Next Article in Journal
B Chromosomes in Free-Living Flatworms of the Genus Macrostomum (Platyhelminthes, Macrostomorpha)
Next Article in Special Issue
Muscle Research: A Tour d’Horizon
Previous Article in Journal
Investigating Pathogenicity and Virulence of Staphylococcus pettenkoferi: An Emerging Pathogen
Previous Article in Special Issue
Available In Vitro Models for Human Satellite Cells from Skeletal Muscle
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Ion Channels and Transporters in Muscle Cell Differentiation

1
Institute for Chemistry and Biochemistry, Freie Universität Berlin, 14195 Berlin, Germany
2
Zhongshan School of Medicine, Sun Yat-Sen University, Guangzhou 510080, China
3
Institute for Molecular Medicine, MSH Medical School Hamburg, 20457 Hamburg, Germany
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2021, 22(24), 13615; https://doi.org/10.3390/ijms222413615
Submission received: 15 November 2021 / Revised: 4 December 2021 / Accepted: 14 December 2021 / Published: 19 December 2021
(This article belongs to the Special Issue Muscular Structure, Physiology and Metabolism)

Abstract

:
Investigations on ion channels in muscle tissues have mainly focused on physiological muscle function and related disorders, but emerging evidence supports a critical role of ion channels and transporters in developmental processes, such as controlling the myogenic commitment of stem cells. In this review, we provide an overview of ion channels and transporters that influence skeletal muscle myoblast differentiation, cardiac differentiation from pluripotent stem cells, as well as vascular smooth muscle cell differentiation. We highlight examples of model organisms or patients with mutations in ion channels. Furthermore, a potential underlying molecular mechanism involving hyperpolarization of the resting membrane potential and a series of calcium signaling is discussed.

1. Introduction

Vertebrates possess three types of muscle tissue classified by morphology, function, and distribution: skeletal, cardiac, and smooth muscle. The movement of living organisms and the functioning of various visceral organs rely on muscle contraction and relaxation. These processes are controlled by endogenous bioelectric signaling mediated by ion channels and transporters. The loss or dysfunction of such transport proteins usually leads to serious diseases [1,2,3,4]. For example, mutations that disrupt the voltage-gated chloride channel ClC-1 [5,6,7] and the voltage-gated sodium channel Nav1.4 [8,9] result in myotonia congenita and paramyotonia congenita, respectively. Ion channels and their pathogenic roles have been extensively studied in mature, excitable muscle cells [10,11,12,13]. Moreover, there is increasing evidence that ion channels also play crucial roles in muscle development. In this review, we focus on these roles of ion channels and transporters. We summarize the molecular mechanisms by which ion channels or transporters regulate the differentiation of non-excitable stem or progenitor cells during myogenesis. Additionally, we emphasize the fundamental role of endogenous bioelectrical signals in developmental processes.
Similar to muscle cells, every living cell possesses a transmembrane potential (Vm) across the plasma membrane due to the uneven distribution of ions that is established or affected by various ion pumps, transporters, and channels. The range of resting membrane potentials varies between cell types. Stem cells and tumor cells tend to have a more positive, depolarized membrane potential, while terminally differentiated cells usually possess a much more negative, hyperpolarized resting potential [14,15,16,17]. For example, embryonic stem cells and skeletal muscle cells have resting membrane potentials of approximately −10 mV and −90 mV, respectively [14,16]. Ca2+ is one of the most important second messengers in vertebrate cells. Numerous physiological and pathophysiological processes are closely related to Ca2+ signaling. In general, the cytoplasmic free Ca2+ concentration is much lower than that of the extracellular environment. An elevated intracellular Ca2+ concentration can result from Ca2+ influx through specific Ca2+ channels in the plasma membrane (voltage-gated, ligand-gated, or store-operated calcium channels) or by the release of Ca2+ from intracellular calcium stores such as the endoplasmic reticulum, lysosomes, or mitochondria.

2. Ion Channels in Skeletal Myogenesis

2.1. Membrane Hyperpolarization

Skeletal muscle formation occurs during the entire lifespan of vertebrates, including embryonic development, postnatal growth, and damage repair in adults [18,19]. Myogenesis from myogenic precursor cells (myoblasts) generally begins with cell cycle withdrawal, followed by the expression of muscle-specific transcription factor myogenin and the subsequent fusion of multiple cells into multinucleated myotubes (Figure 1) [20,21]. Hyperpolarization of the membrane potential is a prerequisite for skeletal muscle myoblast differentiation. This hyperpolarization can result from an efflux of cations, most likely K+ according to the given intracellular and extracellular concentrations, or theoretically by an influx of anions. Primary muscle progenitor cells derived from single satellite cells maintain their stem cell identity rather than undergo myogenic commitment when hyperpolarization is impaired by high external K+ or the Na+,K+-ATPase inhibitor ouabain [22,23,24]. More specifically, upon induction of the myogenic differentiation of human myoblasts, the activation of an ether-à-go-go (EAG) K+ channel has been shown to rapidly hyperpolarize myoblasts from approximately −8 mV to approximately −32 mV [25,26,27]. This is followed by a further drop in the resting membrane potential to approximately −74 mV due to the activation of the inward-rectifying K+ channel Kir2.1 [27,28,29]. The human EAG K+ current density was reported to be low in proliferating myoblasts, to increase in fusion-competent myoblasts, and to decline again in myotubes [27]. By contrast, the Kir2.1 current has been found to be expressed in 40–50% of differentiating myoblasts and in all myotubes [22,27]. Notably, the activation of plasma membrane-localized Kir2.1 channels by dephosphorylation of Tyr242 is considered one of the earliest detectable events during myoblast differentiation [28]. It occurs within the first 6 h of differentiation, several hours before the expression of the two myogenic transcription factors myogenin and myocyte enhancer factor 2 (MEF2) [22,30].
In addition to these contributors of hyperpolarization, several types of ion channels, including ether-à-go-go-related gene (ERG) K+ channels [29], store-operated Ca2+ entry (SOCE) channels [32], and volume-regulated anion channels (VRACs) [33,34], have been shown to affect the resting membrane potential of fusion-competent myoblasts. Inhibition of the human ERG K+ channel activity depolarized myoblasts by approximately 10 mV [29], whereas knockdown of the stromal interaction molecule 1 (STIM1) or Orai1, reducing SOCE, impaired hyperpolarization and consequently inhibited myoblast differentiation [32]. Furthermore, it has been reported that, by activating the intermediate-conductance Ca2+-activated K+ channel (IKCa), extracellular 5’-guanosine-triphosphate (GTP) hyperpolarizes C2C12 cells from a mean value of −15 mV to approximately −75 mV and increases myosin heavy chain (MHC) expression [35,36,37]. VRAC is a plasma membrane channel formed by heteromers of leucine-rich repeat containing family 8 (LRRC8) members that mediates the flux of Cl and organic osmolytes in a variety of physiological processes [38,39,40,41,42]. Using an optical activity sensor [43], VRAC was shown to be transiently activated during the early stage of C2C12 (a mouse skeletal muscle myoblast cell line that expresses all five LRRC8 family members [44]) differentiation, which was also accompanied by a reduction in intracellular chloride [34]. While VRAC is not required for C2C12 proliferation [45], knockdown of the essential LRRC8A subunit [46,47] or pharmacological inhibition of its activity impaired the hyperpolarization and subsequent fusion of C2C12 myoblasts [33]. However, a VRAC-mediated efflux of Cl per se, which is evidenced by the increased cytosolic Cl upon VRAC inhibition, cannot contribute to the hyperpolarization. An explanation may be that VRAC affects other channels such as Kir2.1. VRAC was also proposed to be involved in myotube differentiation by regulating signaling independent of its ion transport activity [48].
The analysis of human patients and animal models provide in vivo evidence for the importance of ion channel function in the development of skeletal muscle. Patients with mutations in the KCNJ2 gene, which encodes the Kir2.1 potassium channel, exhibit severe craniofacial and limb defects, such as cleft palate and brachydactyly (shortened digits) [49,50,51]. Kir2.1 knockout mice also display a cleft palate and patterning defects in their skeletal digits [52,53]. Mice with global or skeletal muscle-specific deletions of STIM1 [54,55,56] and Orai1 [57,58,59] exhibit a drastically reduced muscle mass and much smaller body size compared with their wild-type littermates. Furthermore, mice lacking the essential LRRC8A subunit of the VRAC exhibited severe growth retardation, high prenatal and postnatal lethality, and various organ abnormalities, including thin skeletal muscle bundles [60]. However, targeted deletion of LRRC8A in skeletal muscle resulted in significantly smaller myofibers without affecting total muscle mass [48].

2.2. Ca2+ Signaling

An increase in the free cytoplasmic Ca2+ concentration is required for the expression of myogenic transcription factors and the formation of normal-sized myotubes [23,61,62,63,64]. The hyperpolarization of human myoblasts induced by the sequential activation of EAG and Kir2.1 has been shown to trigger a small but sustained influx of Ca2+ through α1H T-type voltage-gated Ca2+ channels (VGCCs), sufficient to cause a significant increase in the resting intracellular Ca2+ concentration [29,62]. This cytosolic Ca2+ signal activates the calcineurin/NFAT pathway, thereby inducing the expression of myogenin and MEF2 (Figure 2) [23]. Another Ca2+-dependent pathway involving the Ca2+/calmodulin-dependent kinase (CaMK), is required for myogenin expression [65] but does not link to Kir2.1-induced hyperpolarization [23]. Interestingly, a 10 mV depolarization of the resting potential was observed to increase the T-type Ca2+ current and to raise the intracellular free Ca2+ concentration, thus triggering a ten-fold acceleration of human myoblast fusion [29]. However, the involvement of T-type VGCCs as a primary Ca2+ entry mechanism in myoblast differentiation seems to be species-dependent, as it was shown that L-type rather than T-type Ca2+ currents can regulate the expression of myogenin and MHC in murine C2C12 cells [66,67]. A link between L-type VGCCs and calcineurin activity has also been suggested [68].
Intracellular Ca2+ can also be elevated during myoblast differentiation when Ca2+ is released from the endoplasmic reticulum (ER) through inositol 1,4,5 tris-phosphate receptors (IP3Rs), followed by Ca2+ entry through SOCE channels [69,70,71,72,73]. Knockdown of IP3R1 in human myoblasts impaired both endogenous spontaneous Ca2+ oscillations and SOCE, which in turn reduced the activity of two key enzymes of muscle differentiation: calcineurin and CaMKII [74]. By contrast, the overexpression of IP3R1 not only rescued normal differentiation in IP3R1-silenced myoblasts but also increased the percentage of MEF2-positive nuclei after one day of differentiation [74]. In zebrafish, it was shown that, in addition to the IP3 receptor, the Ca2+-dependent ryanodine receptor (RyR) also contributes to the cytosolic Ca2+ signal during myogenesis upon lysosomal Ca2+ release by two-pore channel type 2 (TPC2) activation [75,76]. Upon ER Ca2+ store depletion, the Ca2+ sensor STIM1 triggers Ca2+ influx through SOCE-mediating channels located at the ER-plasma membrane junction, thereby efficiently restoring the ER Ca2+ content [77,78,79]. Here, two classes of channels are involved [80]: Orai channels [77,81] and transient receptor potential canonical channels (TRPCs) [82,83]. The important roles of STIM1 [32,54,55,56,84], Orai1 [32,57,58,59], TRPC1 [85,86,87,88], and TRPC4 [89,90] during myogenesis in mouse and human have been established. Silencing any of them reduced SOCE and myoblast differentiation, whereas the forced expression of STIM1 with Orai1, TRPC,1 or TRPC4 in human myoblasts increased SOCE, accelerated myoblast fusion, and produced hypertrophic myotubes [32,89]. Furthermore, the N-methyl-D-aspartate (NMDA) receptor, a subtype of ionotropic glutamate receptors, was also shown to mediate Ca2+ influx and to promote C2C12 myoblast fusion [91]. It is worth recalling that the graded Ca2+ signal involved in skeletal muscle formation depends on Ca2+ release from intracellular stores as well as Ca2+ influx from the extracellular space [29,69]. However, all of these Ca2+ signals are inhibited when the hyperpolarization process that increases the driving force for Ca2+ is blocked [23,33,36,63]. Notably, SOCE is involved in both hyperpolarization and subsequent Ca2+ signaling. Moreover, muscles from TRPC1 knockout mice display reduced fiber cross-sectional area and contain less myofibrillar proteins [86,92].

2.3. Further Molecular Mechanisms

The failure of myoblasts to exit the cell cycle leads to defective myotube formation [20,93]. It was reported that blocking the Ca2+- and voltage-dependent K+ channel KCa1.1 in human primary myoblasts increased the levels of cytosolic Ca2+ and activated NFκB, which resulted in enhanced cell proliferation and reduced fusion [94]. Interestingly, KCa1.1 expression in myotonic dystrophy type 1 (DM1) myoblasts was found to be significantly decreased [95], whereas introducing functional KCa1.1 α-subunits into DM1 myoblasts reduced their proliferation to normal levels and rescued the expressions of MEF2 and myogenin [94]. While constitutive overexpression of the protein called chloride intracellular channel 5 (CLIC5) partly shifted C2C12 cells from G2/M phase to G0/G1 phase, resulting in decreased cell proliferation and increased expression levels of myogenin and MHC [96], a direct effect of altered ion transport remains to be shown because it is unlikely that CLICs function indeed as chloride channels [97]. The activation of Kv7 channels reduced proliferation and stimulated differentiation of C2C12 myoblasts [98]. In particular, it was reported that the endocannabinoid 2-arachidonoylglycerol inhibits skeletal muscle differentiation via cannabinoid type 1 receptor-mediated inhibition of Kv7.4 channels [99]. Knockdown of Kv7.4 reduced the expression levels of several differentiation markers, but overexpression of Kv7.4 did not enhance myoblast differentiation [100].
Inhibition of mechanosensitive (or stretch-activated) cation channels by pharmacological blockers leads to impaired phenotypic maturation of C2C12 myoblasts, including reduced expression of sarcomeric proteins and MHC and decreased creatine kinase activity [101,102], with contradicting findings on the inhibitory effect on myogenin expression. Several further ion transport proteins have been implicated in skeletal myogenesis, including TRPC3 [103], Pannexin1 and Pannexin3 [104], connexin43 [88,105,106], two-pore domain potassium channels TASK2 and TREK1 [107], nicotinic acetylcholine receptors [63,108], transient receptor potential vanilloid 1 (TRPV1) [109,110], and Na+/K+/2Cl cotransporter 1 (NKCC1) [111]. However, the specific mechanistic roles of these proteins in myogenic differentiation have not yet been elucidated.

3. Ion Channels and Transporters in Cardiac Differentiation

The heart is the first inner organ to form and function in the embryo. After birth, unlike skeletal muscle, the division or generation of cardiac muscle cells only occurs as a very slow process [112,113,114]. Due to the controversy or limitations surrounding cardiac progenitor cells [112,115,116], here, we focus on data describing cardiac differentiation of pluripotent stem cells (PSCs), such as embryonic stem cells (ESCs) and induced pluripotent stem cells (iPSCs). The differentiation of PSCs into cardiomyocytes is usually accomplished by the embryoid body method and is characterized by sequential expression of a series of genes: initial mesoderm and cardiomesoderm markers (brachyury T and mesoderm posterior protein 1 (MESP1)), followed by cardiac-specific transcription factors (NKX2.5, GATA4, and MEF2C), and finally cardiac-specific structural proteins (cardiac Troponin I (cTnI), MYH6, MYH7, and myosin light chain 2a (MLC2A)) (Figure 3) [112,115,117]. Interestingly, the activation of Ca2+-activated K+ channels (SKCas) (mainly the intermediate-conductance SKCa SK4) induced hyperpolarization of the membrane potential in undifferentiated murine ESCs, thereby inducing cardiac differentiation [118]. This suggests that membrane hyperpolarization precedes the signaling cascades of mesoderm commitment and cardiomyocyte specification. Notably, it was proposed that SKCa activation induces efficient cardiac differentiation of ESCs by activating the Ras-Mek1/2-ERK1/2 signal transduction pathway [118]. A recent study showed that the K+ channel ERG1 is involved in the cardiac differentiation of rat ESCs by interacting with integrin β1 and thus activating the AKT pathway (Figure 4) [119].
The sources of intracellular Ca2+ signals vary with the differentiation stage of PSC-derived cardiomyocytes and embryos maintained ex vivo [121,122,123]. In particular, the Na+/Ca2+ exchanger NCX1 is expressed during the earliest stages of heart development [123,124,125] and pharmacological blockade of NCX1 impacted on CaMKII signaling to downregulate the expression of key cardiac markers (Nkx2.5, Myh6, and Tnnt2 (encoding cTnT)), which led to impaired differentiation and failure of cardiac crescent formation [123]. The inhibition of L-type Ca2+ channels that are also expressed in the early stages of cardiac development resulted in the downregulation of Myh6 and Tnnt2 and a reduction in beating embryoid bodies [123]. In addition, the activity of the Na+/H+ exchanger NHE1 was shown to promote cardiomyocyte differentiation; however, the mechanism has so far remained unknown [126]. In contrast with skelatal muscle differentiation, there is no clear data on a contribution of intracellular calcium stores, such as the ER, lysosomes, or mitochondria, in cardiac muscle differentiation.
A mutation that disrupts the ERG K+ channel activity causes severe cardiac phenotypes in human patients, including QT prolongation, functional AV conduction disturbances, and polymorphic ventricular arrhythmias [127]. Similarly, another missense mutation in the human ERG, when introduced into the orthologous mouse gene in mouse ESCs, causes developmental cardiac defects in the right ventricle and its outflow tract. Homozygous mutant offspring died in utero by embryonic day 11.5 [128]. Mutations that result in reduced voltage-dependent channel inactivation of a specific L-type Ca2+ channel, Cav1.2, cause defects in heart development in human patients. These defects include lethal arrhythmias and congenital heart disease [129,130].

4. Ion Channel Activity in Smooth Muscle Cell Differentiation

Compared with skeletal muscle, there is only scarce information on the role of ion channels or transporters in vascular smooth muscle cell (VSMC) differentiation. VSMCs arise from multiple origins during embryonic development [131,132]. Although various in vitro models have been established to investigate the detailed mechanisms of deriving VSMCs from stem cells, it appears that origin-specific VSMCs possess individual regulatory mechanisms regarding the control of differentiation [133,134]. On the other hand, unlike cardiac and skeletal muscle cells, VSMCs do not terminally differentiate but maintain the ability to undergo phenotypic modulation in response to physiological and pathological stimuli, switching between a fully differentiated and contractile phenotype and a highly proliferative, migratory, and synthetic phenotype [135]. The limited progress made in defining VSMC differentiation [136,137,138] makes it more difficult to determine the role of ion channels or transporters in this process.
Many studies have presented confusing or apparently contradicting data on Ca2+ signaling, mostly by influx from the extracellular space [139,140,141,142,143]. Little is known about a potential role of intracellular Ca2+ stores in VSMC differentiation. While the lysosomal channels TRPML1 and TPC2 were shown to contribute to Ca2+ signaling in differentiated smooth muscle cells [144,145], a role for lysosomal as well as mitochondrial Ca2+ in their differentiation is unknown. The relative contributions of plasma membrane and ER Ca2+ channels in the plasma membrane alters when smooth muscle cells switch between a contractile and a proliferative phenotype [146,147]. However, this change in Ca2+ signaling does not seem to affect cell differentiation [148].
Nevertheless, it is beyond doubt that ion channels are critically involved in the development of vascular smooth muscle as well. One example is the ERG1 K+ channel, in which deletion results in defects in the yolk sac and intraembryonic vasculature. Treatment with the specific ERG1 antagonist dofetilide, both in vivo and in vitro, recapitulates this vascular phenotype [149]. Interestingly, a recent study demonstrated the importance of plasma membrane hyperpolarization in VSMC differentiation [150]. First, the induction of contractile differentiation of primary VSMCs by transforming growth factor TGF-β1 treatment caused hyperpolarization of the resting membrane potential. Second, TGF-β1-stimulated VSMC differentiation in the mesenchymal stem cell line C3H10T1/2 was inhibited in a dose-dependent manner in the presence of additional extracellular KCl. Furthermore, it was reported that TRPC6-mediated Ca2+ influx and depolarization suppressed VSMC differentiation by coordinately promoting the interaction of TRPC6 with lipid phosphatase and PTEN (phosphatase and tensin homolog) [150].

5. Conclusions

It is becoming increasingly clear that ion channels and transporters play conserved roles in developmental processes [151]. In contrast with transcriptional networks and signaling mechanisms, the emerging field of bioelectricity is a reservoir of new discoveries to be explored [15,16,17]. Changes in the resting membrane potential—as distinguished from transient membrane potential changes or oscillations during physiological processes—can be an instructive parameter in regulating cell fate decisions [152]. There are challenges to the investigation of the role for ion channels in cell differentiation, such as the fact that they can operate within a very narrow time window and that their regulation is not necessarily associated with changes in protein expression. The integration of various Ca2+ signals further adds to the complexity. Future research is required for identifying the intracellular or extracellular stimuli that control the activity of ion channels and transporters, and their downstream signaling. Elucidating the mechanisms by which ion channels and transporters promote muscle cell differentiation will lead to a better understanding of muscle development or disease and will provide insight for the development of therapeutic strategies relying on drugs or regenerative medicine.

Author Contributions

All authors contributed to the writing, and have read and agreed to the published version of the manuscript.

Funding

This research received no external funding. The publication of this article was funded by Freie Universität Berlin.

Conflicts of Interest

The authors declare no conflict of interest.

Abbreviations

CaMKCa2+/calmodulin-dependent kinase
CLIC5Chloride intracellular channel 5
cTnICardiac Troponin I
DM1Myotonic dystrophy type 1
EAGEther-à-go-go
EREndoplasmic reticulum
ESCEmbryonic stem cell
GTP5’-guanosine-triphosphate
IKCaIntermediate-conductance Ca2+-activated K+ channel
IP3RInositol 1,4,5 trisphosphate receptor
iPSCInduced pluripotent stem cell
LRRC8Leucine-rich repeat containing family 8
MCKMuscle creatine kinase
MEF2Myocyte enhancer factor 2
MESP1Mesoderm posterior protein 1
MHCMyosin heavy chain
MLC2AMyosin light chain 2a
Myf5 Myogenic factor 5
MyoD Myoblast determination protein
NFAT Nuclear factor of activated T-cell
NKCC1Na+/K+/2Cl cotransporter 1
NMDAN-methyl-D-aspartate
PSCPluripotent stem cell
PTEN Phosphatase and tensin homolog
RyRRyanodine receptor
SKCaSmall and intermediate conductance Ca2+-activated K+ channel
SOCEStore-operated Ca2+ entry
STIM1Stromal interaction molecule 1
TPC2Two-pore channel type 2
TRPCTransient receptor potential canonical channel
TRPV1Transient receptor potential vanilloid 1
VGCCVoltage-gated Ca2+ channel
VRACVolume-regulated anion channel
VSMCVascular smooth muscle cell

References

  1. Camerino, D.C.; Tricarico, D.; Desaphy, J.-F. Ion channel pharmacology. Neurotherapeutics 2007, 4, 184–198. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. Planells-Cases, R.; Jentsch, T.J. Chloride channelopathies. Biochim. Biophys. Acta Mol. Basis Dis. 2009, 1792, 173–189. [Google Scholar] [CrossRef] [Green Version]
  3. Enkvetchakul, D. Genetic Disorders of Ion Channels. Mo. Med. 2010, 107, 270–275. [Google Scholar] [PubMed]
  4. Liu, Y.; Wang, K. Exploiting the Diversity of Ion Channels: Modulation of Ion Channels for Therapeutic Indications. In Concepts and Principles of Pharmacology; Springer: Berlin/Heidelberg, Germany, 2019; Volume 260, pp. 187–205. [Google Scholar] [CrossRef]
  5. Koch, M.C.; Steinmeyer, K.; Lorenz, C.; Ricker, K.; Wolf, F.; Otto, M.; Zoll, B.; Lehmann-Horn, F.; Grzeschik, K.-H.; Jentsch, T.J. The Skeletal Muscle Chloride Channel in Dominant and Recessive Human Myotonia. Science 1992, 257, 797–800. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  6. Pedersen, T.H.; Riisager, A.; de Paoli, F.V.; Chen, T.-Y.; Nielsen, O.B. Role of physiological ClC-1 Cl ion channel regulation for the excitability and function of working skeletal muscle. J. Gen. Physiol. 2016, 147, 291–308. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  7. Jeng, C.-J.; Fu, S.-J.; You, C.-Y.; Peng, Y.-J.; Hsiao, C.-T.; Chen, T.-Y.; Tang, C.-Y. Defective Gating and Proteostasis of Human ClC-1 Chloride Channel: Molecular Pathophysiology of Myotonia Congenita. Front. Neurol. 2020, 11, 76. [Google Scholar] [CrossRef] [Green Version]
  8. Cannon, S.C. Sodium Channelopathies of Skeletal Muscle. Mediat. Drugs Gastrointest. Motil. I 2017, 246, 309–330. [Google Scholar] [CrossRef]
  9. Mantegazza, M.; Cestèle, S.; Catterall, W.A. Sodium channelopathies of skeletal muscle and brain. Physiol. Rev. 2021, 101, 1633–1689. [Google Scholar] [CrossRef]
  10. Cannon, S. Channelopathies of Skeletal Muscle Excitability. Muscle 2012, 5, 1013–1022. [Google Scholar] [CrossRef] [Green Version]
  11. Phillips, L.; Trivedi, J.R. Skeletal Muscle Channelopathies. Neurotherapeutics 2018, 15, 954–965. [Google Scholar] [CrossRef] [Green Version]
  12. Kline, J.; Costantini, O. Inherited Cardiac Arrhythmias and Channelopathies. Med Clin. North Am. 2019, 103, 809–820. [Google Scholar] [CrossRef]
  13. Wallace, E.; Howard, L.; Liu, M.; O’Brien, T.; Ward, D.; Shen, S.; Prendiville, T. Long QT Syndrome: Genetics and Future Perspective. Pediatr. Cardiol. 2019, 40, 1419–1430. [Google Scholar] [CrossRef] [Green Version]
  14. Binggeli, R.; Weinstein, R.C. Membrane potentials and sodium channels: Hypotheses for growth regulation and cancer formation based on changes in sodium channels and gap junctions. J. Theor. Biol. 1986, 123, 377–401. [Google Scholar] [CrossRef]
  15. Yang, M.; Brackenbury, W.J. Membrane potential and cancer progression. Front. Physiol. 2013, 4, 185. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  16. Levin, M.; Pezzulo, G.; Finkelstein, J.M. Endogenous Bioelectric Signaling Networks: Exploiting Voltage Gradients for Control of Growth and Form. Annu. Rev. Biomed. Eng. 2017, 19, 353–387. [Google Scholar] [CrossRef] [PubMed]
  17. Levin, M. Molecular bioelectricity: How endogenous voltage potentials control cell behavior and instruct pattern regulation in vivo. Mol. Biol. Cell 2014, 25, 3835–3850. [Google Scholar] [CrossRef] [PubMed]
  18. Bentzinger, C.F.; Wang, Y.X.; Rudnicki, M.A. Building Muscle: Molecular Regulation of Myogenesis. Cold Spring Harb. Perspect. Biol. 2012, 4, a008342. [Google Scholar] [CrossRef]
  19. Chal, J.; Pourquié, O. Making muscle: Skeletal myogenesis in vivo and in vitro. Development 2017, 144, 2104–2122. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  20. Walsh, K.; Perlman, H. Cell cycle exit upon myogenic differentiation. Curr. Opin. Genet. Dev. 1997, 7, 597–602. [Google Scholar] [CrossRef]
  21. Sampath, S.C.; Sampath, S.C.; Millay, D.P. Myoblast fusion confusion: The resolution begins. Skeletal Muscle 2018, 8, 3. [Google Scholar] [CrossRef] [Green Version]
  22. Konig, S.; Hinard, V.; Arnaudeau, S.; Holzer, N.; Potter, G.; Bader, C.R.; Bernheim, L. Membrane Hyperpolarization Triggers Myogenin and Myocyte Enhancer Factor-2 Expression during Human Myoblast Differentiation. J. Biol. Chem. 2004, 279, 28187–28196. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  23. Konig, S.; Béguet, A.; Bader, C.R.; Bernheim, L. The calcineurin pathway links hyperpolarization (Kir2.1)-induced Ca2+ signals to human myoblast differentiation and fusion. Development 2006, 133, 3107–3114. [Google Scholar] [CrossRef] [Green Version]
  24. Fennelly, C.; Wang, Z.; Criswell, T.; Soker, S. Sustained Depolarization of the Resting Membrane Potential Regulates Muscle Progenitor Cell Growth and Maintains Stem Cell Properties In Vitro. Stem Cell Rev. Rep. 2016, 12, 634–644. [Google Scholar] [CrossRef] [PubMed]
  25. Bernheim, L.; Liu, J.H.; Hamann, M.; Haenggeli, C.A.; Fischer-Lougheed, J.; Bader, C.R. Contribution of a non-inactivating potassium current to the resting membrane potential of fusion-competent human myoblasts. J. Physiol. 1996, 493, 129–141. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  26. Bijlenga, P.; Occhiodoro, T.; Liu, J.-H.; Bader, C.R.; Bernheim, L.; Fischer-Lougheed, J. An ether-à-go-go K+ current, Ih-eag, contributes to the hyperpolarization of human fusion-competent myoblasts. J. Physiol. 1998, 512, 317–323. [Google Scholar] [CrossRef]
  27. Liu, J.-H.; Bijlenga, P.; Fischer-Lougheed, J.; Occhiodoro, T.; Kaelin, A.; Bader, C.R.; Bernheim, L. Role of an inward rectifier K+ current and of hyperpolarization in human myoblast fusion. J. Physiol. 1998, 510, 467–476. [Google Scholar] [CrossRef]
  28. Fischer-Lougheed, J.; Liu, J.-H.; Espinos, E.; Mordasini, D.; Bader, C.R.; Belin, D.; Bernheim, L. Human Myoblast Fusion Requires Expression of Functional Inward Rectifier Kir2.1 Channels. J. Cell Biol. 2001, 153, 677–686. [Google Scholar] [CrossRef] [Green Version]
  29. Liu, J.-H.; König, S.; Michel, M.; Arnaudeau, S.; Fischer-Lougheed, J.; Bader, C.R.; Bernheim, L. Acceleration of human myoblast fusion by depolarization: Graded Ca2+ signals involved. Development 2003, 130, 3437–3446. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  30. Hinard, V.; Belin, D.; Konig, S.; Bader, C.R.; Bernheim, L. Initiation of human myoblast differentiation via dephosphorylation of Kir2.1 K+ channels at tyrosine 242. Development 2008, 135, 859–867. [Google Scholar] [CrossRef] [Green Version]
  31. Jang, Y.-N.; Baik, E.J. JAK-STAT pathway and myogenic differentiation. JAK-STAT 2013, 2, e23282. [Google Scholar] [CrossRef] [Green Version]
  32. Darbellay, B.; Arnaudeau, S.; König, S.; Jousset, H.; Bader, C.; Demaurex, N.; Bernheim, L. STIM1- and Orai1-dependent Store-operated Calcium Entry Regulates Human Myoblast Differentiation. J. Biol. Chem. 2009, 284, 5370–5380. [Google Scholar] [CrossRef] [Green Version]
  33. Chen, L.; Becker, T.M.; Koch, U.; Stauber, T. The LRRC8/VRAC anion channel facilitates myogenic differentiation of murine myoblasts by promoting membrane hyperpolarization. J. Biol. Chem. 2019, 294, 14279–14288. [Google Scholar] [CrossRef] [PubMed]
  34. Chen, L.; König, B.; Stauber, T. LRRC8 channel activation and reduction in cytosolic chloride concentration during early differentiation of C2C12 myoblasts. Biochem. Biophys. Res. Commun. 2020, 532, 482–488. [Google Scholar] [CrossRef] [PubMed]
  35. Fioretti, B.; Pietrangelo, T.; Catacuzzeno, L.; Franciolini, F. Intermediate-conductance Ca2+-activated K+ channel is expressed in C2C12 myoblasts and is downregulated during myogenesis. Am. J. Physiol. Physiol. 2005, 289, C89–C96. [Google Scholar] [CrossRef] [Green Version]
  36. Pietrangelo, T.; Fioretti, B.; Mancinelli, R.; Catacuzzeno, L.; Franciolini, F.; Fanò, G.; Fulle, S. Extracellular guanosine-5′-triphosphate modulates myogenesis via intermediate Ca2+-activated K+ currents in C2C12 mouse cells. J. Physiol. 2006, 572, 721–733. [Google Scholar] [CrossRef] [PubMed]
  37. Tanaka, S.; Ono, Y.; Sakamoto, K. DCEBIO facilitates myogenic differentiation via intermediate conductance Ca2+ activated K+ channel activation in C2C12 myoblasts. J. Pharmacol. Sci. 2017, 133, 276–279. [Google Scholar] [CrossRef]
  38. Chen, L.; König, B.; Liu, T.; Pervaiz, S.; Razzaque, Y.S.; Stauber, T. More than just a pressure relief valve: Physiological roles of volume-regulated LRRC8 anion channels. Biol. Chem. 2019, 400, 1481–1496. [Google Scholar] [CrossRef] [PubMed]
  39. Kolobkova, Y.; Pervaiz, S.; Stauber, T. The expanding toolbox to study the LRRC8-formed volume-regulated anion channel VRAC. Curr. Top. Membr. 2021, 88, 119–163. [Google Scholar] [CrossRef] [PubMed]
  40. König, B.; Stauber, T. Biophysics and Structure-Function Relationships of LRRC8-Formed Volume-Regulated Anion Channels. Biophys. J. 2019, 116, 1185–1193. [Google Scholar] [CrossRef] [Green Version]
  41. Osei-Owusu, J.; Yang, J.; Vitery, M.D.C.; Qiu, Z. Molecular Biology and Physiology of Volume-Regulated Anion Channel (VRAC). Curr. Top. Membr. 2018, 81, 177–203. [Google Scholar] [CrossRef]
  42. Strange, K.; Yamada, T.; Denton, J.S. A 30-year journey from volume-regulated anion currents to molecular structure of the LRRC8 channel. J. Gen. Physiol. 2019, 151, 100–117. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  43. König, B.; Hao, Y.; Schwartz, S.; Plested, A.J.; Stauber, T. A FRET sensor of C-terminal movement reveals VRAC activation by plasma membrane DAG signaling rather than ionic strength. eLife 2019, 8, e45421. [Google Scholar] [CrossRef]
  44. Pervaiz, S.; Kopp, A.; Von Kleist, L.; Stauber, T. Absolute Protein Amounts and Relative Abundance of Volume-regulated Anion Channel (VRAC) LRRC8 Subunits in Cells and Tissues Revealed by Quantitative Immunoblotting. Int. J. Mol. Sci. 2019, 20, 5879. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  45. Liu, T.; Stauber, T. The Volume-Regulated Anion Channel LRRC8/VRAC Is Dispensable for Cell Proliferation and Migration. Int. J. Mol. Sci. 2019, 20, 2663. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  46. Qiu, Z.; Dubin, A.E.; Mathur, J.; Tu, B.; Reddy, K.; Miraglia, L.J.; Reinhardt, J.; Orth, A.P.; Patapoutian, A. SWELL1, a Plasma Membrane Protein, Is an Essential Component of Volume-Regulated Anion Channel. Cell 2014, 157, 447–458. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  47. Voss, F.K.; Ullrich, F.; Münch, J.; Lazarow, K.; Lutter, D.; Mah, N.; Andrade-Navarro, M.A.; von Kries, J.P.; Stauber, T.; Jentsch, T.J. Identification of LRRC8 Heteromers as an Essential Component of the Volume-Regulated Anion Channel VRAC. Science 2014, 344, 634–638. [Google Scholar] [CrossRef] [Green Version]
  48. Kumar, A.; Xie, L.; Stainier, D.Y.; Hinton, A.O.; Gunasekar, S.K.; Minerath, R.A.; Shen, K.; Maurer, J.M.; Grueter, C.E.; Abel, E.D.; et al. SWELL1 regulates skeletal muscle cell size, intracellular signaling, adiposity and glucose metabolism. eLife 2020, 9. [Google Scholar] [CrossRef] [PubMed]
  49. Bendahhou, S.; Donaldson, M.R.; Plaster, N.M.; Tristani-Firouzi, M.; Fu, Y.-H.; Ptácek, L.J. Defective Potassium Channel Kir2.1 Trafficking Underlies Andersen-Tawil Syndrome. J. Biol. Chem. 2003, 278, 51779–51785. [Google Scholar] [CrossRef] [Green Version]
  50. Yoon, G.; Oberoi, S.; Tristani-Firouzi, M.; Etheridge, S.; Quitania, L.; Kramer, J.; Miller, B.; Fu, Y.; Ptáček, L. Andersen-Tawil syndrome: Prospective cohort analysis and expansion of the phenotype. Am. J. Med Genet. Part A 2006, 140A, 312–321. [Google Scholar] [CrossRef] [PubMed]
  51. Plaster, N.M.; Tawil, R.; Tristani-Firouzi, M.; Canún, S.; Bendahhou, S.; Tsunoda, A.; Donaldson, M.R.; Iannaccone, S.T.; Brunt, E.; Barohn, R.; et al. Mutations in Kir2.1 cause the developmental and episodic electrical phenotypes of Andersen’s syndrome. Cell 2001, 105, 511–519. [Google Scholar] [CrossRef] [Green Version]
  52. Zaritsky, J.J.; Eckman, D.M.; Wellman, G.C.; Nelson, M.; 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] [PubMed] [Green Version]
  53. Dahal, G.R.; Rawson, J.; Gassaway, B.; Kwok, B.; Tong, Y.; Ptáček, L.J.; Bates, E. An inwardly rectifying K+ channel is required for patterning. Development 2012, 139, 3653–3664. [Google Scholar] [CrossRef] [Green Version]
  54. Stiber, J.; Hawkins, A.; Zhang, Z.-S.; Wang, S.; Burch, J.; Graham, V.; Ward, C.C.; Seth, M.; Finch, E.; Malouf, N.; et al. STIM1 signalling controls store-operated calcium entry required for development and contractile function in skeletal muscle. Nat. Cell Biol. 2008, 10, 688–697. [Google Scholar] [CrossRef] [Green Version]
  55. Varga-Szabo, D.; Braun, A.; Kleinschnitz, C.; Bender, M.; Pleines, I.; Pham, M.; Renné, T.; Stoll, G.; Nieswandt, B. The calcium sensor STIM1 is an essential mediator of arterial thrombosis and ischemic brain infarction. J. Exp. Med. 2008, 205, 1583–1591. [Google Scholar] [CrossRef]
  56. Li, T.; Finch, E.A.; Graham, V.; Zhang, Z.-S.; Ding, J.-D.; Burch, J.; Oh-Hora, M.; Rosenberg, P. STIM1-Ca2+ Signaling Is Required for the Hypertrophic Growth of Skeletal Muscle in Mice. Mol. Cell. Biol. 2012, 32, 3009–3017. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  57. Gwack, Y.; Srikanth, S.; Oh-Hora, M.; Hogan, P.G.; Lamperti, E.D.; Yamashita, M.; Gelinas, C.; Neems, D.S.; Sasaki, Y.; Feske, S.; et al. Hair loss and defective T- and B-cell function in mice lacking ORAI1. Mol. Cell. Biol. 2008, 28, 5209–5222. [Google Scholar] [CrossRef] [Green Version]
  58. Vig, M.; DeHaven, W.I.; Bird, G.S.; Billingsley, J.M.; Wang, H.; Rao, P.E.; Hutchings, A.B.; Jouvin, M.-H.; Putney, J.; Kinet, J.-P. Defective mast cell effector functions in mice lacking the CRACM1 pore subunit of store-operated calcium release–activated calcium channels. Nat. Immunol. 2007, 9, 89–96. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  59. Wei-LaPierre, L.; Carrell, E.M.; Boncompagni, S.; Protasi, F.; Dirksen, R.T. Orai1-dependent calcium entry promotes skeletal muscle growth and limits fatigue. Nat. Commun. 2013, 4, 1–12. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  60. Kumar, L.; Chou, J.; Yee, C.S.; Borzutzky, A.; Vollmann, E.H.; Von Andrian, U.H.; Park, S.-Y.; Hollander, G.; Manis, J.P.; Poliani, P.L.; et al. Leucine-rich repeat containing 8A (LRRC8A) is essential for T lymphocyte development and function. J. Exp. Med. 2014, 211, 929–942. [Google Scholar] [CrossRef]
  61. Przybylski, R.J.; MacBride, R.G.; Kirby, A.C. Calcium regulation of skeletal myogenesis. I. Cell content critical to myotube formation. In Vitro Cell. Dev. Biol. 1989, 25, 830–838. [Google Scholar] [CrossRef]
  62. Bijlenga, P.; Liu, J.-H.; Espinos, E.; Haenggeli, C.-A.; Fischer-Lougheed, J.; Bader, C.R.; Bernheim, L. T-type alpha 1H Ca2+ channels are involved in Ca2+ signaling during terminal differentiation (fusion) of human myoblasts. Proc. Natl. Acad. Sci. USA 2000, 97, 7627–7632. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  63. Constantin, B.; Cognard, C.; Raymond, G. Myoblast fusion requires cytosolic calcium elevation but not activation of voltage-dependent calcium channels. Cell Calcium 1996, 19, 365–374. [Google Scholar] [CrossRef]
  64. Przybylski, R.J.; Szigeti, V.; Davidheiser, S.; Kirby, A.C. Calcium regulation of skeletal myogenesis. II. Extracellular and cell surface effects. Cell Calcium 1994, 15, 132–142. [Google Scholar] [CrossRef]
  65. Xu, Q.; Yu, L.; Liu, L.; Cheung, C.F.; Li, X.; Yee, S.-P.; Yang, X.-J.; Wu, Z. p38 Mitogen-activated Protein Kinase-, Calcium-Calmodulin–dependent Protein Kinase-, and Calcineurin-mediated Signaling Pathways Transcriptionally Regulate Myogenin Expression. Mol. Biol. Cell 2002, 13, 1940–1952. [Google Scholar] [CrossRef] [Green Version]
  66. Porter, G.; Makuck, R.F.; Rivkees, S.A. Reduction in Intracellular Calcium Levels Inhibits Myoblast Differentiation. J. Biol. Chem. 2002, 277, 28942–28947. [Google Scholar] [CrossRef] [Green Version]
  67. Bidaud, I.; Monteil, A.; Nargeot, J.; Lory, P. Properties and role of voltage-dependent calcium channels during mouse skeletal muscle differentiation. J. Muscle Res. Cell Motil. 2006, 27, 75–81. [Google Scholar] [CrossRef]
  68. Spangenburg, E.E.; Bowles, D.; Booth, F.W. Insulin-Like Growth Factor-Induced Transcriptional Activity of the Skeletal α-Actin Gene Is Regulated by Signaling Mechanisms Linked to Voltage-Gated Calcium Channels during Myoblast Differentiation. Endocrinology 2004, 145, 2054–2063. [Google Scholar] [CrossRef]
  69. Arnaudeau, S.; Holzer, N.; König, S.; Bader, C.R.; Bernheim, L. Calcium sources used by post-natal human myoblasts during initial differentiation. J. Cell. Physiol. 2006, 208, 435–445. [Google Scholar] [CrossRef]
  70. Nakanishi, K.; Kakiguchi, K.; Yonemura, S.; Nakano, A.; Morishima, N. Transient Ca2+ depletion from the endoplasmic reticulum is critical for skeletal myoblast differentiation. FASEB J. 2015, 29, 2137–2149. [Google Scholar] [CrossRef]
  71. Araya, R.; Riquelme, M.A.; Brandan, E.; Sáez, J.C. The formation of skeletal muscle myotubes requires functional membrane receptors activated by extracellular ATP. Brain Res. Rev. 2004, 47, 174–188. [Google Scholar] [CrossRef]
  72. Seigneurin-Venin, S.; Parrish, E.; Marty, I.; Rieger, F.; Romey, G.; Villaz, M.; García, L. Involvement of the Dihydropyridine Receptor and Internal Ca2+ Stores in Myoblast Fusion. Exp. Cell Res. 1996, 223, 301–307. [Google Scholar] [CrossRef]
  73. Stiber, J.A.; Tabatabaei, N.; Hawkins, A.F.; Hawke, T.; Worley, P.F.; Williams, R.S.; Rosenberg, P. Homer modulates NFAT-dependent signaling during muscle differentiation. Dev. Biol. 2005, 287, 213–224. [Google Scholar] [CrossRef] [PubMed]
  74. Antigny, F.; Konig, S.; Bernheim, L.; Frieden, M. Inositol 1,4,5 trisphosphate receptor 1 is a key player of human myoblast differentiation. Cell Calcium 2014, 56, 513–521. [Google Scholar] [CrossRef] [PubMed]
  75. Kelu, J.J.; Webb, S.E.; Parrington, J.; Galione, A.; Miller, A.L. Ca2+ release via two-pore channel type 2 (TPC2) is required for slow muscle cell myofibrillogenesis and myotomal patterning in intact zebrafish embryos. Dev. Biol. 2017, 425, 109–129. [Google Scholar] [CrossRef]
  76. Kelu, J.J.; Chan, H.L.; Webb, S.E.; Cheng, A.H.; Ruas, M.; Parrington, J.; Galione, A.; Miller, A.L. Two-Pore Channel 2 activity is required for slow muscle cell-generated Ca2+ signaling during myogenesis in intact zebrafish. Int. J. Dev. Biol. 2015, 59, 313–325. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  77. Luik, R.M.; Wu, M.M.; Buchanan, J.; Lewis, R.S. The elementary unit of store-operated Ca2+ entry: Local activation of CRAC channels by STIM1 at ER–plasma membrane junctions. J. Cell Biol. 2006, 174, 815–825. [Google Scholar] [CrossRef]
  78. Wu, M.M.; Buchanan, J.; Luik, R.M.; Lewis, R.S. Ca2+ store depletion causes STIM1 to accumulate in ER regions closely associated with the plasma membrane. J. Cell Biol. 2006, 174, 803–813. [Google Scholar] [CrossRef] [PubMed]
  79. Jousset, H.; Frieden, M.; Demaurex, N. STIM1 Knockdown Reveals That Store-operated Ca2+ Channels Located Close to Sarco/Endoplasmic Ca2+ ATPases (SERCA) Pumps Silently Refill the Endoplasmic Reticulum. J. Biol. Chem. 2007, 282, 11456–11464. [Google Scholar] [CrossRef] [Green Version]
  80. Choi, J.H.; Jeong, S.Y.; Oh, M.R.; Allen, P.D.; Lee, E.H. TRPCs: Influential Mediators in Skeletal Muscle. Cells 2020, 9, 850. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  81. Xu, P.; Lu, J.; Li, Z.; Yu, X.; Chen, L.; Xu, T. Aggregation of STIM1 underneath the plasma membrane induces clustering of Orai1. Biochem. Biophys. Res. Commun. 2006, 350, 969–976. [Google Scholar] [CrossRef]
  82. Lopez, J.J.; Salido, G.M.; Pariente, J.A.; Rosado, J. Interaction of STIM1 with Endogenously Expressed Human Canonical TRP1 upon Depletion of Intracellular Ca2+ Stores. J. Biol. Chem. 2006, 281, 28254–28264. [Google Scholar] [CrossRef] [Green Version]
  83. Yuan, J.P.; Zeng, W.; Huang, G.N.; Worley, P.F.; Muallem, S. STIM1 heteromultimerizes TRPC channels to determine their function as store-operated channels. Nat. Cell Biol. 2007, 9, 636–645. [Google Scholar] [CrossRef]
  84. Darbellay, B.; Arnaudeau, S.; Ceroni, D.; Bader, C.R.; Konig, S.; Bernheim, L. Human Muscle Economy Myoblast Differentiation and Excitation-Contraction Coupling Use the Same Molecular Partners, STIM1 and STIM2. J. Biol. Chem. 2010, 285, 22437–22447. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  85. Louis, M.; Zanou, N.; Van Schoor, M.; Gailly, P. TRPC1 regulates skeletal myoblast migration and differentiation. J. Cell Sci. 2008, 121, 3951–3959. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  86. Zanou, N.; Schakman, O.; Louis, P.; Ruegg, U.; Dietrich, A.; Birnbaumer, L.; Gailly, P. Trpc1 Ion Channel Modulates Phosphatidylinositol 3-Kinase/Akt Pathway during Myoblast Differentiation and Muscle Regeneration. J. Biol. Chem. 2012, 287, 14524–14534. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  87. Formigli, L.; Sassoli, C.; Squecco, R.; Bini, F.; Martinesi, M.; Chellini, F.; Luciani, G.; Sbrana, F.; Zecchi-Orlandini, S.; Francini, F.; et al. Regulation of transient receptor potential canonical channel 1 (TRPC1) by sphingosine 1-phosphate in C2C12 myoblasts and its relevance for a role of mechanotransduction in skeletal muscle differentiation. J. Cell Sci. 2009, 122, 1322–1333. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  88. Meacci, E.; Bini, F.; Sassoli, C.; Martinesi, M.; Squecco, R.; Chellini, F.; Zecchi-Orlandini, S.; Francini, F.; Formigli, L. Functional interaction between TRPC1 channel and connexin-43 protein: A novel pathway underlying S1P action on skeletal myogenesis. Cell. Mol. Life Sci. 2010, 67, 4269–4285. [Google Scholar] [CrossRef]
  89. Antigny, F.; Koenig, S.; Bernheim, L.; Frieden, M. During post-natal human myogenesis, normal myotube size requires TRPC1 and TRPC4 mediated Ca2+ entry. J. Cell Sci. 2013, 126, 2525–2533. [Google Scholar] [CrossRef] [Green Version]
  90. Antigny, F.; Sabourin, J.; Saüc, S.; Bernheim, L.; Koenig, S.; Frieden, M. TRPC1 and TRPC4 channels functionally interact with STIM1L to promote myogenesis and maintain fast repetitive Ca2+ release in human myotubes. Biochim. Biophys. Acta Bioenerg. 2017, 1864, 806–813. [Google Scholar] [CrossRef] [PubMed]
  91. Lee, K.H.; Park, J.-Y.; Kim, K. NMDA receptor-mediated calcium influx plays an essential role in myoblast fusion. FEBS Lett. 2004, 578, 47–52. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  92. Zanou, N.; Shapovalov, G.; Louis, M.; Tajeddine, N.; Gallo, C.; Van Schoor, M.; Anguish, I.; Cao, M.L.; Schakman, O.; Dietrich, A.; et al. Role of TRPC1 channel in skeletal muscle function. Am. J. Physiol. Physiol. 2010, 298, C149–C162. [Google Scholar] [CrossRef] [Green Version]
  93. Timchenko, N.A.; Iakova, P.; Cai, Z.-J.; Smith, J.R.; Timchenko, L. Molecular Basis for Impaired Muscle Differentiation in Myotonic Dystrophy. Mol. Cell. Biol. 2001, 21, 6927–6938. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  94. Tajhya, R.B.; Hu, X.; Tanner, M.; Huq, R.; Kongchan, N.; Neilson, J.R.; Rodney, G.; Horrigan, F.T.; Timchenko, L.T.; Beeton, C. Functional KCa1.1 channels are crucial for regulating the proliferation, migration and differentiation of human primary skeletal myoblasts. Cell Death Dis. 2016, 7, e2426. [Google Scholar] [CrossRef] [PubMed]
  95. Meola, G.; Cardani, R. Myotonic dystrophies: An update on clinical aspects, genetic, pathology, and molecular pathomechanisms. Biochim. et Biophys. Acta Mol. Basis Dis. 2015, 1852, 594–606. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  96. Li, F.; Yin, J.; Yue, T.; Liu, L.; Zhang, H. The CLIC5 (chloride intracellular channel 5) involved in C2C12 myoblasts proliferation and differentiation. Cell Biol. Int. 2010, 34, 379–384. [Google Scholar] [CrossRef] [PubMed]
  97. Stauber, T.; Jentsch, T.J. Chloride in Vesicular Trafficking and Function. Annu. Rev. Physiol. 2013, 75, 453–477. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  98. Iannotti, F.A.; Panza, E.; Barrese, V.; Viggiano, D.; Soldovieri, M.V.; Taglialatela, M. Expression, Localization, and Pharmacological Role of Kv7 Potassium Channels in Skeletal Muscle Proliferation, Differentiation, and Survival after Myotoxic Insults. J. Pharmacol. Exp. Ther. 2009, 332, 811–820. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  99. Iannotti, F.A.; Silvestri, C.; Mazzarella, E.; Martella, A.; Calvigioni, D.; Piscitelli, F.; Ambrosino, P.; Petrosino, S.; Czifra, G.; Biro, T.; et al. The endocannabinoid 2-AG controls skeletal muscle cell differentiation via CB1 receptor-dependent inhibition of Kv7 channels. Proc. Natl. Acad. Sci. USA 2014, 111, E2472–E2481. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  100. Iannotti, F.A.; Barrese, V.; Formisano, L.; Miceli, F.; Taglialatela, M. Specification of skeletal muscle differentiation by repressor element-1 silencing transcription factor (REST)-regulated Kv7.4 potassium channels. Mol. Biol. Cell 2013, 24, 274–284. [Google Scholar] [CrossRef] [Green Version]
  101. Wedhas, N.; Klamut, H.J.; Dogra, C.; Srivastava, A.K.; Mohan, S.; Kumar, A. Inhibition of mechanosensitive cation channels inhibits myogenic differentiation by suppressing the expression of myogenic regulatory factors and caspase-3 activity. FASEB J. 2005, 19, 1986–1997. [Google Scholar] [CrossRef] [Green Version]
  102. Formigli, L.; Meacci, E.; Sassoli, C.; Squecco, R.; Nosi, D.; Chellini, F.; Naro, F.; Francini, F.; Zecchi-Orlandini, S. Cytoskeleton/stretch-activated ion channel interaction regulates myogenic differentiation of skeletal myoblasts. J. Cell. Physiol. 2007, 211, 296–306. [Google Scholar] [CrossRef] [PubMed]
  103. Woo, J.S.; Cho, C.-H.; Kim, D.H.; Lee, E.H. TRPC3 cation channel plays an important role in proliferation and differentiation of skeletal muscle myoblasts. Exp. Mol. Med. 2010, 42, 614–627. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  104. Langlois, S.; Xiang, X.; Young, K.; Cowan, B.J.; Penuela, S.; Cowan, K.N. Pannexin 1 and Pannexin 3 Channels Regulate Skeletal Muscle Myoblast Proliferation and Differentiation. J. Biol. Chem. 2014, 289, 30717–30731. [Google Scholar] [CrossRef] [Green Version]
  105. Araya, R.; Eckardt, D.; Maxeiner, S.; Krüger, O.; Theis, M.; Willecke, K.; Sáez, J.C. Expression of connexins during differentiation and regeneration of skeletal muscle: Functional relevance of connexin43. J. Cell Sci. 2005, 118, 27–37. [Google Scholar] [CrossRef] [Green Version]
  106. Araya, R.; Eckardt, D.; Riquelme, M.A.; Willecke, K.; Sáez, J.C. Presence and importance of connexin43 during myogenesis. Cell Commun. Adhes. 2003, 10, 451–456. [Google Scholar] [CrossRef]
  107. Afzali, A.M.; Ruck, T.; Herrmann, A.M.; Iking, J.; Sommer, C.; Kleinschnitz, C.; Preuβe, C.; Stenzel, W.; Budde, T.; Wiendl, H.; et al. The potassium channels TASK2 and TREK1 regulate functional differentiation of murine skeletal muscle cells. Am. J. Physiol. Physiol. 2016, 311, C583–C595. [Google Scholar] [CrossRef] [Green Version]
  108. Krause, R.M.; Hamann, M.; Bader, C.R.; Liu, J.H.; Baroffio, A.; Bernheim, L. Activation of nicotinic acetylcholine receptors increases the rate of fusion of cultured human myoblasts. J. Physiol. 1995, 489, 779–790. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  109. Kurosaka, M.; Ogura, Y.; Funabashi, T.; Akema, T. Involvement of Transient Receptor Potential Cation Channel Vanilloid 1 (TRPV1) in Myoblast Fusion. J. Cell. Physiol. 2016, 231, 2275–2285. [Google Scholar] [CrossRef] [PubMed]
  110. Obi, S.; Nakajima, T.; Hasegawa, T.; Nakamura, F.; Sakuma, M.; Toyoda, S.; Tei, C.; Inoue, T. Heat induces myogenic transcription factors of myoblast cells via transient receptor potential vanilloid 1 (Trpv1). FEBS Open Bio 2018, 9, 101–113. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  111. Mandai, S.; Furukawa, S.; Kodaka, M.; Hata, Y.; Mori, T.; Nomura, N.; Ando, F.; Mori, Y.; Takahashi, D.; Yoshizaki, Y.; et al. Loop diuretics affect skeletal myoblast differentiation and exercise-induced muscle hypertrophy. Sci. Rep. 2017, 7, 46369. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  112. Laflamme, M.A.; Murry, C.E. Heart regeneration. Nature 2011, 473, 326–335. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  113. Senyo, S.; Steinhauser, M.L.; Pizzimenti, C.L.; Yang, V.K.; Cai, L.; Wang, M.; Wu, T.-D.; Guerquin-Kern, J.-L.; Lechene, C.P.; Lee, R.T. Mammalian heart renewal by pre-existing cardiomyocytes. Nat. Cell Biol. 2013, 493, 433–436. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  114. Ali, S.R.; Hippenmeyer, S.; Saadat, L.V.; Luo, L.; Weissman, I.L.; Ardehali, R. Existing cardiomyocytes generate cardiomyocytes at a low rate after birth in mice. Proc. Natl. Acad. Sci. USA 2014, 111, 8850–8855. [Google Scholar] [CrossRef] [Green Version]
  115. Später, D.; Hansson, E.M.; Zangi, L.; Chien, K.R. How to make a cardiomyocyte. Development 2014, 141, 4418–4431. [Google Scholar] [CrossRef] [Green Version]
  116. Barreto, S.; Hamel, L.; Schiatti, T.; Yang, Y.; George, V. Cardiac Progenitor Cells from Stem Cells: Learning from Genetics and Biomaterials. Cells 2019, 8, 1536. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  117. Zwi, L.; Caspi, O.; Arbel, G.; Huber, I.; Gepstein, A.; Park, I.-H.; Gepstein, L. Cardiomyocyte Differentiation of Human Induced Pluripotent Stem Cells. Circulation 2009, 120, 1513–1523. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  118. Kleger, A.; Seufferlein, T.; Malan, D.; Tischendorf, M.; Storch, A.; Wolheim, A.; Latz, S.; Protze, S.; Porzner, M.; Proepper, C.; et al. Modulation of Calcium-Activated Potassium Channels Induces Cardiogenesis of Pluripotent Stem Cells and Enrichment of Pacemaker-Like Cells. Circulation 2010, 122, 1823–1836. [Google Scholar] [CrossRef] [Green Version]
  119. Wang, D.; Liu, C.; Liu, H.; Meng, Y.; Lin, F.; Gu, Y.; Wang, H.; Shang, M.; Tong, C.; Sachinidis, A.; et al. ERG1 Plays an Essential Role in Rat Cardiomyocyte Fate Decision by Mediating AKT Signaling. Stem Cells 2021, 39, 443–457. [Google Scholar] [CrossRef]
  120. Jeziorowska, D.; Korniat, A.; Salem, J.-E.; Fish, K.; Hulot, J.-S. Generating patient-specific induced pluripotent stem cells-derived cardiomyocytes for the treatment of cardiac diseases. Expert Opin. Biol. Ther. 2015, 15, 1–11. [Google Scholar] [CrossRef]
  121. Fu, J.; Yu, H.-M.; Wang, R.; Liang, J.; Yang, H.-T. Developmental regulation of intracellular calcium transients during cardiomyocyte differentiation of mouse embryonic stem cells1. Acta Pharmacol. Sin. 2006, 27, 901–910. [Google Scholar] [CrossRef]
  122. Janowski, E.; Cleemann, L.; Sasse, P.; Morad, M. Diversity of Ca2+ Signaling in Developing Cardiac Cells. Ann. New York Acad. Sci. 2006, 1080, 154–164. [Google Scholar] [CrossRef] [PubMed]
  123. Tyser, R.C.; Miranda, A.; Chen, C.-M.; Davidson, S.; Srinivas, S.; Riley, P.R. Calcium handling precedes cardiac differentiation to initiate the first heartbeat. eLife 2016, 5, e17113. [Google Scholar] [CrossRef] [PubMed]
  124. Linask, K.K.; Han, M.-D.; Artman, M.; Ludwig, C.A. Sodium-calcium exchanger (NCX-1) and calcium modulation: NCX protein expression patterns and regulation of early heart development. Dev. Dyn. 2001, 221, 249–264. [Google Scholar] [CrossRef] [PubMed]
  125. Reppel, M.; Sasse, P.; Malan, D.; Nguemo, F.; Reuter, H.; Bloch, W.; Hescheler, J.; Fleischmann, B.K. Functional expression of the Na+/Ca2+ exchanger in the embryonic mouse heart. J. Mol. Cell. Cardiol. 2007, 42, 121–132. [Google Scholar] [CrossRef]
  126. Li, X.; Karki, P.; Lei, L.; Wang, H.; Fliegel, L. Na+/H+ exchanger isoform 1 facilitates cardiomyocyte embryonic stem cell differentiation. Am. J. Physiol. Circ. Physiol. 2009, 296, H159–H170. [Google Scholar] [CrossRef]
  127. Hoorntje, T.; Alders, M.; van Tintelen, P.; van der Lip, K.; Sreeram, N.; van der Wal, A.; Mannens, M.; Wilde, A. Homozygous premature truncation of the HERG protein: The human HERG knockout. Circulation 1999, 100, 1264–1267. [Google Scholar] [CrossRef] [PubMed]
  128. Teng, G.Q.; Zhao, X.; Lees-Miller, J.P.; Quinn, F.R.; Li, P.; Rancourt, D.E.; London, B.; Cross, J.C.; Duff, H.J. Homozygous Missense N629D hERG (KCNH2) Potassium Channel Mutation Causes Developmental Defects in the Right Ventricle and Its Outflow Tract and Embryonic Lethality. Circ. Res. 2008, 103, 1483–1491. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  129. Splawski, I.; Timothy, K.W.; Sharpe, L.M.; Decher, N.; Kumar, P.; Bloise, R.; Napolitano, C.; Schwartz, P.J.; Joseph, R.; Condouris, K.; et al. CaV1.2 Calcium Channel Dysfunction Causes a Multisystem Disorder Including Arrhythmia and Autism. Cell 2004, 119, 19–31. [Google Scholar] [CrossRef] [Green Version]
  130. Splawski, I.; Timothy, K.W.; Decher, N.; Kumar, P.; Sachse, F.; Beggs, A.; Sanguinetti, M.C.; Keating, M.T. Severe arrhythmia disorder caused by cardiac L-type calcium channel mutations. Proc. Natl. Acad. Sci. USA 2005, 102, 8089–8096. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  131. Yoshida, T.; Owens, G.K. Molecular Determinants of Vascular Smooth Muscle Cell Diversity. Circ. Res. 2005, 96, 280–291. [Google Scholar] [CrossRef]
  132. Majesky, M.W. Developmental Basis of Vascular Smooth Muscle Diversity. Arter. Thromb. Vasc. Biol. 2007, 27, 1248–1258. [Google Scholar] [CrossRef] [Green Version]
  133. Xie, C.; Ritchie, R.P.; Huang, H.; Zhang, J.; Chen, Y.E. Smooth muscle cell differentiation in vitro: Models and underlying molecular mechanisms. Arterioscler. Thromb. Vasc. Biol. 2011, 31, 1485–1494. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  134. Sinha, S.; Iyer, D.; Granata, A. Embryonic origins of human vascular smooth muscle cells: Implications for in vitro modeling and clinical application. Cell. Mol. Life Sci 2020, 71, 2271–2288. [Google Scholar] [CrossRef] [Green Version]
  135. Owens, G.K.; Kumar, M.S.; Wamhoff, B.R. Molecular Regulation of Vascular Smooth Muscle Cell Differentiation in Development and Disease. Physiol. Rev. 2004, 84, 767–801. [Google Scholar] [CrossRef]
  136. Mack, C.P. Signaling Mechanisms That Regulate Smooth Muscle Cell Differentiation. Arter. Thromb. Vasc. Biol. 2011, 31, 1495–1505. [Google Scholar] [CrossRef] [Green Version]
  137. Zhang, M.-J.; Zhou, Y.; Chen, L.; Wang, Y.-Q.; Wang, X.; Pi, Y.; Gao, C.-Y.; Li, J.-C.; Zhang, L.-L. An overview of potential molecular mechanisms involved in VSMC phenotypic modulation. Histochem. Cell Biol. 2015, 145, 119–130. [Google Scholar] [CrossRef] [PubMed]
  138. Frismantiene, A.; Filippova, M.; Erne, P.; Resink, T.J. Smooth muscle cell-driven vascular diseases and molecular mechanisms of VSMC plasticity. Cell. Signal. 2018, 52, 48–64. [Google Scholar] [CrossRef]
  139. Kudryavtseva, O.; Aalkjaer, C.; Matchkov, V. Vascular smooth muscle cell phenotype is defined by Ca2+-dependent transcription factors. FEBS J. 2013, 280, 5488–5499. [Google Scholar] [CrossRef] [Green Version]
  140. Wan, X.-J.; Zhao, H.-C.; Zhang, P.; Huo, B.; Shen, B.-R.; Yan, Z.-Q.; Qi, Y.-X.; Jiang, Z.-L. Involvement of BK channel in differentiation of vascular smooth muscle cells induced by mechanical stretch. Int. J. Biochem. Cell Biol. 2015, 59, 21–29. [Google Scholar] [CrossRef]
  141. Johnson, M.T.; Gudlur, A.; Zhang, X.; Xin, P.; Emrich, S.M.; Yoast, R.; Courjaret, R.; Nwokonko, R.M.; Li, W.; Hempel, N.; et al. L-type Ca2+ channel blockers promote vascular remodeling through activation of STIM proteins. Proc. Natl. Acad. Sci. USA 2020, 117, 17369–17380. [Google Scholar] [CrossRef] [PubMed]
  142. Zhang, B.; Chen, L.; Bai, Y.; Song, J.; Cheng, J.; Ma, H.; Ma, J.; Xie, M. miR-137 and its target T-type Ca V 3.1 channel modulate dedifferentiation and proliferation of cerebrovascular smooth muscle cells in simulated microgravity rats by regulating calcineurin/NFAT pathway. Cell Prolif. 2020, 53, e12774. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  143. Wamhoff, B.; Bowles, D.; McDonald, O.; Sinha, S.; Somlyo, A.; Owens, G. L-type Voltage-Gated Ca2+ Channels Modulate Expression of Smooth Muscle Differentiation Marker Genes via a Rho Kinase/Myocardin/SRF–Dependent Mechanism. Circ. Res. 2004, 95, 406–414. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  144. Thakore, P.; Pritchard, H.A.T.; Griffin, C.S.; Yamasaki, E.; Drumm, B.T.; Lane, C.; Sanders, K.M.; Earley, Y.F.; Earley, S. TRPML1 channels initiate Ca2+ sparks in vascular smooth muscle cells. Sci. Signal. 2020, 13, 637. [Google Scholar] [CrossRef] [PubMed]
  145. Durlu-Kandilci, N.T.; Ruas, M.; Chuang, K.-T.; Brading, A.; Parrington, J.; Galione, A. TPC2 Proteins Mediate Nicotinic Acid Adenine Dinucleotide Phosphate (NAADP)- and Agonist-evoked Contractions of Smooth Muscle. J. Biol. Chem. 2010, 285, 24925–24932. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  146. Berra-Romani, R.; Mazzocco-Spezzia, A.; Pulina, M.V.; Golovina, V.A. Ca2+handling is altered when arterial myocytes progress from a contractile to a proliferative phenotype in culture. Am. J. Physiol. Physiol. 2008, 295, C779–C790. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  147. Boie, S.; Chen, J.; Sanderson, M.J.; Sneyd, J. The relative contributions of store-operated and voltage-gated Ca2+ channels to the control of Ca2+ oscillations in airway smooth muscle. J. Physiol. 2016, 595, 3129–3141. [Google Scholar] [CrossRef] [Green Version]
  148. Kim, B.; Molina, R.; Jensen, G.; Poburko, D. A differentiated Ca2+ signalling phenotype has minimal impact on myocardin expression in an automated differentiation assay using A7r5 cells. Cell Calcium 2021, 96, 102369. [Google Scholar] [CrossRef]
  149. Teng, G.; Zhao, X.; Lees-Miller, J.P.; Belke, D.; Shi, C.; Chen, Y.; O’Brien, E.R.; Fedak, P.W.; Bracey, N.; Cross, J.C.; et al. Role of mutation and pharmacologic block of human KCNH2 in vasculogenesis and fetal mortality: Partial rescue by transforming growth factor-β. Circulation. Arrhythmia Electrophysiol. 2015, 8, 420–428. [Google Scholar] [CrossRef] [Green Version]
  150. Numaga-Tomita, T.; Shimauchi, T.; Oda, S.; Tanaka, T.; Nishiyama, K.; Nishimura, A.; Birnbaumer, L.; Mori, Y.; Nishida, M. TRPC6 regulates phenotypic switching of vascular smooth muscle cells through plasma membrane potential-dependent coupling with PTEN. FASEB J. 2019, 33, 9785–9796. [Google Scholar] [CrossRef] [Green Version]
  151. Bates, E. Ion Channels in Development and Cancer. Annu. Rev. Cell Dev. Biol. 2015, 31, 231–247. [Google Scholar] [CrossRef] [PubMed]
  152. Sundelacruz, S.; Levin, M.; Kaplan, D.L. Role of Membrane Potential in the Regulation of Cell Proliferation and Differentiation. Stem Cell Rev. Rep. 2009, 5, 231–246. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Myogenic differentiation. The scheme shows the differentiation of skeletal muscle from the mesoderm that is determined in the skeletal muscle lineage after MyoD and Myf5 expression. Myoblasts proliferate until they withdraw from the cell cycle and differentiate into myocytes, which involves myogenin and MRF4. Myocytes fuse to form myotubes that express skeletal muscle proteins myosin heavy chain (MHC), muscle creatine kinase (MCK), and α-actin. Adapted from [31].
Figure 1. Myogenic differentiation. The scheme shows the differentiation of skeletal muscle from the mesoderm that is determined in the skeletal muscle lineage after MyoD and Myf5 expression. Myoblasts proliferate until they withdraw from the cell cycle and differentiate into myocytes, which involves myogenin and MRF4. Myocytes fuse to form myotubes that express skeletal muscle proteins myosin heavy chain (MHC), muscle creatine kinase (MCK), and α-actin. Adapted from [31].
Ijms 22 13615 g001
Figure 2. Membrane hyperpolarization and calcium signaling in myoblast differentiation. The sequential activity of EAG and Kir2.1 K+ channels leads to membrane hyperpolarization, which in turn is required for Ca2+ signaling. Ca2+ release from the ER, which leads to activation of SOCE, can contribute to the Ca2+ signal. Activity of the Cl channel VRAC contributes to hyperpolarization by an unknown mechanism. See the main text for details.
Figure 2. Membrane hyperpolarization and calcium signaling in myoblast differentiation. The sequential activity of EAG and Kir2.1 K+ channels leads to membrane hyperpolarization, which in turn is required for Ca2+ signaling. Ca2+ release from the ER, which leads to activation of SOCE, can contribute to the Ca2+ signal. Activity of the Cl channel VRAC contributes to hyperpolarization by an unknown mechanism. See the main text for details.
Ijms 22 13615 g002
Figure 3. Cardiac differentiation. The scheme represents the different cell states during cardiac differentiation from pluripotent stem cells with the corresponding molecular markers. Adapted from [120].
Figure 3. Cardiac differentiation. The scheme represents the different cell states during cardiac differentiation from pluripotent stem cells with the corresponding molecular markers. Adapted from [120].
Ijms 22 13615 g003
Figure 4. Membrane hyperpolarization and ion channel activation in cardiac differentiation. The activity of ERG1 and SKCa K+ channels lead to membrane hyperpolarization and signaling to cardiac differentiation. See the main text for details.
Figure 4. Membrane hyperpolarization and ion channel activation in cardiac differentiation. The activity of ERG1 and SKCa K+ channels lead to membrane hyperpolarization and signaling to cardiac differentiation. See the main text for details.
Ijms 22 13615 g004
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Chen, L.; Hassani Nia, F.; Stauber, T. Ion Channels and Transporters in Muscle Cell Differentiation. Int. J. Mol. Sci. 2021, 22, 13615. https://doi.org/10.3390/ijms222413615

AMA Style

Chen L, Hassani Nia F, Stauber T. Ion Channels and Transporters in Muscle Cell Differentiation. International Journal of Molecular Sciences. 2021; 22(24):13615. https://doi.org/10.3390/ijms222413615

Chicago/Turabian Style

Chen, Lingye, Fatemeh Hassani Nia, and Tobias Stauber. 2021. "Ion Channels and Transporters in Muscle Cell Differentiation" International Journal of Molecular Sciences 22, no. 24: 13615. https://doi.org/10.3390/ijms222413615

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