Next Article in Journal
Characterization of Human Medullary Thyroid Carcinoma Glycosphingolipids Identifies Potential Cancer Markers
Next Article in Special Issue
Lysosomal Exocytosis of Olivacine on the Way to Explain Drug Resistance in Cancer Cells
Previous Article in Journal
Bacterial Luciferases from Vibrio harveyi and Photobacterium leiognathi Demonstrate Different Conformational Stability as Detected by Time-Resolved Fluorescence Spectroscopy
Previous Article in Special Issue
Differential Serotonin Uptake Mechanisms at the Human Maternal–Fetal Interface
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Role of K+ and Ca2+-Permeable Channels in Osteoblast Functions

Department of Pharmacology, Graduate School of Medical Sciences, Nagoya City University, Nagoya 467-8601, Japan
*
Authors to whom correspondence should be addressed.
Int. J. Mol. Sci. 2021, 22(19), 10459; https://doi.org/10.3390/ijms221910459
Submission received: 10 September 2021 / Revised: 23 September 2021 / Accepted: 24 September 2021 / Published: 28 September 2021
(This article belongs to the Special Issue Channels and Transporters in Cells and Tissues 3.0)

Abstract

:
Bone-forming cells or osteoblasts play an important role in bone modeling and remodeling processes. Osteoblast differentiation or osteoblastogenesis is orchestrated by multiple intracellular signaling pathways (e.g., bone morphogenetic proteins (BMP) and Wnt signaling pathways) and is modulated by the extracellular environment (e.g., parathyroid hormone (PTH), vitamin D, transforming growth factor β (TGF-β), and integrins). The regulation of bone homeostasis depends on the proper differentiation and function of osteoblast lineage cells from osteogenic precursors to osteocytes. Intracellular Ca2+ signaling relies on the control of numerous processes in osteoblast lineage cells, including cell growth, differentiation, migration, and gene expression. In addition, hyperpolarization via the activation of K+ channels indirectly promotes Ca2+ signaling in osteoblast lineage cells. An improved understanding of the fundamental physiological and pathophysiological processes in bone homeostasis requires detailed investigations of osteoblast lineage cells. This review summarizes the current knowledge on the functional impacts of K+ channels and Ca2+-permeable channels, which critically regulate Ca2+ signaling in osteoblast lineage cells to maintain bone homeostasis.

1. Introduction

Bone modeling initially occurs during skeletal development and comprises two modes: intramembranous ossification and endochondral ossification [1]. During intramembranous ossification, mesenchymal cells are directly converted to osteoblasts, whereas during endochondral ossification, mesenchymal cells are differentiated into cartilage cells before conversion into osteoblasts [1]. Preosteoblasts derived from multipotent mesenchymal osteoprogenitor cells further differentiate into mature osteoblasts according to osteogenic signals. Mature osteoblasts are responsible for bone formation in bone modeling and remodeling. Therefore, the proliferation and differentiation of osteoblast lineage cells play an important role in the regulation of bone homeostasis [2].
Various growth factors and signaling molecules contribute to osteoblastogenesis. For instance, bone morphogenetic proteins (BMP) belonging to the transforming growth factor β (TGF-β) superfamily have a pivotal role in controlling osteoblast differentiation [3]. BMP signaling involves both canonical and non-canonical pathways. Canonical signaling is a SMAD-dependent pathway involving three types of SMADs: receptor-regulated SMADs (R-SMADs), common-mediator SMADs (Co-SMADs), and inhibitory-SMADs. The activation of type I BMP receptors phosphorylates R-SMADs (SMAD1, 5, and 8) to form a heterotrimeric complex with Co-SMAD, SMAD4. In the nucleus, this SMAD complex acts as a transcription factor to regulate the expression of BMP target genes [4]. On the other hand, non-canonical signaling activates SMAD-independent pathways, such as mitogen-activated protein kinase (MAPK) and phosphatidylinositol-3 kinase (PI3K)/Akt [5]. Additionally, integrins, which are heterodimeric transmembrane cell adhesion complexes composed of α and β chains, are expressed in osteoblasts and contribute to bone formation [6]. Integrins promote osteoblast differentiation by interacting with major extracellular matrix components, such as fibronectin and collagen type I (Col1) [7]. The activation of integrin signaling triggers osteogenesis, indicated by the induction of osteoblast genes, such as the transcription factor Runt-related transcription factor 2 (Runx2), phosphorylation of focal adhesion kinase (FAK), and extracellular signal-regulated kinase 1/2 (ERK1/2) [8]. Runx2 forms heterodimers with the core binding factor β (CBFβ) to control the expression of osteoblast marker genes including Col1a1, Spp1, Ibsp, Bglap2, and Fn1, and regulates multiple steps during osteoblast differentiation [9].
Intracellular Ca2+ ([Ca2+]i) signaling has a critical role in osteoblast proliferation and differentiation (Figure 1), mediating direct/indirect Ca2+-regulated proteins, such as Ca2+/calmodulin-dependent kinase II (CaMKII), a Ca2+/CaM-dependent protein phosphatase, calcineurin (Cn), and a Ca2+-dependent endoplasmic reticulum (ER) chaperone, calreticulin (CRT). Several studies have shown the significance of Ca2+ signaling in osteogenic functions. CaMKII activates cAMP response element-binding protein (CREB)/ activating transcription factor (ATF) and ERK signaling pathways in osteoblasts [10]. Cn controls bone formation and resorption by dephosphorylating the osteogenic transcription regulator, nuclear factor of activated T cells (NFAT) [11,12]. CRT regulates osteoblast differentiation via the NFAT signaling pathway [13]. In addition, the Ca2+ leak channel in the ER transmembrane and coiled-coil domains 1 (TMCO1) regulates osteoblast functions via the CaMKII-HDAC4-Runx2 signaling pathway [14]. The non-canonical Wnt signaling pathway is downstream of Ca2+ signaling and is involved in intramembranous and endochondral ossification [15]. Wnt-5a activates a non-canonical Wnt pathway by enhancing the activation of Ca2+-dependent protein kinase C and Ca2+/CaMKII and plays an integral role in BMP2-mediated osteogenic differentiation [16].
In many cases, [Ca2+]i mobilization is controlled by intracellular Ca2+ store activation and Ca2+ influx. Membrane hyperpolarization induced by K+ channel activation increases the electromotive driving force of Ca2+ entry through voltage-independent, store-operated, Ca2+-permeable channels in non-excitable cells [17,18,19]. Therefore, the membrane potential is one of the key biophysical signals in osteoblast lineage cells [20]. The present review summarizes the most recent evidence regarding K+ channels and Ca2+-permeable channels (Orai/stromal interaction protein (STIM), transient receptor potential (TRP), and mechano-sensing Piezo subfamilies) in osteoblast lineage cells, which are a group of cells comprising mesenchymal progenitors, preosteoblasts, osteoblasts (often called mature osteoblasts), bone-lining cells, and osteocytes, focusing on their role in skeletal development.

2. Regulatory Mechanism of Ca2+ Signaling by Membrane Potential in Non-Excitable Cell

[Ca2+]i is a ubiquitous second messenger regulating various cellular functions, such as proliferation, secretion, differentiation, migration, and apoptosis [21]. In quiescent cells, [Ca2+]i is maintained at very low levels (50–100 nM). The stimulation of cells with agonists, such as cytokines, hormones, or growth factors, results in [Ca2+]i rises via Ca2+ influx and/or Ca2+ release from the ER and mitochondria. In electrically non-excitable cells, including osteoblast lineage cells, membrane hyperpolarization increases the driving force for Ca2+ signaling. A direct link has been demonstrated between membrane hyperpolarization and Ca2+ signaling in epithelial cells [22], endothelial cells [23], cancer cells [19], microglia [24], and many types of immune cells [25]. The main Ca2+ entry route in these cells is not through voltage-gated Ca2+ channels but through Orai or TRP channels with high Ca2+ permeability. Membrane hyperpolarization promotes osteogenic differentiation in human mesenchymal stem cells (MSCs) [20], and the K+ channels play a pivotal role in the regulation of the membrane potential hyperpolarization. Therefore, a functional analysis of K+ channels may provide crucial insights into the functions of osteoblast lineage cells (Figure 2).

3. K+ Channel Superfamilies in Osteoblast Lineages

K+ channels are particularly important in maintaining the resting membrane potential and determining the shape and duration of the action potential. K+ channels may also regulate the cell volume, proliferation, differentiation, and migration of a wide range of cell types including osteoblast lineage cells (Table 1). Approximately 80 members are classified into voltage-gated K+ (KV) channel superfamily, inward-rectifier K+ (Kir) channel superfamily, Ca2+-activated K+ (KCa) channel superfamily, and two-pore domain K+ (K2P) channel superfamily [26] (Figure 3).
KV channel is a tetramer composed of four identical subunits consisting of six transmembrane (TM) (S1–S6) domains with a pore region (S5-P-S6), in which seven amino acidic residues, TTVGYGD that forms the structure of the ion-selectivity filter. KV channels are also characterized by containing a voltage-sensor domain in which the S4 contains positively charged amino acids that constitute the voltage-sensing elements. Kir channel is also a tetramer composed of four identical subunits consisting of two transmembrane domains connected by a pore domain, in which the ion-selectivity filter with the characteristic TXGYG signature sequence. KCa1.1 channels have seven transmembranes (S0–S6). KCa2 and KCa3.1 channels have six transmembranes (S1–S6) like KV channels but also contain an intracellular domain to bind CaM. The pore domain of the KCa1.1 channel is assigned to the region contained between S5 and S6 segments, which include the signature sequence of TVGYG. KCa2 and KCa3.1 channels have six TM segments and a pore loop region (between S5 and S6) containing the characteristic signature sequence GYGD. Then, K2P channel subunits are composed of four TM segments and two pore domains with the characteristic TxGy/FG motif; the 4TM/2P structure defines the membership in the K2P channel family [26].

3.1. Voltage-Gated K+ (KV) Channel Superfamily

The KV channel superfamily is the largest group of K+ channel superfamilies. KV channels represent a diverse group of membrane proteins with 12 distinct subfamilies (KV1–KV12). KV channels are well-known for regulating electrophysiological processes such as action potential generation and excitation-contraction coupling in excitable cells, such as neurons and myocytes [43]. In excitable cells, hyperpolarization followed by KV channel activation reduces [Ca2+]i levels by voltage-gated Ca2+ channel blockades, thereby inhibiting neurotransmission and muscle contraction. In contrast, in non-excitable cells, hyperpolarization followed by KV channel activation increases [Ca2+]i levels by facilitating Ca2+ influx through voltage-independent, store-operated Ca2+-permeable channels, thereby promoting cell proliferation, migration, and differentiation [43]. KV channels also influence cell cycle progression through cell volume regulation [43].
Yang et al. [27] showed that treatment with tetraethyl ammonium, a KV channel inhibitor, increased the matrix mineralization of human bone marrow-derived mesenchymal stem cells (BMMSCs) during osteogenic differentiation. The pharmacological blockade of KV7.3 with linopirdine augmented the mineralization during osteoblast differentiation of osteoblast-like MG-63 and Saos-2 cells through the release of glutamate [27]. Furthermore, several studies have shown that KV10.1 (also referred to as ether-à-go-go) enhanced the proliferation of MG-63 and Saos-2 cells [28,29]. The functional expression of the KV2.1 channel substantially contributed to 17β-estradiol-sensitive K+ currents in MG-63 cells [30].

3.2. Inward-Rectifier K+ (Kir) Channel Superfamily

The Kir channel superfamily is best known for its role in maintaining resting membrane potential in cardiac/vascular myocytes, neurons, and pancreatic β cells, regulating muscle contraction/relaxation, action potential firing, and insulin release [44]. Kir channels also regulate bone and cartilage development, as reported by Ozekin et al. [45].
Andersen–Tawil syndrome (ATS) is caused by mutations in the KCNJ2 gene, which encodes Kir2.1, in patients exhibiting cardiac arrhythmia, periodic paralysis, dental defects, cleft lip/palate, micrognathia, hypertelorism, low-set ears, and limb patterning defects. The disruption of Kir2.1 also leads to limb defects, hypoplastic craniofacial structures, and cleft palate in mice, craniofacial defects in Xenopus laevis, and morphological defects in Drosophila [46,47,48]. Recent studies showed that Kir2.1 disruption inhibited mammalian facial development by inhibiting BMP signaling [31]. Sacco et al. [32] demonstrated the role of Kir2.1 in osteoblast during osteogenesis. Osteoblasts derived from healthy control participants were able to produce bone matrix; however, osteoblasts from ATS patients, as well as in Kir2.1 knockout mice, were unable to synthesize bone matrix [32]. On the other hand, the loss of Kir2.1 function induced the dephosphorylation of SMAD and inhibited the BMP signaling pathway, resulting in impaired osteoblastogenesis [49].
ATP-sensitive K+ (KATP) channels, which are heteromultimers (1:1) of Kir6.x, and sulfonylurea receptors, regulate a variety of cellular functions by coupling cell metabolism with membrane potential [44]. Osteogenic differentiation of MSCs strongly upregulated Kir6.2, whereas Kir6.1 and SUR2A showed no significant change. The elevated Kir6.2 expression may contribute to in vitro differentiation of MSC according to the metabolic state of the developing tissue [33]. KATP channels are related to alkaline phosphatase activity and bone mineralization in rat primary osteoblast cells [34]. The activation of KATP channels protects osteoblasts from high-glucose-induced injury by promoting osteoblast proliferation and preventing apoptosis [34].

3.3. Ca2+-Activated K+ (KCa) Channel Superfamily

The KCa channel superfamily consists of eight members: KCa1.1/BK, KCa2.1/SK1, KCa2.2/SK2, KCa2.3/SK3, KCa3.1/SK4/IK, KCa4.1/KNa1.1, KCa4.2/KNa1.2, and KCa5.1. Due to sequence similarity in the pore region and C-terminus CaM-binding domain, the small-conductance (KCa2.1, KCa2.2, KCa2.3) and intermediate-conductance KCa (KCa3.1) channels belong to the same genetic lineage (KCNN genes) [50]. Distinct from KCa2.x and KCa3.1, the large-conductance KCa (KCa1.1) has voltage-dependent gating characteristics with sequence similarities to KV channels. KCa1.1 has unique Ca2+-binding domains at the C-terminus, RCK1/2 domains (referred to as “Ca2+ bowls”), required for Ca2+-dependent channel activation [50].
KCa1.1 is activated by depolarization and [Ca2+]i elevation. Due to these properties, the KCa1.1 activation results in repolarization and the closing of voltage-dependent Ca2+ channels in neurons and smooth muscle cells [51,52]. Therefore, KCa1.1 serves as a negative feedback regulator of membrane potential and [Ca2+]i in excitable cells. In contrast, KCa2.2 activation hyperpolarizes and thereby promotes Ca2+ influx through TRPC in brain capillary endothelial cells [53]. KCa3.1 activation results in elevated [Ca2+]i in human cancer cells [19,54] and immune cells [25]. Taken together, KCa channels have an important role in the positive feedback mechanism in Ca2+ signaling in non-excitable cells.
KCa1.1 especially plays a critical role in osteoclast activation. For instance, KCa1.1 deficiency in mice induced osteopenia due to enhanced osteoclast resorption subsequent to the autonomous release of cathepsin K [55]. KCa1.1 is also functionally expressed in human malignant osteoblast-like osteosarcoma cells (SaM-1, MG-63, and SaOS-2) and osteogenic precursor cells (C1) [35].
Hei et al. [36] showed that KCa1.1-null mice osteoblasts significantly decreased the ability for proliferation and mineralization, thereby decreasing bone mineral density and trabecular bone volume of the tibia and lumbar vertebrae. Integrin-linked kinases regulate bone formation and repair. KCa1.1 proteins interact with integrin β1 in osteoblasts, and the integrin-mediated activation of MAPK-ERK and FAK signaling pathways promotes osteogenesis-related expression of genes, such as Runx2. The blockade of integrin signaling by KCa1.1 knockdown in BMMSCs caused impaired bone formation and osteoblast differentiation [37].
Vascular calcification is a hallmark of cardiovascular disease with chronic kidney disease. One of the key events is the transition of contractile vascular smooth muscle cells (VSMCs) into a non-contractile osteoblast-like phenotype. The activation of KCa1.1 channels ameliorated vascular calcification via Akt/FoxO1 signaling pathways in calcified VSMCs [38]. On the other hand, KCa3.1 channels are expressed in proliferative undifferentiated VSMCs [56]. The pharmacological blockade of KCa3.1 inhibited the calcification medium (Ca2+/PO43−)-induced upregulation of osterix and osteocalcin in VSMCs and suppressed mineralization during VSMC calcification by simultaneously enhancing osteopontin expression [39] and by interfering with TGF-β and nuclear factor-kappa B (NF-κB) signaling [39].
Many studies have firmly established the important roles of vitamin D (VD) in calcium metabolism and bone development [57]. The VD endocrine system has a beneficial effect on bone homeostasis, but VD receptor (VDR) stimulation has direct positive or negative effects on bone mass depending on the osteoblast development stage. Yamamoto et al. [58] showed that VDR in osteoblasts served as a negative regulator of bone homeostasis due to osteoclastogenesis induced by the receptor activator of NF-κB ligand, RANKL. On the other hand, Sooy et al. [59] reported that osteoblast differentiation was promoted in calvarial osteoblasts obtained from VDR-deficient mice. Our recent study showed that the functional activity of KCa3.1 positively regulated the proliferation of mouse preosteoblast MC3T3-E1 cells by enhancing Ca2+ signaling [40]. KCa3.1 activity was suppressed in cells treated with VDR agonists following the downregulation of transcriptional/epigenetic modulators of KCa3.1, such as Fra-1 and HDAC2. Conclusively, a VDR agonist-induced decrease in KCa3.1 activity suppressed cell proliferation in mouse preosteoblasts. Therefore, the combined treatment of KCa3.1 activators and vitamin D preparations may enhance therapeutic effects on bone disorder involving bone mass reduction.

3.4. Two-Pore Domain K+ (K2P) Channel Superfamily

The background leak K2P channel subfamily is involved in diverse physiological functions such as ion homeostasis, hormone secretion, cell development, and excitability and is modulated by mechanical stretch, heat, intracellular/extracellular pH, lipids, and temperature [60]. The human K2P channel subfamily consists of 15 members possessing four transmembrane segments and two-pore-forming domains and assembling as dimers [61]. K2P2.1 (tandem of P-domains in a weakly inward rectifying K+ (TWIK)-related K+ channel 1, TREK1), a mechanosensitive K2P channel member, has been highly expressed in excitable cells [62]. TREK1 is also expressed in human primary osteoblasts and MG-63 cells and is involved in the maintenance of resting membrane potential in human osteoblasts [41]. Of the15 K2P members, five (K2P3.1 (TWIK-related acid-sensitive K+ channel 1, TASK1), K2P5.1 (TASK2), K2P9.1 (TASK3), K2P16.1 (TWIK-related alkaline pH-activated K+ channel 1, TALK1), and K2P17.1 (TALK2)) are known to be influenced by extra- and intracellular pH changes [63]. Acid-sensitive K2P channels (K2P3.1, K2P5.1, and K2P9.1), especially, are functionally expressed in MG-63 cells. The extracellular acidosis-induced inhibition of acid-sensitive K2P channel activity reduced MG-63 cell proliferation, indicating that these channels are associated with osteoblast proliferation [42]. The impacts of K2P channels on significant roles in native osteoblast lineages have not been established. However, Cid et al. [64] showed that K2P5.1 was expressed in hyaline cartilage from the trachea and the articular surface of the knee. Further studies are expected to provide evidence that K2P channels have physiological and pathophysiological roles in bone/cartilage homeostasis.

4. Ca2+-Permeable Channel Superfamilies in Osteoblast Lineages

Ca2+-permeable channels are present in non-excitable as well as in excitable cells [21] (Table 2). In osteoblast lineages, Ca2+-permeable channels play fundamental roles in cellular responses to external stimulation [10,11,12,13,14,15]. A store-operated calcium entry (SOCE), mediated by the activation of Orai/STIM, determines sustained [Ca2+]i increase, which is critical in regulating a variety of cellular functions, including proliferation and, more specifically, differentiation of osteoblast lineage cells (Table 3). Mechanical stimuli may also regulate the activation of mechanosensing Ca2+-permeable channels such as TRP (Table 4) and Piezo channels (Table 5).

4.1. Orai/STIM

Orai1 is a pore-forming subunit of the Ca2+ release-activated Ca2+ (CRAC) channel that mediates Ca2+ influx in most non-excitable cells via a store-operated Ca2+ entry (SOCE) mechanism [95]. [Ca2+]i changes are important signaling pathways in the regulation of osteoblast linage function [10], but causal roles remain highly controversial.
Studies on Orai1-deficient mice provided the first evidence of the importance of Orai1 for osteoblastic bone formation. The genetic deletion of Orai1 impaired bone development in the mice and resulted in osteopenia due to defective osteoblastic bone formation [65,66]. In addition, Choi et al. [67] showed that Orai1 contributed to the proliferation and differentiation of osteoblast lineage cells at various stages of maturation and that Orai1 deficiency attenuated the proliferation, differentiation, and Col1 secretion of mesenchymal progenitors and primary calvarial osteoblasts in Orai1-deficient mice. Orai1 deficiency also decreased the expression of an osteoid osteocyte marker gene, phosphate-regulating gene with homologies to endopeptidases on the X chromosome (Phex), a mineralizing osteocyte marker gene, dentin matrix protein 1 (DMP1), and a mature osteocyte marker gene, fibroblast growth factor 23 (FGF23) in mouse cortical long bone [67]. Taken together, Orai1 plays a crucial role in the differentiation of osteoblast lineage cells from mesenchymal progenitors into osteocytes.
Additionally, Liu et al. [68] showed that overexpression of Orai1 promoted human cartilage-derived MSC differentiation to osteoblasts. Lee et al. [69] reported on the Orai1 contribution to osteogenic effects of BMP2. Differentiation of BMMSCs into bone-forming osteoblasts requires the orchestrated regulation of signaling pathways, such as BMP signaling [96]. BMP signaling induces SMAD1/5/8-dependent signal transduction via BMP receptor types I and II, which activates the expression of Runx2, osterix, and osteoblast differentiation markers [97]. BMP-induced osteogenic conditions failed to induce increased osteogenic differentiation markers in BMMSCs of Orai1-deficient mice. These results suggest that Orai1 regulates osteogenic differentiation through BMP signaling. Interestingly, constitutively activated BMP signaling reversed the inhibition of osteogenic differentiation in Orai1-deficient mice. Collectively, BMP signaling is the downstream pathway of Orai1-mediated osteogenic differentiation.
FGF23, mainly produced by osteocytes and osteoblasts, is a pivotal regulator of renal functions such as phosphate/calcium handling and vitamin D3 homeostasis [98,99,100]. SOCE through Orai1 induces FGF23 production in rat osteoblast-like UMR106 cells [70,71]. Glosse et al. [101] showed that the blockade of SOCE decreased upregulated FGF23 expression induced by AMP-dependent kinase (AMPK) inhibition. Therefore, inhibitory AMPK-mediated FGF23 production may be regulated by Orai1-mediated SOCE activity in osteoblasts [101].
Peroxisome proliferator-activated receptors (PPARs) are nuclear transcriptional factors and are composed of PPARα, PPARβ/δ, and PPARγ [102]. PPARα is expressed in many organs, including bone [103]. Stimulation with a PPARα agonist induces AMPK activation in osteoblast lineage cells, resulting in a decrease in SOCE activity, whereas the pharmacological and genetic blockade of PPARα promotes FGF23 expression [104]. Thus, PPARα activation downregulated FGF23 expression through the AMPK-mediated suppression of SOCE in osteoblasts.
The ER Ca2+ sensor STIM1 is an essential component of the SOCE process. Overexpression of STIM1 in MC3T3-E1 cells promotes osteoblast differentiation and matrix mineralization with the upregulation of osteogenic markers, such as Runx2, Col1, and BMP4 [72]. Therefore, the loss of STIM1 function leading to SOCE dysregulation may be associated with impaired ER Ca2+ signaling in several skeletal- and tooth-related disorders [105]. Furthermore, gain-of-function mutations of Orai1 and STIM1 gave rise to tubular aggregate myopathy (TAM) and Stormorken syndrome (STRMK), forming a clinical spectrum encompassing muscle weakness, myalgia, and cramps, and additional multi-systemic signs, including short stature [106]. Silva-Rojas et al. [107] generated a mouse model carrying the most recurrent STIM1 gain-of-function mutation, STIM1 R304W, found in patients with TAM/STRMK. Moreover, microcomputed tomography analyses revealed a severe skeletal phenotype with thinner and more compact bone in STIM1 R304W mice [107,108]. Increased osteoclastogenesis was observed in bone marrow cells derived from patients with TAM/STRMK [109]. Bone formation and resorption are intricately balanced processes driven by bone-forming osteoblasts and bone-resorbing osteoclasts. Therefore, to gain further in vivo information on the contributions of SOCE activity to osteoblastogenesis, the role of SOCE activity in osteoclastogenesis needs to be comprehensively examined.

4.2. TRP Channel Superfamilies

TRP channels are voltage-independent, Ca2+-permeable, non-selective cation channels. Approximately 30 members are classified into TRPC (canonical superfamily), TRPV (vanilloid superfamily), TRPM (melastatin superfamily), TRPA (ankyrin superfamily), TRPN (NompC superfamily), TRPML (mucolipin superfamily), and TRPP (polycystin superfamily) [110]. TRP channels are activated by many physical or chemical stimuli (e.g., temperature, membrane potential, and pH). Members of the TRPM, TRPV, and TRPP subfamilies are involved in extracellular calcium homeostasis and intracellular Ca2+ signaling in osteoblast lineage cells [111].

4.2.1. TRPM Channels

Shear stress is physiologically generated by blood flow and interstitial fluid [112]. Fluid flow is a potent stimulator of osteoblasts and generally can be of three types: (1) steady (fixed in magnitude and direction); (2) pulsating (oscillating in magnitude but fixed in direction); and (3) oscillatory (oscillating in magnitude and periodically reversing direction). Roy et al. [73] demonstrated that oscillatory fluid flow, but not steady and pulsating flow, induced high Ca2+ flicker activity at Ca2+ microdomains with a localized high [Ca2+]i in MG-63 cells. The localization of TRPM7 in lipid rafts as Ca2+ microdomains was critical for Ca2+ flicker activity [73]. In human MSCs, TRPM7 stimulated by fluid flow promoted [Ca2+]i, resulting in osteogenic differentiation [74,75].
Clinical pulsed electromagnetic field stimulation is widely applied to promote bone regeneration in both the United States and Europe as a therapeutic approach for patients with musculoskeletal disorders [113]. Although the underlying mechanisms of electromagnetically induced osteogenesis are not yet completely elucidated, several studies have indicated the benefit of applying electromagnetic stimulation in bone-healing processes. Magnetically induced electrostimulation is an attractive approach to osteogenic differentiation [114] and osteoporosis prevention [115]. Indeed, the stimulation by electric fields upregulated TRPM7 expression, leading to the migration of human osteoblasts from femoral heads of patients undergoing total hip replacement [76].
Mechanical stresses regulate bone formation by upregulating RANKL expression in osteoblasts. Acute mechanical stress induced by hypoosmotic shock promoted TRPM3 and TRPV4 activation, resulting in increased Ca2+-mediated RANKL and NFATc1 expressions in primary mouse calvarial osteoblasts [77]. NFATc1 is required for parathyroid hormone (PTH)-related peptide-induced RANKL expression in C2C12 and primary-cultured mouse calvarial cells [116]. The [Ca2+]i rise through TRPM3 and TRPV4 as a mechanosensor are critical to NFATc1 activity and the subsequent physiological activity of osteoblasts.

4.2.2. TRPV Channels

TRPV1 deficiency results in the reduced osteogenic differentiation potential of BMMSCs [78]. On the other hand, bone marrow cells derived from TRPV1- and TRPV4-deficient mice showed enhanced osteoblast differentiation activity [79]. Gain-of-function mutations of TRPV4 induced congenital skeletal dysplasias, including metatropic dysplasia (MD) [117]. Patients with MD exhibit short limbs at birth due to a defect in long bone development and progressive kyphoscoliosis. Nonaka et al. [118] reported that the novel gain-of-function mutation TRPV4 L619F was involved in the chondrogenic differentiation of MSCs and accelerated the osteogenic differentiation of MSCs obtained from patients with MD [80]. Osteoblasts expressing TRPV4 L619F exhibited increased [Ca2+]i, levels; thus, calcification was enhanced with upregulated Runx2 and osteocalcin expression. Indeed, the expression and nuclear translocation of NFATc1 were upregulated in osteoblasts expressing TRPV4 L619F. These findings indicate that MD-associated disorganized endochondral ossification is induced by disordered osteogenic differentiation through the TRPV4-Ca2+/NFATc1 signaling axis [80].
A biologically active N-terminal fragment of human PTH, PTH (1-34), is used clinically to increase bone volume in osteoporosis. Pozo et al. [81] showed that PTH (1-34) promoted Ca2+ influx through TRPV4 in MG-63 cells via a PKA-dependent pathway. Interestingly, the PTH-induced Ca2+ influx significantly inhibited MG-63 cell migration. When PTH (1-34) is administered to humans or rodents, bone mass was increased due to an increased number of osteoblasts and bone formation rate [119]. PTH (1-34) also suppressed osteoblast apoptosis [120] and converted quiescent bone-lining cells into active osteoblasts [121]. Further studies will be needed to clarify the relationship between cytosolic Ca2+ dynamics and osteoblast functions via PTH stimulation.

4.2.3. TRPP Channels

Autosomal dominant polycystic kidney disease is caused by inactivating mutations in the gene that encode either TRPP1 or TRPP2 [122]. TRPP1 binds to TRPP2 through their respective C-terminal coiled-coil domains to form the functional TRPP complex in the cell membrane [123]. This complex between TRPP1 and TRPP2 functions as a mechanosensor in bone and kidney [124].
Recent studies indicated that the TRPP complex played a pivotal role in osteoblastogenesis [82,83,85,86,125]. TRPP1 has been shown as a mechanosensing molecule involved in osteoblastogenesis and bone remodeling. Dalagiorgou et al. [125] reported that mechanical stress upregulated Runx2 expression by activating the TRPP1-JAK2/STAT3 signaling pathway, resulting in osteoblastic differentiation in human osteoblastic cells and, ultimately, bone formation. In addition, TRPP1 and the transcriptional coactivator TAZ formed a mechanosensing complex and contributed to osteoblastic differentiation in MSCs [82,83].
Osteoporosis is a disease characterized by low bone volume, skeletal fragility, and an increased risk of fracture. Bicaudal C homolog 1 (Bicc1) is one of the genetic determinants of patient bone mineral density, and Bicc1-deficient mice had low bone mineral density. Mesner et al. [85] indicated that TRPP2 was a downstream target of Bicc1 involved in osteoblast differentiation. In addition, a human genome-wide association study (GWAS) meta-analysis showed that single nucleotide polymorphisms (SNPs) in TRPP2 are associated with decreased bone mineral density. These findings suggest that Bicc1 is a critical determinant of osteoblastogenesis and bone mineral density by regulating TRPP2 transcription.
The osteoblast-specific depletion of TRPP2 resulted in reduced bone volume associated with decreased bone mineral density, mineral apposition rate, trabecular bone volume, and cortical thickness, and impaired biomechanical properties of bone. TRPP2-deficient osteoblasts exhibited lower basal [Ca2+]i and impaired response to flow-induced Ca2+ influx [86]. Additionally, the short hairpin RNA (shRNA)-mediated blockade of TRPP1 suppressed Ca2+ signaling to shear fluid stress in MG-63 cells [84]. These studies suggest that Ca2+-permeable channel TRPP2 is coupled to mechanosensing TRPP1 in response to mechanical stress-loaded osteoblasts during osteoblast differentiation.

4.3. Piezo Channels

Piezo channels are emerging as important regulators of various aspects of mechanosensing [126,127,128]. Piezo2 is predominantly expressed in sensory neurons, such as dorsal root ganglia neurons, whereas Piezo1 is primarily expressed in non-sensory tissues and non-neuronal cells and senses various mechanical stimuli, including shear stress, static pressure, and membrane stretch. Piezo channels are permeable to divalent ions such as Ca2+, as well as to monovalent ions [129]. Recent GWAS reports showed that human Piezo1 SNPs were associated with shorter adult height [130] and reduced bone mineral density [131]. Sugimoto et al. [87] showed that hydrostatic pressure induced the differentiation of BMMSCs into osteoblasts with enhanced BMP2 expression via the ERK1/2 and p38 MAPK signaling through Piezo1. In MC3T3-E1 cells, the combined action of both Piezo1 and TRPV4 is involved in the sensing of mechanical fluid shear stress [88], and Piezo1 is essential for Runx2 expression through the Akt/GSK-3β/β-catenin signaling pathway [89]. Indeed, Piezo1 is highly expressed in various bone tissues of mice and plays a critical role in bone formation [90] (Figure 4).
Osteoblast lineage-specific, Piezo1-deficient mice (Piezo1OcnCre) showed severely impaired bone formation and blunt, mechanical unloading-induced bone loss [90]. Moreover, Piezo1-conditional-deficient mice, Piezo1Prrx1Cre, exhibited multiple skeletal fractures in long bones, the radius, and ulna; however, Piezo2 conditional-deficient mice (Piezo2Prrx1Cre) had normal bone development with no fractures. Piezo1/2 double conditional knockout neonatal mice exhibited more severe skeletal defects in the appendicular skeleton [94]. Piezo1 and Piezo1/2 knockout mice showed decreased activity in the Hippo-Yap1/TAZ and Wnt/β-catenin signaling pathway in osteoblasts, thus promoting osteoblastogenic differentiation [94]. In addition, Li et al. [91] reported that osteoblast/osteocyte-specific deletion of Piezo1 decreased bone formation and bone mass in other Piezo1 knockout mice, Piezo1Dmp1Cre. The shRNA-mediated blockade of Piezo1 in MLO-Y4 osteocyte-like cells reduced the expression of Ptgs2 and Tnfrsf11b [89], well-known targets of fluid shear stress in osteocytes [132,133]. Furthermore, Sasaki et al. [92] reported that mechanical stimulation suppressed an osteocyte-derived negative regulator of bone formation, sclerostin mRNA expression, in murine osteocyte IDG-SW3 cells through the Piezo1-Akt pathway.
Piezo1 knockout mice, Piezo1Runx2Cre, demonstrated a more severe osteoporotic phenotype than Piezo1Dmp1Cre [93]. Because Runx2 was expressed in chondrocytes, as well as in osteoblasts, Piezo1 activity in chondrocytes was involved in this aggravation of osteoporosis in Piezo1Runx2Cre mice [132]. Lee et al. [134] have described the presence of mechanosensory Piezo channels in chondrocytes, which function synergistically in response to injurious mechanical loading.
Wang et al. [135] reported that depletion of osteoblastic Piezo1 promoted bone resorption in Piezo1Prx1Cre mice. Piezo1 in osteoblasts regulated the YAP-dependent expression of Col types II and IV, which control osteoclast differentiation, suggesting that Piezo1 can coordinate the crosstalk between osteoblasts and osteoclasts by directly sensing mechanical stress in osteoblast lineage cells.

5. Conclusions/Future Perspectives

Osteoporosis is a major skeletal disorder that influences bone structure and composition. Gradual bone loss results in bone fragility and significantly increases the risk of bone fracture. An estimated 1.5 million fractures occur per year in the United States [136]. Bone homeostasis depends on intracellular Ca2+ signaling, as well as on the external calcium balance that regulates the function and differentiation of chondrocytes, osteoclasts, and osteoblast lineage cells. A number of ion channels critically contribute to the regulation of these processes, including intestinal calcium absorption and renal calcium reabsorption [137].
This review has highlighted the critical role of K+ channels and Ca2+-permeable channels in osteoblast functions. However, the lack of understanding of their specific functions in bone formation poses a definite limit. Elucidating their involvement in bone signaling pathways could help identify molecules targeting bone biology. Emphasis should be given on data describing the cooperative coupling of K+ and Ca2+ signaling in regulating the physiological and pathophysiological frameworks of bone homeostasis.

Author Contributions

Conceptualization, H.K. and S.O.; writing—original draft preparation, H.K.; writing—review and editing, S.O.; visualization, H.K.; supervision, S.O.; project administration, H.K. and S.O; funding acquisition, H.K. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by JSPS KAKENHI, grant numbers JP18KK0218 and JP21K06600.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Berendsen, A.D.; Olsen, B.R. Bone development. Bone 2015, 80, 14–18. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. Salhotra, A.; Shah, H.N.; Levi, B.; Longaker, M.T. Mechanisms of bone development and repair. Nat. Rev. Mol. Cell Biol. 2020, 21, 696–711. [Google Scholar] [CrossRef] [PubMed]
  3. Wu, M.; Chen, G.; Li, Y.P. TGF-b and BMP signaling in osteoblast, skeletal development, and bone formation, homeostasis and disease. Bone Res. 2016, 4, 16009. [Google Scholar] [CrossRef] [PubMed]
  4. Horbelt, D.; Denkis, A.; Knaus, P. A portrait of Transforming Growth Factor b superfamily signalling: Background matters. Int. J. Biochem. Cell Biol. 2012, 44, 469–474. [Google Scholar] [CrossRef]
  5. Zhang, Y.E. Non-Smad Signaling Pathways of the TGF-b Family. Cold Spring Harb. Perspect. Biol. 2017, 9, a022129. [Google Scholar] [CrossRef] [PubMed]
  6. Marie, P.J. Targeting integrins to promote bone formation and repair. Nat. Rev. Endocrinol. 2013, 9, 288–295. [Google Scholar] [CrossRef]
  7. Damsky, C.H. Extracellular matrix-integrin interactions in osteoblast function and tissue remodeling. Bone 1999, 25, 95–96. [Google Scholar] [CrossRef]
  8. Yuh, D.Y.; Maekawa, T.; Li, X.; Kajikawa, T.; Bdeir, K.; Chavakis, T.; Hajishengallis, G. The secreted protein DEL-1 activates a b3 integrin-FAK-ERK1/2-RUNX2 pathway and promotes osteogenic differentiation and bone regeneration. J. Biol. Chem. 2020, 295, 7261–7273. [Google Scholar] [CrossRef] [Green Version]
  9. Komori, T. Regulation of Proliferation, Differentiation and Functions of Osteoblasts by Runx2. Int. J. Mol. Sci. 2019, 20, 1694. [Google Scholar] [CrossRef] [Green Version]
  10. Zayzafoon, M.; Fulzele, K.; McDonald, J.M. Calmodulin and calmodulin-dependent kinase IIalpha regulate osteoblast differentiation by controlling c-fos expression. J. Biol. Chem. 2005, 280, 7049–7059. [Google Scholar] [CrossRef] [Green Version]
  11. Sun, L.; Blair, H.C.; Peng, Y.; Zaidi, N.; Adebanjo, O.A.; Wu, X.B.; Wu, X.Y.; Iqbal, J.; Epstein, S.; Abe, E.; et al. Calcineurin regulates bone formation by the osteoblast. Proc. Natl. Acad. Sci. USA 2005, 102, 17130–17135. [Google Scholar] [CrossRef] [Green Version]
  12. Winslow, M.M.; Pan, M.; Starbuck, M.; Gallo, E.M.; Deng, L.; Karsenty, G.; Crabtree, G.R. Calcineurin/NFAT signaling in osteoblasts regulates bone mass. Dev. Cell 2006, 10, 771–782. [Google Scholar] [CrossRef] [Green Version]
  13. Pilquil, C.; Alvandi, Z.; Opas, M. Calreticulin regulates a switch between osteoblast and chondrocyte lineages derived from murine embryonic stem cells. J. Biol. Chem. 2020, 295, 6861–6875. [Google Scholar] [CrossRef] [Green Version]
  14. Li, J.; Liu, C.; Li, Y.; Zheng, Q.; Xu, Y.; Liu, B.; Sun, W.; Li, Y.; Ji, S.; Liu, M.; et al. TMCO1-mediated Ca2+ leak underlies osteoblast functions via CaMKII signaling. Nat. Commun. 2019, 10, 1589. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  15. Houschyar, K.S.; Tapking, C.; Borrelli, M.R.; Popp, D.; Duscher, D.; Maan, Z.N.; Chelliah, M.P.; Li, J.; Harati, K.; Wallner, C.; et al. Wnt Pathway in Bone Repair and Regeneration—What Do We Know So Far. Front. Cell Dev. Biol. 2019, 6, 170. [Google Scholar] [CrossRef] [PubMed]
  16. Nemoto, E.; Ebe, Y.; Kanaya, S.; Tsuchiya, M.; Nakamura, T.; Tamura, M.; Shimauchi, H. Wnt5a signaling is a substantial constituent in bone morphogenetic protein-2-mediated osteoblastogenesis. Biochem. Biophys. Res. Commun. 2012, 422, 627–632. [Google Scholar] [CrossRef]
  17. Nakakura, S.; Matsui, M.; Sato, A.; Ishii, M.; Endo, K.; Muragishi, S.; Murase, M.; Kito, H.; Niguma, H.; Kurokawa, N.; et al. Pathophysiological significance of the two-pore domain K+ channel K2P5.1 in splenic CD4+CD25 T cell subset from a chemically-induced murine inflammatory bowel disease model. Front. Physiol. 2015, 6, 299. [Google Scholar] [CrossRef] [Green Version]
  18. Kito, H.; Yamamura, H.; Suzuki, Y.; Ohya, S.; Asai, K.; Imaizumi, Y. Membrane hyperpolarization induced by endoplasmic reticulum stress facilitates Ca2+ influx to regulate cell cycle progression in brain capillary endothelial cells. J. Pharmacol. Sci. 2014, 125, 227–232. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  19. Ohya, S.; Kanatsuka, S.; Hatano, N.; Kito, H.; Matsui, A.; Fujimoto, M.; Matsuba, S.; Niwa, S.; Zhan, P.; Suzuki, T.; et al. Downregulation of the Ca2+-activated K+ channel KCa3.1 by histone deacetylase inhibition in human breast cancer cells. Pharmacol. Res. Perspect. 2016, 4, e00228. [Google Scholar] [CrossRef] [PubMed]
  20. Sundelacruz, S.; Levin, M.; Kaplan, D.L. Membrane potential controls adipogenic and osteogenic differentiation of mesenchymal stem cells. PLoS ONE 2008, 3, e3737. [Google Scholar] [CrossRef] [Green Version]
  21. 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]
  22. Ohba, T.; Sawada, E.; Suzuki, Y.; Yamamura, H.; Ohya, S.; Tsuda, H.; Imaizumi, Y. Enhancement of Ca2+ influx and ciliary beating by membrane hyperpolarization due to ATP-sensitive K+ channel opening in mouse airway epithelial cells. J. Pharmacol. Exp. Ther. 2013, 347, 145–153. [Google Scholar] [CrossRef] [Green Version]
  23. Yamazaki, D.; Kito, H.; Yamamoto, S.; Ohya, S.; Yamamura, H.; Asai, K.; Imaizumi, Y. Contribution of Kir2 potassium channels to ATP-induced cell death in brain capillary endothelial cells and reconstructed HEK293 cell model. Am. J. Physiol. Cell Physiol. 2011, 300, C75–C86. [Google Scholar] [CrossRef] [PubMed]
  24. Laprell, L.; Schulze, C.; Brehme, M.L.; Oertner, T.G. The role of microglia membrane potential in chemotaxis. J. Neuroinflamm. 2021, 18, 21. [Google Scholar] [CrossRef] [PubMed]
  25. Feske, S.; Wulff, H.; Skolnik, E.Y. Ion channels in innate and adaptive immunity. Annu. Rev. Immunol. 2015, 33, 291–353. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  26. Gonzalez, C.; Baez-Nieto, D.; Valencia, I.; Oyarzun, I.; Rojas, P.; Naranjo, D.; Latorre, R. K+ channels: Function-structural overview. Compr. Physiol. 2012, 2, 2087–2149. [Google Scholar]
  27. Yang, J.E.; Song, M.S.; Shen, Y.; Ryu, P.D.; Lee, S.Y. The Role of KV7.3 in Regulating Osteoblast Maturation and Mineralization. Int. J. Mol. Sci. 2016, 17, 407. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  28. Wu, J.; Zhong, D.; Fu, X.; Liu, Q.; Kang, L.; Ding, Z. Silencing of Ether a go-go 1 by shRNA inhibits osteosarcoma growth and cell cycle progression. Int. J. Mol. Sci. 2014, 15, 5570–5581. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  29. Wu, X.; Zhong, D.; Gao, Q.; Zhai, W.; Ding, Z.; Wu, J. MicroRNA-34a inhibits human osteosarcoma proliferation by downregulating ether a go-go 1 expression. Int. J. Med. Sci. 2013, 10, 676–682. [Google Scholar] [CrossRef] [Green Version]
  30. Li, X.; Zheng, S.; Dong, X.; Xiao, J. 17β-Estradiol inhibits outward voltage-gated K⁺ currents in human osteoblast-like MG63 cells. J. Membr. Biol. 2013, 246, 39–45. [Google Scholar] [CrossRef]
  31. Belus, M.T.; Rogers, M.A.; Elzubeir, A.; Josey, M.; Rose, S.; Andreeva, V.; Yelick, P.C.; Bates, E.A. Kir2.1 is important for efficient BMP signaling in mammalian face development. Dev. Biol. 2018, 444 (Suppl. 1), S297–S307. [Google Scholar] [CrossRef]
  32. Sacco, S.; Giuliano, S.; Sacconi, S.; Desnuelle, C.; Barhanin, J.; Amri, E.Z.; Bendahhou, S. The inward rectifier potassium channel Kir2.1 is required for osteoblastogenesis. Hum. Mol. Genet. 2015, 24, 471–479. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  33. Diehlmann, A.; Bork, S.; Saffrich, R.; Veh, R.W.; Wagner, W.; Derst, C. KATP channels in mesenchymal stromal stem cells: Strong up-regulation of Kir6.2 subunits upon osteogenic differentiation. Tissue Cell 2011, 43, 331–336. [Google Scholar] [CrossRef]
  34. Liu, Y.; Liu, J.; Li, X.; Wang, F.; Xu, X.; Wang, C. Exogenous H2S prevents high glucose-induced damage to osteoblasts through regulation of KATP channels. Biochimie 2017, 137, 151–157. [Google Scholar] [CrossRef] [PubMed]
  35. Hirukawa, K.; Muraki, K.; Ohya, S.; Imaizumi, Y.; Togari, A. Electrophysiological properties of a novel Ca2+-activated K+ channel expressed in human osteoblasts. Calcif. Tissue Int. 2008, 83, 222–229. [Google Scholar] [CrossRef]
  36. Hei, H.; Gao, J.; Dong, J.; Tao, J.; Tian, L.; Pan, W.; Wang, H.; Zhang, X. BK Knockout by TALEN-Mediated Gene Targeting in Osteoblasts: KCNMA1 Determines the Proliferation and Differentiation of Osteoblasts. Mol. Cells 2016, 39, 530–535. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  37. Wang, Y.; Guo, Q.; Hei, H.; Tao, J.; Zhou, Y.; Dong, J.; Xin, H.; Cai, H.; Gao, J.; Yu, K.; et al. BK ablation attenuates osteoblast bone formation via integrin pathway. Cell Death Dis. 2019, 10, 738. [Google Scholar] [CrossRef] [PubMed]
  38. Ning, F.L.; Tao, J.; Li, D.D.; Tian, L.L.; Wang, M.L.; Reilly, S.; Liu, C.; Cai, H.; Xin, H.; Zhang, X.M. Activating BK channels ameliorates vascular smooth muscle calcification through Akt signaling. Acta Pharmacol. Sin. 2021, 1–10. [Google Scholar] [CrossRef]
  39. Freise, C.; Querfeld, U. Inhibition of vascular calcification by block of intermediate conductance calcium-activated potassium channels with TRAM-34. Pharmacol. Res. 2014, 85, 6–14. [Google Scholar] [CrossRef]
  40. Kito, H.; Morihiro, H.; Sakakibara, Y.; Endo, K.; Kajikuri, J.; Suzuki, T.; Ohya, S. Downregulation of the Ca2+-activated K+ channel KCa3.1 in mouse preosteoblast cells treated with vitamin D receptor agonist. Am. J. Physiol. Cell Physiol. 2020, 319, C345–C358. [Google Scholar] [CrossRef]
  41. Hughes, S.; Magnay, J.; Foreman, M.; Publicover, S.J.; Dobson, J.P.; El Haj, A.J. Expression of the mechanosensitive 2PK+ channel TREK-1 in human osteoblasts. J. Cell. Physiol. 2006, 206, 738–748. [Google Scholar] [CrossRef]
  42. Li, X.; Dong, X.; Zheng, S.; Xiao, J. Expression and localization of TASK-1, -2 and -3 channels in MG63 human osteoblast-like cells. Oncol. Lett. 2013, 5, 865–869. [Google Scholar] [CrossRef]
  43. Gutman, G.A.; Chandy, K.G.; Grissmer, S.; Lazdunski, M.; McKinnon, D.; Pardo, L.A.; Robertson, G.A.; Rudy, B.; Sanguinetti, M.C.; Stuhmer, W.; et al. International Union of Pharmacology. LIII. Nomenclature and molecular relationships of voltage-gated potassium channels. Pharmacol. Rev. 2005, 57, 473–508. [Google Scholar] [CrossRef] [PubMed]
  44. Hibino, H.; Inanobe, A.; Furutani, K.; Murakami, S.; Findlay, I.; Kurachi, Y. Inwardly rectifying potassium channels: Their structure, function, and physiological roles. Physiol. Rev. 2010, 90, 291–366. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  45. Ozekin, Y.H.; Isner, T.; Bates, E.A. Ion Channel Contributions to Morphological Development: Insights from the Role of Kir2.1 in Bone Development. Front. Mol. Neurosci. 2020, 13, 99. [Google Scholar] [CrossRef] [PubMed]
  46. Adams, D.S.; Uzel, S.G.; Akagi, J.; Wlodkowic, D.; Andreeva, V.; Yelick, P.C.; Devitt-Lee, A.; Pare, J.F.; Levin, M. Bioelectric signalling via potassium channels: A mechanism for craniofacial dysmorphogenesis in KCNJ2-associated Andersen-Tawil Syndrome. J. Physiol. 2016, 594, 3245–3270. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  47. Dahal, G.R.; Rawson, J.; Gassaway, B.; Kwok, B.; Tong, Y.; Ptacek, L.J.; Bates, E. An inwardly rectifying K+ channel is required for patterning. Development 2012, 139, 3653–3664. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  48. Zaritsky, J.J.; Eckman, D.M.; Wellman, G.C.; Nelson, M.T.; Schwarz, T.L. Targeted disruption of Kir2.1 and Kir2.2 genes reveals the essential role of the inwardly rectifying K+ current in K+-mediated vasodilation. Circ. Res. 2000, 87, 160–166. [Google Scholar] [CrossRef] [Green Version]
  49. Pini, J.; Giuliano, S.; Matonti, J.; Gannoun, L.; Simkin, D.; Rouleau, M.; Bendahhou, S. Osteogenic and Chondrogenic Master Genes Expression Is Dependent on the Kir2.1 Potassium Channel Through the Bone Morphogenetic Protein Pathway. J. Bone Miner. Res. 2018, 33, 1826–1841. [Google Scholar] [CrossRef]
  50. Wei, A.D.; Gutman, G.A.; Aldrich, R.; Chandy, K.G.; Grissmer, S.; Wulff, H. International Union of Pharmacology. LII. Nomenclature and molecular relationships of calcium-activated potassium channels. Pharmacol. Rev. 2005, 57, 463–472. [Google Scholar] [CrossRef] [Green Version]
  51. Berkefeld, H.; Fakler, B.; Schulte, U. Ca2+-activated K+ channels: From protein complexes to function. Physiol. Rev. 2010, 90, 1437–1459. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  52. Bolton, T.B.; Imaizumi, Y. Spontaneous transient outward currents in smooth muscle cells. Cell Calcium 1996, 20, 141–152. [Google Scholar] [CrossRef]
  53. Yamazaki, D.; Aoyama, M.; Ohya, S.; Muraki, K.; Asai, K.; Imaizumi, Y. Novel functions of small conductance Ca2+-activated K+ channel in enhanced cell proliferation by ATP in brain endothelial cells. J. Biol. Chem. 2006, 281, 38430–38439. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  54. Matsui, M.; Kajikuri, J.; Kito, H.; Endo, K.; Hasegawa, Y.; Murate, S.; Ohya, S. Inhibition of Interleukin 10 Transcription through the SMAD2/3 Signaling Pathway by Ca2+-Activated K+ Channel KCa3.1 Activation in Human T-Cell Lymphoma HuT-78 Cells. Mol. Pharmacol. 2019, 95, 294–302. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  55. Sausbier, U.; Dullin, C.; Missbach-Guentner, J.; Kabagema, C.; Flockerzie, K.; Kuscher, G.M.; Stuehmer, W.; Neuhuber, W.; Ruth, P.; Alves, F.; et al. Osteopenia due to enhanced cathepsin K release by BK channel ablation in osteoclasts. PLoS ONE 2011, 6, e21168. [Google Scholar]
  56. Bi, D.; Toyama, K.; Lemaitre, V.; Takai, J.; Fan, F.; Jenkins, D.P.; Wulff, H.; Gutterman, D.D.; Park, F.; Miura, H. The intermediate conductance calcium-activated potassium channel KCa3.1 regulates vascular smooth muscle cell proliferation via controlling calcium-dependent signaling. J. Biol. Chem. 2013, 288, 15843–15853. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  57. St-Arnaud, R. The direct role of vitamin D on bone homeostasis. Arch. Biochem. Biophys. 2008, 473, 225–230. [Google Scholar] [CrossRef] [PubMed]
  58. Yamamoto, Y.; Yoshizawa, T.; Fukuda, T.; Shirode-Fukuda, Y.; Yu, T.; Sekine, K.; Sato, T.; Kawano, H.; Aihara, K.; Nakamichi, Y.; et al. Vitamin D receptor in osteoblasts is a negative regulator of bone mass control. Endocrinology 2013, 154, 1008–1020. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  59. Sooy, K.; Sabbagh, Y.; Demay, M.B. Osteoblasts lacking the vitamin D receptor display enhanced osteogenic potential in vitro. J. Cell. Biochem. 2005, 94, 81–87. [Google Scholar] [CrossRef] [PubMed]
  60. Feliciangeli, S.; Chatelain, F.C.; Bichet, D.; Lesage, F. The family of K2P channels: Salient structural and functional properties. J. Physiol. 2015, 593, 2587–2603. [Google Scholar] [CrossRef] [Green Version]
  61. Enyedi, P.; Czirjak, G. Molecular background of leak K+ currents: Two-pore domain potassium channels. Physiol. Rev. 2010, 90, 559–605. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  62. Djillani, A.; Mazella, J.; Heurteaux, C.; Borsotto, M. Role of TREK-1 in Health and Disease, Focus on the Central Nervous System. Front. Pharmacol. 2019, 10, 379. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  63. Endo, K.; Kurokawa, N.; Kito, H.; Nakakura, S.; Fujii, M.; Ohya, S. Molecular identification of the dominant-negative, splicing isoform of the two-pore domain K+ channel K2P5.1 in lymphoid cells and enhancement of its expression by splicing inhibition. Biochem. Pharmacol. 2015, 98, 440–452. [Google Scholar] [CrossRef]
  64. Cid, L.P.; Roa-Rojas, H.A.; Niemeyer, M.I.; Gonzalez, W.; Araki, M.; Araki, K.; Sepulveda, F.V. TASK-2: A K2P K+ channel with complex regulation and diverse physiological functions. Front. Physiol. 2013, 4, 198. [Google Scholar] [CrossRef] [Green Version]
  65. Hwang, S.Y.; Foley, J.; Numaga-Tomita, T.; Petranka, J.G.; Bird, G.S.; Putney, J.W., Jr. Deletion of Orai1 alters expression of multiple genes during osteoclast and osteoblast maturation. Cell Calcium 2012, 52, 488–500. [Google Scholar] [CrossRef] [Green Version]
  66. Robinson, L.J.; Mancarella, S.; Songsawad, D.; Tourkova, I.L.; Barnett, J.B.; Gill, D.L.; Soboloff, J.; Blair, H.C. Gene disruption of the calcium channel Orai1 results in inhibition of osteoclast and osteoblast differentiation and impairs skeletal development. Lab. Investig. 2012, 92, 1071–1083. [Google Scholar] [CrossRef]
  67. Choi, H.; Srikanth, S.; Atti, E.; Pirih, F.Q.; Nervina, J.M.; Gwack, Y.; Tetradis, S. Deletion of Orai1 leads to bone loss aggravated with aging and impairs function of osteoblast lineage cells. Bone Rep. 2018, 8, 147–155. [Google Scholar] [CrossRef]
  68. Liu, S.; Takahashi, M.; Kiyoi, T.; Toyama, K.; Mogi, M. Genetic Manipulation of Calcium Release-Activated Calcium Channel 1 Modulates the Multipotency of Human Cartilage-Derived Mesenchymal Stem Cells. J. Immunol. Res. 2019, 2019, 7510214. [Google Scholar] [CrossRef]
  69. Lee, S.H.; Park, Y.; Song, M.; Srikanth, S.; Kim, S.; Kang, M.K.; Gwack, Y.; Park, N.H.; Kim, R.H.; Shin, K.H. Orai1 mediates osteogenic differentiation via BMP signaling pathway in bone marrow mesenchymal stem cells. Biochem. Biophys. Res. Commun. 2016, 473, 1309–1314. [Google Scholar] [CrossRef] [Green Version]
  70. Zhang, B.; Yan, J.; Umbach, A.T.; Fakhri, H.; Fajol, A.; Schmidt, S.; Salker, M.S.; Chen, H.; Alexander, D.; Spichtig, D.; et al. NFκB-sensitive Orai1 expression in the regulation of FGF23 release. J. Mol. Med. 2016, 94, 557–566. [Google Scholar] [CrossRef]
  71. Feger, M.; Hase, P.; Zhang, B.; Hirche, F.; Glosse, P.; Lang, F.; Föller, M. The production of fibroblast growth factor 23 is controlled by TGF-β2. Sci. Rep. 2017, 7, 4982. [Google Scholar] [CrossRef] [Green Version]
  72. Chen, Y.; Ramachandran, A.; Zhang, Y.; Koshy, R.; George, A. The ER Ca2+ sensor STIM1 can activate osteoblast and odontoblast differentiation in mineralized tissues. Connect. Tissue Res. 2018, 59, 6–12. [Google Scholar] [CrossRef] [Green Version]
  73. Roy, B.; Das, T.; Mishra, D.; Maiti, T.K.; Chakraborty, S. Oscillatory shear stress induced calcium flickers in osteoblast cells. Integr. Biol. 2014, 6, 289–299. [Google Scholar] [CrossRef]
  74. Xiao, E.; Yang, H.Q.; Gan, Y.H.; Duan, D.H.; He, L.H.; Guo, Y.; Wang, S.Q.; Zhang, Y. Brief reports: TRPM7 Senses mechanical stimulation inducing osteogenesis in human bone marrow mesenchymal stem cells. Stem Cells 2015, 33, 615–621. [Google Scholar] [CrossRef]
  75. Liu, Y.S.; Liu, Y.A.; Huang, C.J.; Yen, M.H.; Tseng, C.T.; Chien, S.; Lee, O.K. Mechanosensitive TRPM7 mediates shear stress and modulates osteogenic differentiation of mesenchymal stromal cells through Osterix pathway. Sci. Rep. 2015, 5, 16522. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  76. Rohde, M.; Ziebart, J.; Kirschstein, T.; Sellmann, T.; Porath, K.; Kühl, F.; Delenda, B.; Bahls, C.; van Rienen, U.; Bader, R.; et al. Human Osteoblast Migration in DC Electrical Fields Depends on Store Operated Ca2+-Release and Is Correlated to Upregulation of Stretch-Activated TRPM7 Channels. Front. Bioeng. Biotechnol. 2019, 7, 422. [Google Scholar] [CrossRef] [Green Version]
  77. Son, A.; Kang, N.; Kang, J.Y.; Kim, K.W.; Yang, Y.M.; Shin, D.M. TRPM3/TRPV4 regulates Ca2+-mediated RANKL/NFATc1 expression in osteoblasts. J. Mol. Endocrinol. 2018, 61, 207–218. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  78. He, L.H.; Liu, M.; He, Y.; Xiao, E.; Zhao, L.; Zhang, T.; Yang, H.Q.; Zhang, Y. TRPV1 deletion impaired fracture healing and inhibited osteoclast and osteoblast differentiation. Sci. Rep. 2017, 7, 42385. [Google Scholar] [CrossRef] [PubMed]
  79. Nishimura, H.; Kawasaki, M.; Tsukamoto, M.; Menuki, K.; Suzuki, H.; Matsuura, T.; Baba, K.; Motojima, Y.; Fujitani, T.; Ohnishi, H.; et al. Transient receptor potential vanilloid 1 and 4 double knockout leads to increased bone mass in mice. Bone Rep. 2020, 12, 100268. [Google Scholar] [CrossRef]
  80. Han, X.; Kato, H.; Sato, H.; Hirofuji, Y.; Fukumoto, S.; Masuda, K. Accelerated osteoblastic differentiation in patient-derived dental pulp stem cells carrying a gain-of-function mutation of TRPV4 associated with metatropic dysplasia. Biochem. Biophys. Res. Commun. 2020, 523, 841–846. [Google Scholar] [CrossRef] [PubMed]
  81. Pozo, A.; Regnier, M.; Lizotte, J.; Martineau, C.; Scorza, T.; Moreau, R. Cyclic adenosine monophosphate-dependent activation of transient receptor potential vanilloid 4 (TRPV4) channels in osteoblast-like MG-63 cells. Cell. Signal. 2020, 66, 109486. [Google Scholar] [CrossRef] [PubMed]
  82. Xiao, Z.; Baudry, J.; Cao, L.; Huang, J.; Chen, H.; Yates, C.R.; Li, W.; Dong, B.; Waters, C.M.; Smith, J.C.; et al. Polycystin-1 interacts with TAZ to stimulate osteoblastogenesis and inhibit adipogenesis. J. Clin. Investig. 2018, 128, 157–174. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  83. Merrick, D.; Mistry, K.; Wu, J.; Gresko, N.; Baggs, J.E.; Hogenesch, J.B.; Sun, Z.; Caplan, M.J. Polycystin-1 regulates bone development through an interaction with the transcriptional coactivator TAZ. Hum. Mol. Genet. 2019, 28, 16–30. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  84. Qiu, N.; Zhou, H.; Xiao, Z. Downregulation of PKD1 by shRNA results in defective osteogenic differentiation via cAMP/PKA pathway in human MG-63 cells. J. Cell. Biochem. 2012, 113, 967–976. [Google Scholar] [CrossRef] [Green Version]
  85. Mesner, L.D.; Ray, B.; Hsu, Y.H.; Manichaikul, A.; Lum, E.; Bryda, E.C.; Rich, S.S.; Rosen, C.J.; Criqui, M.H.; Allison, M.; et al. Bicc1 is a genetic determinant of osteoblastogenesis and bone mineral density. J. Clin. Investig. 2014, 124, 2736–2749. [Google Scholar] [CrossRef] [Green Version]
  86. Xiao, Z.; Cao, L.; Liang, Y.; Huang, J.; Stern, A.R.; Dallas, M.; Johnson, M.; Quarles, L.D. Osteoblast-specific deletion of Pkd2 leads to low-turnover osteopenia and reduced bone marrow adiposity. PLoS ONE 2014, 9, e114198. [Google Scholar] [CrossRef]
  87. Sugimoto, A.; Miyazaki, A.; Kawarabayashi, K.; Shono, M.; Akazawa, Y.; Hasegawa, T.; Ueda-Yamaguchi, K.; Kitamura, T.; Yoshizaki, K.; Fukumoto, S.; et al. Piezo type mechanosensitive ion channel component 1 functions as a regulator of the cell fate determination of mesenchymal stem cells. Sci. Rep. 2017, 7, 17696. [Google Scholar] [CrossRef] [Green Version]
  88. Yoneda, M.; Suzuki, H.; Hatano, N.; Nakano, S.; Muraki, Y.; Miyazawa, K.; Goto, S.; Muraki, K. PIEZO1 and TRPV4, which Are Distinct Mechano-Sensors in the Osteoblastic MC3T3-E1 Cells, Modify Cell-Proliferation. Int. J. Mol. Sci. 2019, 20, 4960. [Google Scholar] [CrossRef] [Green Version]
  89. Song, J.; Liu, L.; Lv, L.; Hu, S.; Tariq, A.; Wang, W.; Dang, X. Fluid shear stress induces Runx-2 expression via upregulation of PIEZO1 in MC3T3-E1 cells. Cell Biol. Int. 2020, 44, 1491–1502. [Google Scholar] [CrossRef] [PubMed]
  90. Sun, W.; Chi, S.; Li, Y.; Ling, S.; Tan, Y.; Xu, Y.; Jiang, F.; Li, J.; Liu, C.; Zhong, G.; et al. The mechanosensitive Piezo1 channel is required for bone formation. eLife 2019, 8, e47454. [Google Scholar] [CrossRef] [PubMed]
  91. Li, X.; Han, L.; Nookaew, I.; Mannen, E.; Silva, M.J.; Almeida, M.; Xiong, J. Stimulation of Piezo1 by mechanical signals promotes bone anabolism. eLife 2019, 8, e49631. [Google Scholar] [CrossRef]
  92. Sasaki, F.; Hayashi, M.; Mouri, Y.; Nakamura, S.; Adachi, T.; Nakashima, T. Mechanotransduction via the Piezo1-Akt pathway underlies Sost suppression in osteocytes. Biochem. Biophys. Res. Commun. 2020, 521, 806–813. [Google Scholar] [CrossRef] [PubMed]
  93. Hendrickx, G.; Fischer, V.; Liedert, A.; von Kroge, S.; Haffner-Luntzer, M.; Brylka, L.; Pawlus, E.; Schweizer, M.; Yorgan, T.; Baranowsky, A.; et al. Piezo1 Inactivation in Chondrocytes Impairs Trabecular Bone Formation. J. Bone Miner. Res. 2021, 36, 369–384. [Google Scholar] [CrossRef] [PubMed]
  94. Zhou, T.; Gao, B.; Fan, Y.; Liu, Y.; Feng, S.; Cong, Q.; Zhang, X.; Zhou, Y.; Yadav, P.S.; Lin, J.; et al. Piezo1/2 mediate mechanotransduction essential for bone formation through concerted activation of NFAT-YAP1-ss-catenin. eLife 2020, 9, e52779. [Google Scholar] [CrossRef] [PubMed]
  95. Prakriya, M.; Feske, S.; Gwack, Y.; Srikanth, S.; Rao, A.; Hogan, P.G. Orai1 is an essential pore subunit of the CRAC channel. Nature 2006, 443, 230–233. [Google Scholar] [CrossRef]
  96. Zuo, C.; Huang, Y.; Bajis, R.; Sahih, M.; Li, Y.P.; Dai, K.; Zhang, X. Osteoblastogenesis regulation signals in bone remodeling. Osteoporos. Int. 2012, 23, 1653–1663. [Google Scholar] [CrossRef]
  97. Salazar, V.S.; Gamer, L.W.; Rosen, V. BMP signalling in skeletal development, disease and repair. Nat. Rev. Endocrinol. 2016, 12, 203–221. [Google Scholar] [CrossRef]
  98. Kuro, O.M.; Moe, O.W. FGF23-alphaKlotho as a paradigm for a kidney-bone network. Bone 2017, 100, 4–18. [Google Scholar] [CrossRef]
  99. Raya, A.I.; Rios, R.; Pineda, C.; Rodriguez-Ortiz, M.E.; Diez, E.; Almaden, Y.; Munoz-Castaneda, J.R.; Rodriguez, M.; Aguilera-Tejero, E.; Lopez, I. Energy-dense diets increase FGF23, lead to phosphorus retention and promote vascular calcifications in rats. Sci. Rep. 2016, 6, 36881. [Google Scholar] [CrossRef] [Green Version]
  100. Smith, E.R.; Holt, S.G.; Hewitson, T.D. alphaKlotho-FGF23 interactions and their role in kidney disease: A molecular insight. Cell. Mol. Life Sci. 2019, 76, 4705–4724. [Google Scholar] [CrossRef]
  101. Glosse, P.; Feger, M.; Mutig, K.; Chen, H.; Hirche, F.; Hasan, A.A.; Gaballa, M.M.S.; Hocher, B.; Lang, F.; Föller, M. AMP-activated kinase is a regulator of fibroblast growth factor 23 production. Kidney Int. 2018, 94, 491–501. [Google Scholar] [CrossRef]
  102. Li, S.; Yang, B.; Du, Y.; Lin, Y.; Liu, J.; Huang, S.; Zhang, A.; Jia, Z.; Zhang, Y. Targeting PPARα for the Treatment and Understanding of Cardiovascular Diseases. Cell Physiol. Biochem. 2018, 51, 2760–2775. [Google Scholar] [CrossRef]
  103. Syversen, U.; Stunes, A.K.; Gustafsson, B.I.; Obrant, K.J.; Nordsletten, L.; Berge, R.; Thommesen, L.; Reseland, J.E. Different skeletal effects of the peroxisome proliferator activated receptor (PPAR)α agonist fenofibrate and the PPARg agonist pioglitazone. BMC Endocr. Disord. 2009, 9, 10. [Google Scholar] [CrossRef] [Green Version]
  104. Ewendt, F.; Hirche, F.; Feger, M.; Föller, M. Peroxisome proliferator-activated receptor α (PPARα)-dependent regulation of fibroblast growth factor 23 (FGF23). Pflug. Arch. Eur. J. Physiol. 2020, 472, 503–511. [Google Scholar] [CrossRef] [PubMed]
  105. Lacruz, R.S.; Feske, S. Diseases caused by mutations in ORAI1 and STIM1. Ann. N. Y. Acad. Sci. 2015, 1356, 45–79. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  106. Silva-Rojas, R.; Laporte, J.; Bohm, J. STIM1/ORAI1 Loss-of-Function and Gain-of-Function Mutations Inversely Impact on SOCE and Calcium Homeostasis and Cause Multi-Systemic Mirror Diseases. Front. Physiol. 2020, 11, 604941. [Google Scholar] [CrossRef] [PubMed]
  107. Misceo, D.; Holmgren, A.; Louch, W.E.; Holme, P.A.; Mizobuchi, M.; Morales, R.J.; De Paula, A.M.; Stray-Pedersen, A.; Lyle, R.; Dalhus, B.; et al. A dominant STIM1 mutation causes Stormorken syndrome. Hum. Mutat. 2014, 35, 556–564. [Google Scholar] [CrossRef] [PubMed]
  108. Gamage, T.H.; Lengle, E.; Gunnes, G.; Pullisaar, H.; Holmgren, A.; Reseland, J.E.; Merckoll, E.; Corti, S.; Mizobuchi, M.; Morales, R.J.; et al. STIM1 R304W in mice causes subgingival hair growth and an increased fraction of trabecular bone. Cell Calcium 2020, 85, 102110. [Google Scholar] [CrossRef] [PubMed]
  109. Huang, Y.; Li, Q.; Feng, Z.; Zheng, L. STIM1 controls calcineurin/Akt/mTOR/NFATC2-mediated osteoclastogenesis induced by RANKL/M-CSF. Exp. Ther. Med. 2020, 20, 736–747. [Google Scholar] [CrossRef]
  110. Li, H. TRP Channel Classification. Adv. Exp. Med. Biol. 2017, 976, 1–8. [Google Scholar]
  111. Lieben, L.; Carmeliet, G. The Involvement of TRP Channels in Bone Homeostasis. Front. Endocrinol. 2012, 3, 99. [Google Scholar] [CrossRef] [Green Version]
  112. Wittkowske, C.; Reilly, G.C.; Lacroix, D.; Perrault, C.M. In Vitro Bone Cell Models: Impact of Fluid Shear Stress on Bone Formation. Front. Bioeng. Biotechnol. 2016, 4, 87. [Google Scholar] [CrossRef] [Green Version]
  113. Cadossi, R.; Massari, L.; Racine-Avila, J.; Aaron, R.K. Pulsed Electromagnetic Field Stimulation of Bone Healing and Joint Preservation: Cellular Mechanisms of Skeletal Response. J. Am. Acad. Orthop. Surg. Glob. Res. Rev. 2020, 4, e1900155. [Google Scholar] [CrossRef] [PubMed]
  114. Hiemer, B.; Ziebart, J.; Jonitz-Heincke, A.; Grunert, P.C.; Su, Y.; Hansmann, D.; Bader, R. Magnetically induced electrostimulation of human osteoblasts results in enhanced cell viability and osteogenic differentiation. Int. J. Mol. Med. 2016, 38, 57–64. [Google Scholar] [CrossRef] [PubMed]
  115. Wang, R.; Wu, H.; Yang, Y.; Song, M. Effects of electromagnetic fields on osteoporosis: A systematic literature review. Electromagn. Biol. Med. 2016, 35, 384–390. [Google Scholar] [CrossRef] [PubMed]
  116. Park, H.J.; Baek, K.; Baek, J.H.; Kim, H.R. The cooperation of CREB and NFAT is required for PTHrP-induced RANKL expression in mouse osteoblastic cells. J. Cell. Physiol. 2015, 230, 667–679. [Google Scholar] [CrossRef]
  117. Nilius, B.; Voets, T. The puzzle of TRPV4 channelopathies. EMBO Rep. 2013, 14, 152–163. [Google Scholar] [CrossRef]
  118. Nonaka, K.; Han, X.; Kato, H.; Sato, H.; Yamaza, H.; Hirofuji, Y.; Masuda, K. Novel gain-of-function mutation of TRPV4 associated with accelerated chondrogenic differentiation of dental pulp stem cells derived from a patient with metatropic dysplasia. Biochem. Biophys. Rep. 2019, 19, 100648. [Google Scholar] [CrossRef]
  119. Jilka, R.L. Molecular and cellular mechanisms of the anabolic effect of intermittent PTH. Bone 2007, 40, 1434–1446. [Google Scholar] [CrossRef] [Green Version]
  120. Jilka, R.L.; Weinstein, R.S.; Bellido, T.; Roberson, P.; Parfitt, A.M.; Manolagas, S.C. Increased bone formation by prevention of osteoblast apoptosis with parathyroid hormone. J. Clin. Investig. 1999, 104, 439–446. [Google Scholar] [CrossRef] [Green Version]
  121. Kim, S.W.; Pajevic, P.D.; Selig, M.; Barry, K.J.; Yang, J.Y.; Shin, C.S.; Baek, W.Y.; Kim, J.E.; Kronenberg, H.M. Intermittent parathyroid hormone administration converts quiescent lining cells to active osteoblasts. J. Bone Miner. Res. 2012, 27, 2075–2084. [Google Scholar] [CrossRef] [Green Version]
  122. Torres, V.E.; Harris, P.C.; Pirson, Y. Autosomal dominant polycystic kidney disease. Lancet 2007, 369, 1287–1301. [Google Scholar] [CrossRef]
  123. Wang, Z.; Ng, C.; Liu, X.; Wang, Y.; Li, B.; Kashyap, P.; Chaudhry, H.A.; Castro, A.; Kalontar, E.M.; Ilyayev, L.; et al. The ion channel function of polycystin-1 in the polycystin-1/polycystin-2 complex. EMBO Rep. 2019, 20, e48336. [Google Scholar] [CrossRef] [PubMed]
  124. Nigro, E.A.; Boletta, A. Role of the polycystins as mechanosensors of extracellular stiffness. Am. J. Physiol. Ren. Physiol. 2021, 320, F693–F705. [Google Scholar] [CrossRef] [PubMed]
  125. Dalagiorgou, G.; Piperi, C.; Adamopoulos, C.; Georgopoulou, U.; Gargalionis, A.N.; Spyropoulou, A.; Zoi, I.; Nokhbehsaim, M.; Damanaki, A.; Deschner, J.; et al. Mechanosensor polycystin-1 potentiates differentiation of human osteoblastic cells by upregulating Runx2 expression via induction of JAK2/STAT3 signaling axis. Cell. Mol. Life Sci. 2017, 74, 921–936. [Google Scholar] [CrossRef]
  126. Gudipaty, S.A.; Lindblom, J.; Loftus, P.D.; Redd, M.J.; Edes, K.; Davey, C.F.; Krishnegowda, V.; Rosenblatt, J. Mechanical stretch triggers rapid epithelial cell division through Piezo1. Nature 2017, 543, 118–121. [Google Scholar] [CrossRef] [Green Version]
  127. Feng, J.; Luo, J.; Yang, P.; Du, J.; Kim, B.S.; Hu, H. Piezo2 channel-Merkel cell signaling modulates the conversion of touch to itch. Science 2018, 360, 530–533. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  128. Li, J.; Hou, B.; Tumova, S.; Muraki, K.; Bruns, A.; Ludlow, M.J.; Sedo, A.; Hyman, A.J.; McKeown, L.; Young, R.S.; et al. Piezo1 integration of vascular architecture with physiological force. Nature 2014, 515, 279–282. [Google Scholar] [CrossRef] [PubMed]
  129. Gnanasambandam, R.; Bae, C.; Gottlieb, P.A.; Sachs, F. Ionic Selectivity and Permeation Properties of Human PIEZO1 Channels. PLoS ONE 2015, 10, e0125503. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  130. Marouli, E.; Graff, M.; Medina-Gomez, C.; Lo, K.S.; Wood, A.R.; Kjaer, T.R.; Fine, R.S.; Lu, Y.; Schurmann, C.; Highland, H.M.; et al. Rare and low-frequency coding variants alter human adult height. Nature 2017, 542, 186–190. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  131. Morris, J.A.; Kemp, J.P.; Youlten, S.E.; Laurent, L.; Logan, J.G.; Chai, R.C.; Vulpescu, N.A.; Forgetta, V.; Kleinman, A.; Mohanty, S.T.; et al. An atlas of genetic influences on osteoporosis in humans and mice. Nat. Genet. 2019, 51, 258–266. [Google Scholar] [CrossRef]
  132. Wadhwa, S.; Godwin, S.L.; Peterson, D.R.; Epstein, M.A.; Raisz, L.G.; Pilbeam, C.C. Fluid flow induction of cyclo-oxygenase 2 gene expression in osteoblasts is dependent on an extracellular signal-regulated kinase signaling pathway. J. Bone Miner. Res. 2002, 17, 266–274. [Google Scholar] [CrossRef]
  133. Zhao, N.; Nociti, F.H., Jr.; Duan, P.; Prideaux, M.; Zhao, H.; Foster, B.L.; Somerman, M.J.; Bonewald, L.F. Isolation and Functional Analysis of an Immortalized Murine Cementocyte Cell Line, IDG-CM6. J. Bone Miner. Res. 2016, 31, 430–442. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  134. Lee, W.; Leddy, H.A.; Chen, Y.; Lee, S.H.; Zelenski, N.A.; McNulty, A.L.; Wu, J.; Beicker, K.N.; Coles, J.; Zauscher, S.; et al. Synergy between Piezo1 and Piezo2 channels confers high-strain mechanosensitivity to articular cartilage. Proc. Natl. Acad. Sci. USA 2014, 111, E5114–E5122. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  135. Wang, L.; You, X.; Lotinun, S.; Zhang, L.; Wu, N.; Zou, W. Mechanical sensing protein PIEZO1 regulates bone homeostasis via osteoblast-osteoclast crosstalk. Nat. Commun. 2020, 11, 282. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  136. Alejandro, P.; Constantinescu, F. A Review of Osteoporosis in the Older Adult: An Update. Rheum. Dis. Clin. 2018, 44, 437–451. [Google Scholar] [CrossRef] [PubMed]
  137. Suzuki, Y.; Landowski, C.P.; Hediger, M.A. Mechanisms and regulation of epithelial Ca2+ absorption in health and disease. Annu. Rev. Physiol. 2008, 70, 257–271. [Google Scholar] [CrossRef]
Figure 1. The Wnt/Ca2+ signaling is initiated when Wnt ligands bind to the Frizzled receptor. Then, the Disheveled (DVL) is recruited and activates phospholipase C (PLC), leading to Ca2+ release. Calreticulin (CRT) is a Ca2+-binding multifunctional molecular chaperone in the endoplasmic reticulum (ER). CRT is involved in IP3-mediated Ca2+ release from ER. [Ca2+]i increase via IP3R, transmembrane, and coiled-coil domains 1 (TMCO-1), and Ca2+-permeable channels cause calmodulin (CaM) activation. The Ca2+/CaM complex regulates Ca2+/CaM-dependent protein kinase and phosphatase such as CaMKII and calcineurin (Cn). These Ca2+-dependent signalings induce the activation of transcriptional factors, including the nuclear factor of activated T cells (NFAT), nuclear factor-kappa B (NF-κB), cAMP response element-binding protein (CREB), Runt-related transcription factor 2 (Runx2), and activation protein 1 (AP-1), leading to the osteoblast differentiation.
Figure 1. The Wnt/Ca2+ signaling is initiated when Wnt ligands bind to the Frizzled receptor. Then, the Disheveled (DVL) is recruited and activates phospholipase C (PLC), leading to Ca2+ release. Calreticulin (CRT) is a Ca2+-binding multifunctional molecular chaperone in the endoplasmic reticulum (ER). CRT is involved in IP3-mediated Ca2+ release from ER. [Ca2+]i increase via IP3R, transmembrane, and coiled-coil domains 1 (TMCO-1), and Ca2+-permeable channels cause calmodulin (CaM) activation. The Ca2+/CaM complex regulates Ca2+/CaM-dependent protein kinase and phosphatase such as CaMKII and calcineurin (Cn). These Ca2+-dependent signalings induce the activation of transcriptional factors, including the nuclear factor of activated T cells (NFAT), nuclear factor-kappa B (NF-κB), cAMP response element-binding protein (CREB), Runt-related transcription factor 2 (Runx2), and activation protein 1 (AP-1), leading to the osteoblast differentiation.
Ijms 22 10459 g001
Figure 2. In non-excitable cells, voltage-independent Ca2+ channels are the main Ca2+-permeable channels. Membrane hyperpolarization followed by K+ channel activation increases the driving force for Ca2+ influx. Consequently, K+ channels indirectly regulate osteoblast functions, including proliferation, differentiation, and mineralization, by controlling intracellular Ca2+ signaling in osteoblast lineage cells.
Figure 2. In non-excitable cells, voltage-independent Ca2+ channels are the main Ca2+-permeable channels. Membrane hyperpolarization followed by K+ channel activation increases the driving force for Ca2+ influx. Consequently, K+ channels indirectly regulate osteoblast functions, including proliferation, differentiation, and mineralization, by controlling intracellular Ca2+ signaling in osteoblast lineage cells.
Ijms 22 10459 g002
Figure 3. Phylogenetic tree topology of K+ channels, voltage-gated K+ (KV) channels (A), Ca2+-activated K+ channels (B), inward-rectifier K+ (Kir) channels (C), and two-pore K+ (K2P) channels (D).
Figure 3. Phylogenetic tree topology of K+ channels, voltage-gated K+ (KV) channels (A), Ca2+-activated K+ channels (B), inward-rectifier K+ (Kir) channels (C), and two-pore K+ (K2P) channels (D).
Ijms 22 10459 g003
Figure 4. Mechanical stress stimulates Piezo1 to promote Ca2+ influx in osteoblast lineage cells. The [Ca2+]i rises regulate various signal transduction pathways, resulting in osteoblastogenesis.
Figure 4. Mechanical stress stimulates Piezo1 to promote Ca2+ influx in osteoblast lineage cells. The [Ca2+]i rises regulate various signal transduction pathways, resulting in osteoblastogenesis.
Ijms 22 10459 g004
Table 1. Functional expression of K+ channels in osteoblast lineage cells.
Table 1. Functional expression of K+ channels in osteoblast lineage cells.
FamilyMemberCell TypesFunctionReference
voltage-gated K+ channelKv7.3human mesenchymal stem cells, MG-63, and Saos-2 cellsosteoblast differentiation and mineralization[27]
Kv10.1MG-63 cellsproliferation/ cell cycle progression[28]
MG-63, Saos-2, and hFOB 1.19 cellsprolifeation[29]
Kv2.1MG-63 cellsfunctional expression[30]
inward rectifier K+ channelKir2.1myoblasts form Andersen-Tawil syndrome (ATS) patientsMineralization [31]
induced pluripotent stem cells form ATS patientsosteoblast differentiation and mineralization[32]
Kir6.2human mesenchamal stromal cellsosteogenic diferentiation[33]
Kir6/SUR1rat calvarial osteoblastsproliferation, apoptosis, and mineralization[34]
Ca2+-activated K+ channelKCa1.1SaM-1, MG-63, SaOS-2, and HOS cellsfunctional expression[35]
ROS17/2.8 and MC3T3-E1 cellsosteoblast proliferation and differentiation[36]
mouse bone marrow mesenchymal stem cells, ROS17/2.8, and MC3T3-E1 cellsbone loss and defects in osteoblast formation[37]
rat/mouse vascular smooth muscle cellsvascular calcification (-)[38]
KCa3.1rat/mouse vascular smooth muscle cellsvascular calcification (+)[39]
MC3T3-E1 cellsprolifeataion[40]
two-pore K+ channelK2P2.1human osteoblasts and MG-63 cellsfunctional expression[41]
K2P3.1, K2P5.1, K2P9.1MG-63 cellsprolifeartion[42]
Table 2. Nomenclature of the Ca2+-permeable channel superfamilies.
Table 2. Nomenclature of the Ca2+-permeable channel superfamilies.
IUPHARHUGOSynonyms
TRPC
TRPC1
TRPC2
TRPC3
TRPC4
TRPC5
TRPC6
TRPC7
TRPC1
pseudogene
TRPC3
TRPC4
TRPC5
TRPC6
TRPC7
TRP1

TRP3
TRP4
TRP5
TRP6
TRP7
TRPV
TRPV1
TRPV2
TRPV3
TRPV4
TRPV5
TRPV6
TRPV1
TRPV2
TRPV3
TRPV4
TRPV5
TRPV6
VR1
VRL/VRL1
VR3
VR2/TRP12/OTRPC4
ECAC1/CAT2/OTRPC3
ECAC2/CAT1/CATL
TRPM
TRPM1
TRPM2
TRPM3
TRPM4
TRPM5
TRPM6
TRPM7
TRPM8
TRPM1
TRPM2
TRPM3
TRPM4
TRPM5
TRPM6
TRPM7
TRPM8
LTRPC1/MLSN1
LTRPC2/TRPC7
LTRPC3/ MLSN3
LTRPC4
LTRPC5/MTR1
CHAK2/HOMG1
LTRPC7/CHAK1
LTRPC6/TRPP8
TRPA
TRPA1TRPA1ANK/TM1
TRPN
TRPN1trpn1 (fish)nompC
TRPML
TRPML1
TRPML2
TRPML3
MCOLN1
MCOLN2
MCOLN3
mucolipin1
mucolipin2
mucolipin3
TRPP
TRPP2
TRPP3
TRPP5
PKD2
PKD2L1
PKD2L2
polycystin 2
polycystin 2L1
polycystin 2L2
Orai
Orai1
Orai2
Orai3
ORAI1
ORAI2
ORAI3
CRACM1/TMEM142A
TMEM142B
TMEM142C
Piezo
Piezo1
Piezo2
PIEZO1
PIEZO2
Table 3. Functional expression of Orai/STIM in osteoblast lineage cells.
Table 3. Functional expression of Orai/STIM in osteoblast lineage cells.
MemberCell TypesFunctionReference
Orai1mouse bone marrow mesenchymal stromal cells and MC3T3-E1 cellsosteoblast differentiation and mineralization[65]
human osteoprogenitor cells (CC-2538)osteoblast differentiation and mineralization[66]
mouse calvaria osteoblasts and mesenchymal progenitorsosteoblast differentiation and mineralization[67]
human cartilage derived mesenchamal stem cellsosteoblast differentiation[68]
mouse bone marrow mesenchymal stromal cellsosteoblast differentiation and mineralization[69]
UMR106 cellsFGF23 expression[70,71]
STIM1MC3T3-E1 cellsosteoblast differentiation and mineralization[72]
Table 4. Functional expression of TRP channels in osteoblast lineage cells.
Table 4. Functional expression of TRP channels in osteoblast lineage cells.
MemberCell TypesFunctionReference
TRPM7MG-63 cells
human mesenchymal stem cells
human osteoblasts
human osteoblasts
Ca2+ flicker activity
osteoblast differentiation
osteoblast differentiation
migration
[73]
[74,75]
[75]
[76]
TRPM3, TRPV4mouse calvarial osteoblastsRANKL and NFATc1 expression[77]
TRPV1mouse bone marrow stromal cellsosteoblast differentiation and mineralization[78]
TRPV1, TRPV4mouse bone marrow cellsosteoblast differentiation[79]
TRPV4mesenchymal stem cells from metatropic dysplasia patients
MG-63 cells
osteoblast differentiation
osteoblast proliferation and differentiation
[80]
[81]
TRPP1human osteoblastic cells
MG-63 cells
osteoblast differentiation
osteoblast differentiation and mineralization
[82,83]
[84]
TRPP1, TRPP2mouse calvarial osteoblastsosteoblast differentiation and mineralization[85]
TRPP2mouse calvarial osteoblastsosteoblast differentiation and mineralization[86]
Table 5. Functional expression of Piezo channels in osteoblast lineage cells.
Table 5. Functional expression of Piezo channels in osteoblast lineage cells.
MemberCell TypesFunctionReference
Piezo1human mesenchymal stem cells, UE7T-13 cells, and SDP11 cellosteoblast differentiation[87]
MC3T3-E1 cellsproliferation[88]
MC3T3-E1 cellsRunx2 expression[89]
MC3T3-E1 cells and mouse calvarial osteoblastsosteoblast differentiation[90]
mouse bone marrow stromal cellsosteoblast differentiation[91]
MLO-Y4 cellsmechanotransduction[91]
IDG-SW3 cellsSost expression[92]
mouse calvarial osteoblastsosteoblast differentiation[93]
Piezo1, Piezo2mouse bone marrow stromal cellsosteoblast differentiation [94]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Kito, H.; Ohya, S. Role of K+ and Ca2+-Permeable Channels in Osteoblast Functions. Int. J. Mol. Sci. 2021, 22, 10459. https://doi.org/10.3390/ijms221910459

AMA Style

Kito H, Ohya S. Role of K+ and Ca2+-Permeable Channels in Osteoblast Functions. International Journal of Molecular Sciences. 2021; 22(19):10459. https://doi.org/10.3390/ijms221910459

Chicago/Turabian Style

Kito, Hiroaki, and Susumu Ohya. 2021. "Role of K+ and Ca2+-Permeable Channels in Osteoblast Functions" International Journal of Molecular Sciences 22, no. 19: 10459. https://doi.org/10.3390/ijms221910459

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