, R.; Rolfs, A.; Frech, M.J. Spontaneous calcium transients in human neural progenitor cells mediated by transient receptor potential channels. Stem Cells Dev. 2013, 22, 2477–2486. [Google Scholar] [CrossRef] [PubMed]
  183. Santoni, G.; Maggi, F.; Morelli, M.B.; Santoni, M.; Marinelli, O. Transient Receptor Potential Cation Channels in Cancer Therapy. Med. Sci. 2019, 7, 108. [Google Scholar] [CrossRef] [PubMed]
  184. Ohashi, K.; Deyashiki, A.; Miyake, T.; Nagayasu, K.; Shibasaki, K.; Shirakawa, H.; Kaneko, S. TRPV4 is functionally expressed in oligodendrocyte precursor cells and increases their proliferation. Pflugers Arch. 2018, 470, 705–716. [Google Scholar] [CrossRef] [PubMed]
  185. Torres, V.E.; Harris, P.C. Polycystic kidney disease: Genes, proteins, animal models, disease mechanisms and therapeutic opportunities. J. Intern. Med. 2007, 261, 17–31. [Google Scholar] [CrossRef]
  186. Giamarchi, A.; Padilla, F.; Coste, B.; Raoux, M.; Crest, M.; Honore, E.; Delmas, P. The versatile nature of the calcium-permeable cation channel TRPP2. EMBO Rep. 2006, 7, 787–793. [Google Scholar] [CrossRef] [PubMed]
  187. Franco, S.J.; Muller, U. Shaping our minds: Stem and progenitor cell diversity in the mammalian neocortex. Neuron 2013, 77, 19–34. [Google Scholar] [CrossRef]
  188. Li, Y.; Wright, J.M.; Qian, F.; Germino, G.G.; Guggino, W.B. Polycystin 2 interacts with type I inositol 1,4,5-trisphosphate receptor to modulate intracellular Ca2+ signaling. J. Biol. Chem. 2005, 280, 41298–41306. [Google Scholar] [CrossRef] [PubMed]
  189. Li, Y.; Santoso, N.G.; Yu, S.; Woodward, O.M.; Qian, F.; Guggino, W.B. Polycystin-1 interacts with inositol 1,4,5-trisphosphate receptor to modulate intracellular Ca2+ signaling with implications for polycystic kidney disease. J. Biol. Chem. 2009, 284, 36431–36441. [Google Scholar] [CrossRef]
  190. Tian, P.F.; Sun, M.M.; Hu, X.Y.; Du, J.; He, W. TRPP2 ion channels: The roles in various subcellular locations. Biochimie 2022, 201, 116–127. [Google Scholar] [CrossRef]
  191. Winokurow, N.; Schumacher, S. A role for polycystin-1 and polycystin-2 in neural progenitor cell differentiation. Cell Mol. Life Sci. 2019, 76, 2851–2869. [Google Scholar] [CrossRef]
  192. Gaiano, N.; Nye, J.S.; Fishell, G. Radial glial identity is promoted by Notch1 signaling in the murine forebrain. Neuron 2000, 26, 395–404. [Google Scholar] [CrossRef] [PubMed]
  193. Hong, S.; Song, M.R. Signal transducer and activator of transcription-3 maintains the stemness of radial glia at mid-neurogenesis. J. Neurosci. 2015, 35, 1011–1023. [Google Scholar] [CrossRef] [PubMed]
  194. Trenker, R.; Jura, N. Receptor tyrosine kinase activation: From the ligand perspective. Curr. Opin. Cell Biol. 2020, 63, 174–185. [Google Scholar] [CrossRef] [PubMed]
  195. Munaron, L. Calcium signalling and control of cell proliferation by tyrosine kinase receptors (review). Int. J. Mol. Med. 2002, 10, 671–676. [Google Scholar] [CrossRef] [PubMed]
  196. Diliberto, P.A.; Hubbert, T.; Herman, B. Early PDGF-induced alterations in cytosolic free calcium are required for mitogenesis. Res. Commun. Chem. Pathol. Pharmacol. 1991, 72, 3–12. [Google Scholar] [PubMed]
  197. Moolenaar, W.H.; Aerts, R.J.; Tertoolen, L.G.; de Laat, S.W. The epidermal growth factor-induced calcium signal in A431 cells. J. Biol. Chem. 1986, 261, 279–284. [Google Scholar] [CrossRef] [PubMed]
  198. Merlet, E.; Lipskaia, L.; Marchand, A.; Hadri, L.; Mougenot, N.; Atassi, F.; Liang, L.; Hatem, S.N.; Hajjar, R.J.; Lompre, A.M. A calcium-sensitive promoter construct for gene therapy. Gene Ther. 2013, 20, 248–254. [Google Scholar] [CrossRef] [PubMed]
  199. Gomez, T.M.; Zheng, J.Q. The molecular basis for calcium-dependent axon pathfinding. Nat. Rev. Neurosci. 2006, 7, 115–125. [Google Scholar] [CrossRef] [PubMed]
  200. Wen, Z.; Zheng, J.Q. Directional guidance of nerve growth cones. Curr. Opin. Neurobiol. 2006, 16, 52–58. [Google Scholar] [CrossRef]
  201. Zheng, J.Q.; Poo, M.M. Calcium signaling in neuronal motility. Annu. Rev. Cell Dev. Biol. 2007, 23, 375–404. [Google Scholar] [CrossRef]
  202. Nelson, W.J. Remodeling epithelial cell organization: Transitions between front-rear and apical-basal polarity. Cold Spring Harb. Perspect. Biol. 2009, 1, a000513. [Google Scholar] [CrossRef] [PubMed]
  203. Kang, D.S.; Yang, Y.R.; Lee, C.; Park, B.; Park, K.I.; Seo, J.K.; Seo, Y.K.; Cho, H.; Cocco, L.; Suh, P.G. Netrin-1/DCC-mediated PLCγ1 activation is required for axon guidance and brain structure development. EMBO Rep. 2019, 20, e48117. [Google Scholar] [CrossRef] [PubMed]
  204. Chen, P.; Xie, H.; Sekar, M.C.; Gupta, K.; Wells, A. Epidermal growth factor receptor-mediated cell motility: Phospholipase C activity is required, but mitogen-activated protein kinase activity is not sufficient for induced cell movement. J. Cell Biol. 1994, 127, 847–857. [Google Scholar] [CrossRef] [PubMed]
  205. Suh, P.G.; Park, J.I.; Manzoli, L.; Cocco, L.; Peak, J.C.; Katan, M.; Fukami, K.; Kataoka, T.; Yun, S.; Ryu, S.H. Multiple roles of phosphoinositide-specific phospholipase C isozymes. BMB Rep. 2008, 41, 415–434. [Google Scholar] [CrossRef] [PubMed]
  206. Fietz, S.A.; Huttner, W.B. Cortical progenitor expansion, self-renewal and neurogenesis-a polarized perspective. Curr. Opin. Neurobiol. 2011, 21, 23–35. [Google Scholar] [CrossRef] [PubMed]
  207. Fietz, S.A.; Kelava, I.; Vogt, J.; Wilsch-Brauninger, M.; Stenzel, D.; Fish, J.L.; Corbeil, D.; Riehn, A.; Distler, W.; Nitsch, R.; et al. OSVZ progenitors of human and ferret neocortex are epithelial-like and expand by integrin signaling. Nat. Neurosci. 2010, 13, 690–699. [Google Scholar] [CrossRef]
  208. Pollard, T.D. Mechanics of cytokinesis in eukaryotes. Curr. Opin. Cell Biol. 2010, 22, 50–56. [Google Scholar] [CrossRef] [PubMed]
  209. Sun, Y.; Goderie, S.K.; Temple, S. Asymmetric distribution of EGFR receptor during mitosis generates diverse CNS progenitor cells. Neuron 2005, 45, 873–886. [Google Scholar] [CrossRef] [PubMed]
  210. Reinchisi, G.; Parada, M.; Lois, P.; Oyanadel, C.; Shaughnessy, R.; Gonzalez, A.; Palma, V. Sonic Hedgehog modulates EGFR dependent proliferation of neural stem cells during late mouse embryogenesis through EGFR transactivation. Front. Cell Neurosci. 2013, 7, 166. [Google Scholar] [CrossRef]
  211. Ayuso-Sacido, A.; Graham, C.; Greenfield, J.P.; Boockvar, J.A. The duality of epidermal growth factor receptor (EGFR) signaling and neural stem cell phenotype: Cell enhancer or cell transformer? Curr. Stem Cell Res. Ther. 2006, 1, 387–394. [Google Scholar] [CrossRef]
  212. Zhang, X.; Xiao, G.; Johnson, C.; Cai, Y.; Horowitz, Z.K.; Mennicke, C.; Coffey, R.; Haider, M.; Threadgill, D.; Eliscu, R.; et al. Bulk and mosaic deletions of Egfr reveal regionally defined gliogenesis in the developing mouse forebrain. iScience 2023, 26, 106242. [Google Scholar] [CrossRef] [PubMed]
  213. Sibilia, M.; Steinbach, J.P.; Stingl, L.; Aguzzi, A.; Wagner, E.F. A strain-independent postnatal neurodegeneration in mice lacking the EGF receptor. EMBO J. 1998, 17, 719–731. [Google Scholar] [CrossRef] [PubMed]
  214. Sabbah, D.A.; Hajjo, R.; Sweidan, K. Review on Epidermal Growth Factor Receptor (EGFR) Structure, Signaling Pathways, Interactions, and Recent Updates of EGFR Inhibitors. Curr. Top. Med. Chem. 2020, 20, 815–834. [Google Scholar] [CrossRef] [PubMed]
  215. Galvez-Contreras, A.Y.; Quinones-Hinojosa, A.; Gonzalez-Perez, O. The role of EGFR and ErbB family related proteins in the oligodendrocyte specification in germinal niches of the adult mammalian brain. Front. Cell Neurosci. 2013, 7, 258. [Google Scholar] [CrossRef] [PubMed]
  216. Talukdar, S.; Emdad, L.; Das, S.K.; Fisher, P.B. EGFR: An essential receptor tyrosine kinase-regulator of cancer stem cells. Adv. Cancer Res. 2020, 147, 161–188. [Google Scholar] [CrossRef] [PubMed]
  217. Lee, J.A.; Jang, D.J.; Kaang, B.K. Two major gate-keepers in the self-renewal of neural stem cells: Erk1/2 and PLCγ1 in FGFR signaling. Mol. Brain 2009, 2, 15. [Google Scholar] [CrossRef] [PubMed]
  218. Paek, H.; Gutin, G.; Hebert, J.M. FGF signaling is strictly required to maintain early telencephalic precursor cell survival. Development 2009, 136, 2457–2465. [Google Scholar] [CrossRef] [PubMed]
  219. Goetz, R.; Mohammadi, M. Exploring mechanisms of FGF signalling through the lens of structural biology. Nat. Rev. Mol. Cell Biol. 2013, 14, 166–180. [Google Scholar] [CrossRef]
  220. Katoh, M. FGFR inhibitors: Effects on cancer cells, tumor microenvironment and whole-body homeostasis (Review). Int. J. Mol. Med. 2016, 38, 3–15. [Google Scholar] [CrossRef]
  221. Repetto, M.; Crimini, E.; Giugliano, F.; Morganti, S.; Belli, C.; Curigliano, G. Selective FGFR/FGF pathway inhibitors: Inhibition strategies, clinical activities, resistance mutations, and future directions. Expert. Rev. Clin. Pharmacol. 2021, 14, 1233–1252. [Google Scholar] [CrossRef]
  222. Son, D.I.; Hong, S.; Shin, K.S.; Kang, S.J. PARP-1 regulates mouse embryonic neural stem cell proliferation by regulating PDGFRα expression. Biochem. Biophys. Res. Commun. 2020, 526, 986–992. [Google Scholar] [CrossRef] [PubMed]
  223. Guerit, E.; Arts, F.; Dachy, G.; Boulouadnine, B.; Demoulin, J.B. PDGF receptor mutations in human diseases. Cell Mol. Life Sci. 2021, 78, 3867–3881. [Google Scholar] [CrossRef] [PubMed]
  224. Lewandowski, S.A.; Fredriksson, L.; Lawrence, D.A.; Eriksson, U. Pharmacological targeting of the PDGF-CC signaling pathway for blood-brain barrier restoration in neurological disorders. Pharmacol. Ther. 2016, 167, 108–119. [Google Scholar] [CrossRef] [PubMed]
  225. Heldin, C.H. Targeting the PDGF signaling pathway in the treatment of non-malignant diseases. J. Neuroimmune Pharmacol. 2014, 9, 69–79. [Google Scholar] [CrossRef] [PubMed]
  226. Heldin, C.H. Targeting the PDGF signaling pathway in tumor treatment. Cell Commun. Signal. 2013, 11, 97. [Google Scholar] [CrossRef] [PubMed]
  227. Hu, Q.; Lee, S.Y.; O’Kusky, J.R.; Ye, P. Signalling through the type 1 insulin-like growth factor receptor (IGF1R) interacts with canonical Wnt signalling to promote neural proliferation in developing brain. ASN Neuro 2012, 4, 253–265. [Google Scholar] [CrossRef] [PubMed]
  228. Hakuno, F.; Takahashi, S.I. IGF1 receptor signaling pathways. J. Mol. Endocrinol. 2018, 61, T69–T86. [Google Scholar] [CrossRef] [PubMed]
  229. Mitsiades, C.S.; Mitsiades, N.S.; McMullan, C.J.; Poulaki, V.; Shringarpure, R.; Akiyama, M.; Hideshima, T.; Chauhan, D.; Joseph, M.; Libermann, T.A.; et al. Inhibition of the insulin-like growth factor receptor-1 tyrosine kinase activity as a therapeutic strategy for multiple myeloma, other hematologic malignancies, and solid tumors. Cancer Cell 2004, 5, 221–230. [Google Scholar] [CrossRef] [PubMed]
  230. Garcia-Lazaro, R.S.; Lamdan, H.; Caligiuri, L.G.; Lorenzo, N.; Berengeno, A.L.; Ortega, H.H.; Alonso, D.F.; Farina, H.G. In vitro and in vivo antitumor activity of Yerba Mate extract in colon cancer models. J. Food Sci. 2020, 85, 2186–2197. [Google Scholar] [CrossRef]
  231. Buck, E.; Gokhale, P.C.; Koujak, S.; Brown, E.; Eyzaguirre, A.; Tao, N.; Rosenfeld-Franklin, M.; Lerner, L.; Chiu, M.I.; Wild, R.; et al. Compensatory insulin receptor (IR) activation on inhibition of insulin-like growth factor-1 receptor (IGF-1R): Rationale for cotargeting IGF-1R and IR in cancer. Mol. Cancer Ther. 2010, 9, 2652–2664. [Google Scholar] [CrossRef]
  232. Islam, O.; Loo, T.X.; Heese, K. Brain-derived neurotrophic factor (BDNF) has proliferative effects on neural stem cells through the truncated TRK-B receptor, MAP kinase, AKT, and STAT-3 signaling pathways. Curr. Neurovasc. Res. 2009, 6, 42–53. [Google Scholar] [CrossRef]
  233. Bartkowska, K.; Paquin, A.; Gauthier, A.S.; Kaplan, D.R.; Miller, F.D. Trk signaling regulates neural precursor cell proliferation and differentiation during cortical development. Development 2007, 134, 4369–4380. [Google Scholar] [CrossRef]
  234. Huang, E.J.; Reichardt, L.F. Trk receptors: Roles in neuronal signal transduction. Annu. Rev. Biochem. 2003, 72, 609–642. [Google Scholar] [CrossRef]
  235. Wilkie, N.; Wingrove, P.B.; Bilsland, J.G.; Young, L.; Harper, S.J.; Hefti, F.; Ellis, S.; Pollack, S.J. The non-peptidyl fungal metabolite L-783,281 activates TRK neurotrophin receptors. J. Neurochem. 2001, 78, 1135–1145. [Google Scholar] [CrossRef]
  236. Josephy-Hernandez, S.; Jmaeff, S.; Pirvulescu, I.; Aboulkassim, T.; Saragovi, H.U. Neurotrophin receptor agonists and antagonists as therapeutic agents: An evolving paradigm. Neurobiol. Dis. 2017, 97, 139–155. [Google Scholar] [CrossRef]
  237. Cocco, E.; Scaltriti, M.; Drilon, A. NTRK fusion-positive cancers and TRK inhibitor therapy. Nat. Rev. Clin. Oncol. 2018, 15, 731–747. [Google Scholar] [CrossRef]
  238. Wang, J.M.; Zeng, Y.S.; Liu, R.Y.; Huang, W.L.; Xiong, Y.; Wang, Y.H.; Chen, S.J.; Teng, Y.D. Recombinant adenovirus vector-mediated functional expression of neurotropin-3 receptor (TrkC) in neural stem cells. Exp. Neurol. 2007, 203, 123–127. [Google Scholar] [CrossRef] [PubMed]
  239. Nielsen, M.S.; Axelsen, L.N.; Sorgen, P.L.; Verma, V.; Delmar, M.; Holstein-Rathlou, N.H. Gap junctions. Compr. Physiol. 2012, 2, 1981–2035. [Google Scholar] [CrossRef]
  240. Sohl, G.; Willecke, K. An update on connexin genes and their nomenclature in mouse and man. Cell Commun. Adhes. 2003, 10, 173–180. [Google Scholar] [CrossRef] [PubMed]
  241. Cina, C.; Bechberger, J.F.; Ozog, M.A.; Naus, C.C. Expression of connexins in embryonic mouse neocortical development. J. Comp. Neurol. 2007, 504, 298–313. [Google Scholar] [CrossRef] [PubMed]
  242. Bittman, K.; Owens, D.F.; Kriegstein, A.R.; LoTurco, J.J. Cell coupling and uncoupling in the ventricular zone of developing neocortex. J. Neurosci. 1997, 17, 7037–7044. [Google Scholar] [CrossRef]
  243. Nadarajah, B.; Jones, A.M.; Evans, W.H.; Parnavelas, J.G. Differential expression of connexins during neocortical development and neuronal circuit formation. J. Neurosci. 1997, 17, 3096–3111. [Google Scholar] [CrossRef]
  244. Oyamada, M.; Oyamada, Y.; Takamatsu, T. Regulation of connexin expression. Biochim. Biophys. Acta 2005, 1719, 6–23. [Google Scholar] [CrossRef]
  245. Liu, X.; Hashimoto-Torii, K.; Torii, M.; Ding, C.; Rakic, P. Gap junctions/hemichannels modulate interkinetic nuclear migration in the forebrain precursors. J. Neurosci. 2010, 30, 4197–4209. [Google Scholar] [CrossRef]
  246. Malmersjo, S.; Rebellato, P.; Smedler, E.; Planert, H.; Kanatani, S.; Liste, I.; Nanou, E.; Sunner, H.; Abdelhady, S.; Zhang, S.; et al. Neural progenitors organize in small-world networks to promote cell proliferation. Proc. Natl. Acad. Sci. USA 2013, 110, E1524–E1532. [Google Scholar] [CrossRef] [PubMed]
  247. Lo Turco, J.J.; Kriegstein, A.R. Clusters of coupled neuroblasts in embryonic neocortex. Science 1991, 252, 563–566. [Google Scholar] [CrossRef]
  248. Ulrich, H.; Abbracchio, M.P.; Burnstock, G. Extrinsic purinergic regulation of neural stem/progenitor cells: Implications for CNS development and repair. Stem Cell Rev. Rep. 2012, 8, 755–767. [Google Scholar] [CrossRef]
  249. Elias, L.A.; Kriegstein, A.R. Gap junctions: Multifaceted regulators of embryonic cortical development. Trends Neurosci. 2008, 31, 243–250. [Google Scholar] [CrossRef]
  250. Wiencken-Barger, A.E.; Djukic, B.; Casper, K.B.; McCarthy, K.D. A role for Connexin43 during neurodevelopment. Glia 2007, 55, 675–686. [Google Scholar] [CrossRef] [PubMed]
  251. Khodosevich, K.; Zuccotti, A.; Kreuzberg, M.M.; Le Magueresse, C.; Frank, M.; Willecke, K.; Monyer, H. Connexin45 modulates the proliferation of transit-amplifying precursor cells in the mouse subventricular zone. Proc. Natl. Acad. Sci. USA 2012, 109, 20107–20112. [Google Scholar] [CrossRef]
  252. Bittman, K.S.; LoTurco, J.J. Differential regulation of connexin 26 and 43 in murine neocortical precursors. Cereb. Cortex 1999, 9, 188–195. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Scheme of the cellular phenotype generated during development from a heterogeneous population of neural stem cells. NSCs proliferate symmetrically (Ijms 25 04073 i001) to increase their pool or asymmetrically (black) to give rise, directly or indirectly, in a time-dependent manner, to neurons (green arrows), astrocytes (blue arrows), and oligodendrocytes (red arrow). NSCs = neural stem cells; IPCs = intermediate progenitor cells. Blue doted arrow = repressed gliogenic signals; Blunt green arrow = pro-neuronal genes repressing gliogenic signals; Blunt blue arrow = gliogenic signals repressing neuronal genes. Created using BioRender.com.
Figure 1. Scheme of the cellular phenotype generated during development from a heterogeneous population of neural stem cells. NSCs proliferate symmetrically (Ijms 25 04073 i001) to increase their pool or asymmetrically (black) to give rise, directly or indirectly, in a time-dependent manner, to neurons (green arrows), astrocytes (blue arrows), and oligodendrocytes (red arrow). NSCs = neural stem cells; IPCs = intermediate progenitor cells. Blue doted arrow = repressed gliogenic signals; Blunt green arrow = pro-neuronal genes repressing gliogenic signals; Blunt blue arrow = gliogenic signals repressing neuronal genes. Created using BioRender.com.
Ijms 25 04073 g001
Figure 2. Calcium entry stimulated via GPCR activation in neural stem cells. GαqPCR is activated through ligand binding promoting PLC activation, Ca2+ is released from the endoplasmic reticulum (ER), Ca2+ depletion in the ER is sensed by the stromal interaction molecules (STIM) and its conformational change to move and contact Ca2+ channels (ORAI and TRPC1) in the cell membrane to promote Ca2+ entrance for ER refill through sarcoendoplasmic reticulum Ca2+ ATPase (SERCA) and bind to proteins with roles in cell proliferation. Created using BioRender.com.
Figure 2. Calcium entry stimulated via GPCR activation in neural stem cells. GαqPCR is activated through ligand binding promoting PLC activation, Ca2+ is released from the endoplasmic reticulum (ER), Ca2+ depletion in the ER is sensed by the stromal interaction molecules (STIM) and its conformational change to move and contact Ca2+ channels (ORAI and TRPC1) in the cell membrane to promote Ca2+ entrance for ER refill through sarcoendoplasmic reticulum Ca2+ ATPase (SERCA) and bind to proteins with roles in cell proliferation. Created using BioRender.com.
Ijms 25 04073 g002
Figure 3. TRPCs in neural stem cells. Scheme showing the two pathways for TRPCs activation: the store-operated (light green arrows) and receptor-operated activation (purple arrows) pathways. Black arrows = agonist GPCR pathways involving Gαq/11 and downstream signaling pathway for Ca2+ ER release; green arrows in the middle of TRPCs and IP3R = extracellular/intracellular Ca2+ entrance and release from de ER, respectively. And, Blunt brown arrow = negative regulation of CaM on TRPCs. Created using BioRender.com.
Figure 3. TRPCs in neural stem cells. Scheme showing the two pathways for TRPCs activation: the store-operated (light green arrows) and receptor-operated activation (purple arrows) pathways. Black arrows = agonist GPCR pathways involving Gαq/11 and downstream signaling pathway for Ca2+ ER release; green arrows in the middle of TRPCs and IP3R = extracellular/intracellular Ca2+ entrance and release from de ER, respectively. And, Blunt brown arrow = negative regulation of CaM on TRPCs. Created using BioRender.com.
Ijms 25 04073 g003
Figure 4. TRPVs and neural stem cell proliferation. The Scheme shows the TRPVs’ activation signals and related pathways for neural stem cell proliferation. Purple arrow = PI3K participation in TRPV cell membrane localization. Once in the membrane (black arrow in the left) TRPVs can be activated promoting Ca2+ entrance (green arrows). TRPV3 and 4 are negatively regulated by CaM (blunt brown arrow) and can be activated by PKC (blue arrow) through GPCR DAG and Ca2+ release (black arrows in the right). Yellow arrow = extracellular Ca2+ influence TRPV4 activity. Created using BioRender.com.
Figure 4. TRPVs and neural stem cell proliferation. The Scheme shows the TRPVs’ activation signals and related pathways for neural stem cell proliferation. Purple arrow = PI3K participation in TRPV cell membrane localization. Once in the membrane (black arrow in the left) TRPVs can be activated promoting Ca2+ entrance (green arrows). TRPV3 and 4 are negatively regulated by CaM (blunt brown arrow) and can be activated by PKC (blue arrow) through GPCR DAG and Ca2+ release (black arrows in the right). Yellow arrow = extracellular Ca2+ influence TRPV4 activity. Created using BioRender.com.
Ijms 25 04073 g004
Figure 5. TRPPs. Scheme showing the interaction of PC2 with other channels and receptors in the plasma membrane and the endoplasmic reticulum membrane (ER). Created using BioRender.com.
Figure 5. TRPPs. Scheme showing the interaction of PC2 with other channels and receptors in the plasma membrane and the endoplasmic reticulum membrane (ER). Created using BioRender.com.
Ijms 25 04073 g005
Figure 6. RTK pathways. Scheme showing the pathways activated by RTKs. The pathways in the black arrows are the PLCγs and the cross-taking with GPCR receptors, both promoting an increase in [Ca2+]i. Created using BioRender.com.
Figure 6. RTK pathways. Scheme showing the pathways activated by RTKs. The pathways in the black arrows are the PLCγs and the cross-taking with GPCR receptors, both promoting an increase in [Ca2+]i. Created using BioRender.com.
Ijms 25 04073 g006
Table 1. Gαq-coupled receptors that affect cell proliferation.
Table 1. Gαq-coupled receptors that affect cell proliferation.
Receptor Function Ligand Agonists Antagonist
5-HT2AR
(serotonin 2A receptor)
Promotes NSC proliferation ex vivo, in vitro, and in vivo [69]
Its antagonism increases NSC differentiation in vitro [70]
Serotonin LSD [71], DOI [72], 25CN-NBOH [73], Mescaline [74], Pimavanserin [75], S 16924 [76]Spiperone [71,77], compound 3b [PMID: 28943244] [78], M100,907 (volinanserin), Pirenperone [71], (−)-MBP (meta-bromo-phenylisopropylamine) [79],
Eplivanserin hemi fumarate (SR-46349B) [80], Sarpogrelate, Naftidrofuryl [81], Risperidone, Pipamperone [82], Olanzapine, Ketanserin, Clozapine, Zotepine, Ziprasidone, ORG5222, Tiaspirone, Ocaperidone [83], Ketanserin, Altanserin [84]
5-HT2CR
(serotonin 2C receptor)
Its antagonism increases NSC differentiation in vitro [70]. Serotonin LSD [71], DOI [72], Mescaline [74], (−)-MBP (meta-bromo-phenylisopropylamine) [79], RO 60-0175 [85], Lorcaserin [86], MK-212, WAY-161503 [87]Spiperone [71], RS-102221[87], Mianserin, 1-NP, ICI 169,369, LY 53857 [84], SB206553, SB242084 [88], Nefazodone, Mirtazapine [89], Ritanserin, Mesulergine [90]
M1R
(muscarinic type 1 receptor)
Promotes NSC differentiation in vitro [91,92]. Acetylcholine AF102B, AF150 (S), AF267B [93], Xanomeline [94], Sabcomeline [95], AC-42, TBPB, N-desmethylclozapine [96], Pirenzepine, Carbachol [97], 77-LH-28-1 [98]Pirenzepin [99], Telenzepine [100], Biperiden [101], Clemastine [102]
H1R
(Histamine type 1 receptor)
Increases NSC neuron differentiation in vitro. Its antagonism reduces neurogenesis in vivo [32]. Histamine PEA [103], Beta-histine [104], Histaprodifen [105], 2-TEA [106]Diphenhydramine, Pyrilamine [106], Chlorpheniramine, Mepyramine [107], Promethazine [108], Cetirizine [109], Hydroxyzine [110], Clemastine [111], Loratadine, Desloratadine [112], Fexofenadine, Levocetirizine [113], Azelastine, Acrivastine, Astemizole, Ebastine, Fexodenadine, Ketotifen, Mizolastine, Terfenadine [114]
ADRα-1AR
(adrenergic α1-receptor)
Increases NSC proliferation in vitro [115].Norepinephrine (noradrenaline)
Epinephrine (adrenaline)
Phenylephrine [116], Methoxamine [117], Metaraminol, Midodrine, Xylometazoline, Oxymetazoline, Naphazoline, Tetrahydrozoline [118], Clonidine, Cirazoline, Sgd 101/75, St 587, Amidephrine, SKF89748, SDZ NVI 085, SK&F 102652, ST-1059, A-61603, A-204176, NS-49, ABT-866, BMY 7378 [119]Prazosin [116], Terazosin [117], Tamsulosin [120], Phentolamine [121], Doxazosin [122], Alfuzosin [123]
CCK1R
(Cholecystokinin type 1 receptor)
Increases NSC proliferation and differentiation into neurons in vitro [124]. CCK-8 A-71623 [125], SR-146131 [126], FPL 14294, AR-R 15849, A-71623, PD149164, PD170292, PD151932, GI 18177 [127]SR 27897 [126], L-364,718, Devazepide [128], Dexloxiglumide, Lorglumide, Proglumide [129], and MK 329 (devazepide) [127]
CaSR
(Ca2+-sensing receptor)
NSCs differentiate to the oligodendrocyte [130].Ca2+, Mg2+,
L-tryptophan [130], spermine [131]
Neomycin [132], Vitamin D, Velcalcetide [133]Calcilytics, Phosphate [134], Ronacaleret [135], 2-methyl-3-phenethyl-3H-pyrimidin-4-one [136], compound (S)-3h [PMID: 15686947] [137], compound 17 [PMID: 15300839] [138], 1-arylmethylpyrrolidin-2-yl ethanol amine [139]
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

Díaz-Piña, D.A.; Rivera-Ramírez, N.; García-López, G.; Díaz, N.F.; Molina-Hernández, A. Calcium and Neural Stem Cell Proliferation. Int. J. Mol. Sci. 2024, 25, 4073. https://doi.org/10.3390/ijms25074073

AMA Style

Díaz-Piña DA, Rivera-Ramírez N, García-López G, Díaz NF, Molina-Hernández A. Calcium and Neural Stem Cell Proliferation. International Journal of Molecular Sciences. 2024; 25(7):4073. https://doi.org/10.3390/ijms25074073

Chicago/Turabian Style

Díaz-Piña, Dafne Astrid, Nayeli Rivera-Ramírez, Guadalupe García-López, Néstor Fabián Díaz, and Anayansi Molina-Hernández. 2024. "Calcium and Neural Stem Cell Proliferation" International Journal of Molecular Sciences 25, no. 7: 4073. https://doi.org/10.3390/ijms25074073

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