Next Article in Journal
Specific NLRP3 Inflammasome Assembling and Regulation in Neutrophils: Relevance in Inflammatory and Infectious Diseases
Next Article in Special Issue
Store-Operated Calcium Entry and Its Implications in Cancer Stem Cells
Previous Article in Journal
Mitotic Maturation Compensates for Premature Centrosome Splitting and PCM Loss in Human cep135 Knockout Cells
Previous Article in Special Issue
CRACking the Molecular Regulatory Mechanism of SOCE during Platelet Activation in Thrombo-Occlusive Diseases
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

On the Connections between TRPM Channels and SOCE

by
Guilherme H. Souza Bomfim
1,
Barbara A. Niemeyer
2,
Rodrigo S. Lacruz
1,* and
Annette Lis
3,*
1
Department of Molecular Pathobiology, New York University College of Dentistry, New York, NY 10010, USA
2
Department of Molecular Biophysics, Center for Integrative Physiology and Molecular Medicine, School of Medicine, Saarland University, 66421 Homburg, Germany
3
Department of Biophysics, Center for Integrative Physiology and Molecular Medicine, School of Medicine, Saarland University, 66421 Homburg, Germany
*
Authors to whom correspondence should be addressed.
Cells 2022, 11(7), 1190; https://doi.org/10.3390/cells11071190
Submission received: 31 January 2022 / Revised: 23 March 2022 / Accepted: 30 March 2022 / Published: 1 April 2022
(This article belongs to the Collection STIM and Orai Communication in Health and Disease)

Abstract

:
Plasma membrane protein channels provide a passageway for ions to access the intracellular milieu. Rapid entry of calcium ions into cells is controlled mostly by ion channels, while Ca2+-ATPases and Ca2+ exchangers ensure that cytosolic Ca2+ levels ([Ca2+]cyt) are maintained at low (~100 nM) concentrations. Some channels, such as the Ca2+-release-activated Ca2+ (CRAC) channels and voltage-dependent Ca2+ channels (CACNAs), are highly Ca2+-selective, while others, including the Transient Receptor Potential Melastatin (TRPM) family, have broader selectivity and are mostly permeable to monovalent and divalent cations. Activation of CRAC channels involves the coupling between ORAI1-3 channels with the endoplasmic reticulum (ER) located Ca2+ store sensor, Stromal Interaction Molecules 1-2 (STIM1/2), a pathway also termed store-operated Ca2+ entry (SOCE). The TRPM family is formed by 8 members (TRPM1-8) permeable to Mg2+, Ca2+, Zn2+ and Na+ cations, and is activated by multiple stimuli. Recent studies indicated that SOCE and TRPM structure-function are interlinked in some instances, although the molecular details of this interaction are only emerging. Here we review the role of TRPM and SOCE in Ca2+ handling and highlight the available evidence for this interaction.

1. Introduction

Cation-conducting channel proteins in the plasma membrane play important roles in a multitude of cellular processes [1]. Most of these channels experience conformational changes from closed to open states allowing the passage of thousands of ions in response to chemical or mechanical signals [2]. Although these channels provide selective permeability to cations over anions determined by the amino acids lining the pore and the pore diameter, they can adopt multiple roles in cell signaling due to the variable selectivity for cations [2]. This is the case for the non-selective transient receptor potential melastatin (TRPM) family, which is capable of conducting monovalent (Na+ and K+) and divalent (Mg2+ and Ca2+) cations [3]. Less selective cation channels of the TRPC family (i.e., review by Saul and Hoth [4]) and of other families also contribute, but are not included in the present overview. By contrast, the specialized Ca2+ release-activated Ca2+ (CRAC) channels, mediating store-operated Ca2+ entry (SOCE), which are generated by the ORAI1-3 proteins, are a 1000-fold more selective for Ca2+ than Na+ ions [5]. Recent studies indicate that the TRPM and the CRAC channels interact in some expected ways but also by novel mechanisms [6,7]. Here we specifically review the role of TRPM channels in Ca2+ homeostasis and highlight recent advances toward understanding the potential synergy between TRPM and ORAI channels.

1.1. General Features of TRPM Channels

TRPM channels are a subfamily of the TRP superfamily composed of eight members denoted as TRPM1 to TRPM8 and grouped based on sequence homology as follows: (1) TRPM1 and TRPM3; (2) TRPM2 and TRPM8; (3) TRPM4 and TRPM5; and (4) TRPM6 and TRPM7 [3]. With the exception of the monovalent selective TRPM4 and TRPM5 channels, the remainder of the proteins in the family are non-selective cation channels that participate in a heterogeneous range of physiological processes including temperature and redox sensing, light sensing, embryonic development and Mg2+ homeostasis [3,8,9]. Although TRPM4 and TRPM5 are Ca2+ activated channels, they are impermeable to Ca2+ [3,10,11]. The general structural architecture of TRP channels consists of six transmembrane helical domains (TM1-TM6), with a loop between TM5 and TM6 forming the channel pore, and N- and C-terminal regions located in the cytosol [3,11]. Each TRPM subunit has a ~850 amino acids cytosolic domain, making TRPM members the largest proteins of the TRP superfamily [9,10]. The ion channel regions of TRPM2, TRPM6 and TRPM7 are linked to an intrinsic enzymatic domain within the C-terminus [12,13]. TRPM6 and TRPM7 contain serine/threonine α-kinase domains, while TRPM2 channel has a NUDT9-H domain [12,14]. These chanzymes modulate cellular functions either by inward ion currents through the pore and/or by phosphorylating downstream proteins via its enzymatic domain [3,10]. Table 1 lists features of the TRPM channels including their ion permeability and activation. Mutations in the genes encoding TRPM channels result in channelopathies including cancer [15], neuropathic pain [16], inflammation [17], hypertension [18,19], diabetes [20] and hypomineralization [21,22]. The TRPM channel pore domain located between TM5-6, surrounded by the S1-S4 domain and connected through the S4-S5 portion, appear to play an important role in channel gating [3,10]. All members of the TRPM family have a conserved Ca2+ binding site, but available data show that this binding site is important for the gating of only TRPM2 [23] and TRPM8 [24]. Several studies have indicated that TRPM channels are expressed in the membranes of intracellular organelles in addition to their plasmalemmal localization [3,6,25], but their functions in organelles are not completely understood and will not be discussed here.

1.2. General Features of the CRAC Channels

CRAC channels are formed by the endoplasmic reticulum (ER) resident Ca2+ sensors stromal interaction molecules 1-2 (STIM1/2) [26,27], and by the highly Ca2+-selective conductance pore ORAI1-3 proteins [28,29], forming the dominant store-operated Ca2+ entry denominated as SOCE [30,31,32]. The most common activation mechanism of SOCE involves the binding of a ligand to receptors in the plasma membrane (PM) which then couple and activate phospholipase C enzymes (PLC) to produce inositol 1,4,5 triphosphate (IP3) and diacylglycerol (DAG). The ensuing binding of IP3 to its receptor in the ER membrane results in a rapid decline in ER Ca2+ concentration, which subsequently causes oligomerization, migration of STIM1/2 to ER-PM contact sites, where the unfolded C-terminal CAD/SOAR domains trap and couple to the ORAI1-3 channels activating SOCE [5,33,34]. The hetero multimerization between STIM1 and STIM2 and different ORAI1-3 subunits result in distinct SOCE and CRAC biophysical properties [35,36,37,38,39]. Reports showed that loss-of-function mutations or knockdown of ORAI2 and ORAI3 genes results in an enhancement of SOCE [33,40,41,42,43]. Also, ORAI3 tunes-down efficient STIM1 gating when in a heteromeric complex with ORAI1 channels [44].
There are several well-known pharmacological inhibitors of SOCE and a major challenge has been to identify molecules with suitable characteristics (i.e., potency, selectivity, toxicology) to offer positive clinical applications [32,45]. The first generation of SOCE inhibitors such as MRS-1845, lanthanides (Gd3+/La3+), imidazoles (SKF-96365) and 2-APB, helped define basic SOCE cellular properties [46], although their poor selectivity significantly limited their use [32]. Newer SOCE inhibitors including compounds such as synta-66, BTP2 (YM-58483), RO2959, AnCoA4 and GSK-7975A have shown improved pharmacological characteristics in terms of potency and selectivity [46,47]. Synta-66, CM4620, RO2959 and AnCoA4 appear to be effective tools showing no significant off-target effects on TRP or voltage-gated channels [32,48,49]. The in vivo potency, efficacy and selectivity of these SOCE inhibitors remain less explored, but recent studies indicate that several new compounds are now in clinical trials [32,47,50,51].

2. TRPM1, TRPM2 and TRPM3 Channels

The first member of the TRPM family cloned and identified was TRPM1 in 1998 [52]. TRPM1 channels are involved in photoresponses in retinal cells in Drosophila and in mice [53,54]. In humans, TRPM1 is also associated with skin pigmentation and homozygous loss of TRPM1 results in retinal blindness [53,55]. Although TRPM1 and TRPM3 channels share ~70% sequence homology, their functions are quite different with TRPM3 acting as a thermoreceptor in detecting noxious hot (~40 °C) temperatures and heat-associated inflammation [54,56]. In addition to Ca2+, TRPM1 is permeable to other divalent ions such as Mn2+ and Mg2+, while TRPM3 has permeability to both monovalent and divalent cations [57]. The ionic conductance of TRPM1 channels is ~76 pS and ~65 to 130 pS for TRPM3 [53,57,58].
The activation mechanism of the TRPM1 channels has not been fully elucidated yet. Recent findings suggest that the intracellular uncoupling of the Gαo/Gβγ subunit of G-proteins, after its activation, leads to the channel’s closure [59]. Also, the activation of protein kinase C-alpha (PKCα) reduces the inhibition of TRPM1 by Mg2+ ions [60]. The heat compound capsaicin has been used to investigate TRPM1 function although it is a non-selective agonist [61]. TRPM3 channels can be activated by pregnenolone sulfate, CIM0216 and hypotonic solutions [56,62] that induces an increase in [Ca2+]cyt leading to Ca2+/calmodulin modulation and subsequent activation of mitogen-activated protein kinases (MAPKs) [57,62]. The lack of selective agonists of TRPM3 channels is evidenced by its activation by several metabolites and synthetic and plant-derived compounds including cholesterol, mefenamic acid, and the antidiabetic PPARγ-agonists rosiglitazone and troglitazone [56,57,62]. Functional studies using HEK-293 cells overexpressing TRPM3 channels showed that flavanones reduced pregnenolone sulfate-induced [Ca2+]cyt elevation [63]. Also, changes in ion concentration can negatively affect channel activity with increases in [Ca2+]cyt inhibiting TRPM1 and TRPM3 channels, intracellular Zn2+ blocking TRPM1, and Mg2+ inhibiting TRPM3 channels [8,64].
TRPM2 channels, which contain an enzymatic domain, sense warm temperatures and are also associated with the inflammatory cascade [65]. For example, Trpm2-deficient mice are prone to Listeria-mediated infections [66]. These channels are also considered as oxidative stress-sensitive and Ca2+-permeable ion channels [65]. The biophysical features of TRPM2 include a large pore, an intracellular Ca2+ binding site linked to the pore and a cation non-selectivity without voltage-dependence [23]. The single-channel conductance is ~60–80 pS and the PCa/PNa permeability ratio is ~0.7–0.9, indicating that cation influx is predominantly of Na+ ions [67,68]. Its permeability to Ca2+ and Mg2+ is maintained by amino acid residues located between the pore helix and the selectivity filter, being regulated by ADP ribose (ADPR), reactive oxygen and nitrogen species (ROS/RNS), and the phospholipid phosphatidylinositol 4,5-bisphosphate (PIP2) [19,65]. Recent evidence suggests that the activation of TRPM2 channels is linked to its enzymatic domain [69,70,71,72]. Interestingly, TRPM2 is catalytically inactive in humans but not in invertebrates [72]. At least one study reported a connection between TRPM2 and SOCE, although this link is indirect [73]. Salivary glands require SOCE for fluid secretion and radiation treatment of the gland activated a TRPM2-dependent pathway involving mitochondria which caused a caspase-mediated cleavage of STIM1 and loss of SOCE [73].

3. TRPM4/TRPM5 Channels

Unlike other TRPM family members, TRPM4 and TRPM5 channels are unique due to their much higher permeability to monovalent cations compared to divalent ions [74,75]. Their role in altering Ca2+ homeostasis is therefore more indirect. However, both channels require increases in intracellular Ca2+ levels to become activated and are also regulated by PIP2. TRPM5 channels are 5 to 10-fold more sensitive to Ca2+ when compared to TRPM4 [74,76,77]. TRPM5 was originally identified as MTR1, and was later found to be co-expressed with the taste signaling molecule α-gustducin [78,79]. TRPM4 has been identified as a homolog of MLSN (TRPM1) and is highly expressed in heart, kidney, prostate and colon and as a Ca2+ activated cation channel mediating membrane depolarization [80,81]. CAM binding sites in the C-terminus of TRPM4 are essential for regulating the sensitivity of direct Ca2+ dependent activation [76,82]. However, both TRPM4 and TRPM5 have a direct Ca2+ binding site on the intracellular side of the S1–S4 domain [83]. TRPM4, but not TRPM5, is inhibited by extracellular adenosine nucleotides (AMP, ADP, ATP) and its activity can be enhanced by PKC dependent phosphorylation of the TRP domain [84].
Increased TRPM4 expression correlates with decreased SOCE due to changes in the driving force for Ca2+ as has initially been demonstrated in prostate cancer cells [85]. A later report correlating the expression of TRPM4 with proliferation, cell cycle progression and invasion by colorectal cancer cells [86], but see also its role for cell spreading, migration and contractile behavior [87]. Within the heart, direct pathophysiological roles for TRPM4 that are likely independent from a concomitant inhibition of SOCE, are linked to smooth muscle depolarization and subsequent myogenic vasoconstriction [88]. Also, a recent review [8] provided the role of TRPM channels in several human diseases. In the context of SOCE regulation, the role of TRPM4 in the immune system is of particular interest. Bone marrow derived mast cells from TRPM4 knockout mice showed a greater release of leukotrienes and TNF-α as well as of histamine [89], but so far it has not been formally investigated to what extent these effects are dependent or independent from differential driving forces for Ca2+ influx from the outside. In T-lymphocytes, TRPM4-dependent alterations of cytokine production and altered Ca2+ oscillations as well as differential effects on NFATc1 localization [90,91], might also be caused by alterations of the electrical driving force. Still, the exact mechanism and/or the differential contribution of TRPM4 to immune responses remains to be fully understood and investigated for the response to differential T cell agonists. How ORAI and TRP channels interact has also been reviewed by Saul et al. [4]. As both TRPM4 and TRPM5 need Ca2+ for activation, a dual interaction with SOCE is likely with SOCE providing a source of Ca2+ and TRPM4/5 mediated depolarization subsequently providing negative feedback on SOCE mediated Ca2+ influx.

4. TRPM6 Channels

TRPM6 channels are involved in maintaining Mg2+ and Ca2+ homeostasis [13] and their activity is essential in the kidney and small intestine and in mammary epithelial cells and colon cells [12]. Mutations in the gene encoding TRPM6 cause hereditary disease of familial hypomagnesemia with secondary hypocalcemia [8]. TRPM6 channels can induce inward divalent cation currents when the intracellular Mg2+ levels are ~500 µM [12,14,92]. The permeability of TRPM6 to divalent cations is dependent on key acid residues present in its selectivity filter [12,93] and its ion conductance is estimated to be around 82–84 pS [12,94]. Molecular analysis based on amino acid sequences has revealed that TRPM6 and TRPM7 channels are close homologues, sharing a key feature of harboring the C-terminal serine/threonine protein kinase domain [3,10]. As shown in HEK-293 cells, TRPM6 specifically interacts with TRPM7 proteins forming complexes in the PM [95]. The TRPM6/7 complex has different biophysical properties compared to homomers of TRPM6, including its permeability to Ni2+, pore structure, inhibition by 2-APB, sensitivity to low (4–6) pH and conductance, ranging between 40 to 105 pS for TRPM7 and 56.6 pS for TRPM6/7 heteromeric form [3,12,94]. Additionally, the activation of TRPM6 channels is modified in the heteromeric form [92]. While TRPM6 homomers are inactive under basal Mg2+ levels, in the oligomeric form TRPM6 can be active after TRPM7 phosphorylation and lacks sensitivity to Mg2+ [92,94]. Despite the close homology of TRPM6 and TRPM7 channels they have different functions, with TRPM6 being involved in intestinal uptake and renal reabsorption of Mg2+, and TRPM7 regulating cellular Mg2+ homeostasis [12,14].

5. TRPM7 Channels

The TRPM7 presents a unique combination of an ion channel with a serine/threonine kinase and contributes to numerous physiological functions [96] (Figure 1). Besides being implicated in maintaining intracellular and systemic Mg2+ homeostasis [97,98,99,100,101,102], TRPM7 has been linked to cell motility, proliferation, differentiation, volume regulation, migration, and apoptosis [98,103,104,105,106,107,108,109,110,111,112,113,114,115,116,117,118,119]. This channel-kinase symbiosis paired with its ubiquitous expression gives it a central and non-redundant role in cellular processes. Its pathophysiology is broad being linked to neurodegenerative disorders, hypertension, and tissue fibrosis and to atypical immune responses [120,121,122,123,124,125,126,127]. As versatile as the function of TRPM7 is, so is its regulation by a variety of internal and external cellular factors. They range from intracellular cations, Mg-ATP, Cl and Br concentration, and intracellular pH to hydrolysis of the PIP2 [128,129,130,131]. The channel is constitutively active and conducts preferentially Zn2+, Mg2+ and Ca2+, and trace metals [103,115,132,133,134]. The constitutive active current is suppressed by intracellular levels of Mg2+ and Mg-ATP [103] and external Mg2+ acts as a permanent blocker of the pore [92,103,130,134]. These features led to the earlier designation of the native TRPM7 current as MagNuM (Mg2+-nucleotide-regulated metal ion current) [103,135] or MIC (Mg2+-inhibited cation current) [36]. TRPM7 permeability for Zn2+ is 4-fold higher than Ca2+ [134] and its deletion protects cells from Zn2+ and Ca2+ induced toxicity [136,137,138].
The kinase domain of TRPM7 belongs to the family of the atypical α-kinases [139] with the predisposition to phosphorylate serine and threonine residues in the context of an α-helix. It shows close homology to eukaryotic elongation factor-2 kinase (eEF2K), Dictyostelium myosin heavy chain kinases (MHCK) A-C and alpha-kinases 1-3 [140]. The first targets of the TRPM7 kinase were identified in vitro, including annexin A1 [141], PLC [142], and the MHCK A-C isoforms [105,143]. Novel mouse models made it possible to expand this diversity by the identification of native kinase substrates like SMAD2 [127,144]. Furthermore, the autophosphorylation of kinases supports target recognition and subsequent phosphorylation of substrates and appears not to be crucial for its catalytic activity [145,146,147]. Beside the necessity of Mn2+ or Mg2+ and Mg-ATP for the kinases activation and phosphorylation [146], the phosphorylation of some targets seems to be a Ca2+-dependent process [105,141,143]. This dependence suggests that channel activity somehow induces Ca2+ influx to support the kinase-target interaction. Additionally, the TRPM7 kinase can be cleaved by caspases to release the kinase domain, without losing its phosphotransferase activity from the channel and act on apoptotic signaling through the Fas receptor [119]. Interestingly, the cleaved kinase translocates to the nucleus affecting mRNA expression of TRPM7-dependent genes by modifying phosphorylation of serines and threonines on specific histone residues in a Zn2+ -dependent way [148].
The combination of a channel and a kinase poses a significant challenge to researchers when investigating the causes of pathophysiological processes. Since the discovery of TRPM7, the research has focused on the relationship between the channel and kinase activity and the physiological roles of the channel versus the kinase domain. This interplay between the TRPM7 channel and α-kinase activity affect each other, but the functional significance of this coupling is not clear and still the subject of ongoing investigation and debate. Reported inconsistencies result mainly from using a heterologous expression of TRPM7 mutants or by complete deletion of the kinase and the use of different tissue types [92,97,98,115,129,145,149]. A mouse model K1646R point mutation is introduced at TRPM7′s enzyme active site [96] opens the door to investigate the role of the kinase, uncoupled from the channel activity on physiological functions. Overall, it appears that the kinase activity of TRPM7 is not essential for the native channel function [96] and may play a more structural role in channel assembly or subcellular localization [97,98,145]. However, the kinase domain might still be important for Mg2+ levels ([Mg2+]cyt) mediated modulation of the TRPM7 channel itself, since the complete deletion of the domain increases [Mg2+]cyt sensitivity of the channel [97,98,128] in contrast to inactive kinase point mutant K1646R [145]. Accordingly, the defect in Mg2+ homeostasis and TRPM7 current reduction is only found in heterozygous delta-kinase but not in K1646R mice [96,97,127].

6. TRPM8 Channels

The seminal discovery of TRPV1 by the Julius laboratory in 1999 [150] with the follow-up study of impaired nociception and pain sensation in mice lacking TRPV1 [151] and of the cold sensitive TRPM8 published in 2002 by the Julius [152] and Patapoutian [17] groups were groundbreaking. These studies were a critical step in recognizing that sensitization of primary afferent neurons via ion channels is responsible for detecting skin surface temperatures and thermal-related neuropathic pain. TRPM8 was originally cloned by screening cDNA isolated from trigeminal sensory neurons for their ability to respond to menthol and cold stimuli and from DRG neurons looking for novel sensory TRP channels [17,152], placing the TRPM subfamily at the center stage of thermal somatosensation [153]. However, the ability to sense changes in temperature is not restricted to a given TRP subfamily. The significance of these findings and the discovery of the touch-sensitive Piezo channels [154] won the discoverers Dr. David Julius and Dr. Ardem Patapoutian the 2021 Nobel Prize in Physiology [152,155]. TRPM8 channels are non-selective Ca2+-permeable channels exhibiting multi-gating mechanisms, being activated by innocuous cool to cold temperatures and regulated by crucial molecules such as Ca2+ and PIP2 [155,156]. The early work by the Latorre group identified the C-terminally located PIP2 sensing domain as one determinant of temperature sensitivity [157,158]. The permeability ratio between Ca2+ and Na+ ions (PCa/PNa) range from 0.97 to 3.2, with monovalent ions conductance series of Cs+ > K+ > Na+ [159]. Also, TRPM8 channels can depolarize cells and activate voltage-gated Na+ and Ca2+ channels [153], leading to increased ion influx. Basal cytosolic Ca2+ is required for TRPM8 activation by the cooling agonist, icilin, but not for menthol activity [24,160]. A major break-through in the understanding of TRPM8 modulation is based on several highly relevant reports published in 2017 to 2019 describing the cryo-electron microscopy structures of TRPM8. Herein, Yin et al. initially resolved the cryo-electron microscopy (Cryo-EM) structure of a full-length TRPM8 from the collared flycatcher [24]. Their structures revealed a complex layered architecture with the menthol binding site located within the voltage-sensor like domain [24,161]. This was followed by a report from the Julius group, using TRPM8 from another bird (Parus major) to more clearly resolve the transmembrane and selectivity filter domains of TRPM8 in the antagonist or Ca2+ bound configurations [153]. A second follow-up report from the Lee group in 2019 further revealed allosteric coupling between binding of PIP2 and cooling compounds and also revealed that intracellular Ca2+ is not necessary for cold- or menthol-dependent TRPM8 activation, however it is necessary for TRPM8 activation by icilin [162]. In contrast to competition for binding of PIP2 versus agonist binding in TRPV1 [163], TRPM8 has nearby but distinct binding sites for PIP2 and agonists and PIP2 is required as a cofactor for channel activation by cooling agonists and cold temperatures [153,164]. PIP2 itself might be sufficient to activate TRPM8, whereas depletion of basal PIP2 levels results in channel desensitization [153]. Endogenous ligands including testosterone, artemin and Pirt (phosphoinositide interacting regulator of TRP) protein have been also proposed as physiological ligands of TRPM8 channels [165]. The modulation of TRPM8 channels have been considered a promising approach to develop novel therapeutic tools for chronic pain and noxious cold sensitization [165] and have also been relevant in the treatment of cancer, neuropathic pain and inflammation [15,165].
Table 1. TRPM1-8 channels ion influx characteristics including, enzymatic domain, gating, ion permeability, function, SOCE interaction and pharmacology.
Table 1. TRPM1-8 channels ion influx characteristics including, enzymatic domain, gating, ion permeability, function, SOCE interaction and pharmacology.
TRPM Channels: Ion Influx Characteristics
Name:Enzymatic DomainGating Ion PermeabilityFunctionSOCE InteractionPharmacologyReferences
TRPM1Noo and Gβγ subunits of G-proteinsDivalent (Ca2+/Mg2+/Mn2+)Skin Pigmentation Retinal Photoresponse NoActivator(s): Pregnenolone Inhibitor(s): ↑[Zn2+]cyt[51,52,58,60]
TRPM2Yes (NUDT9-H)ADP-ribose and Ca2+Monovalent (Na+/K+/Cs+) Divalent (Ca2+/Mg2+/Ba2+) Body Temperature Control Insulin/ROS/Immune Response Yes (Indirectly)Activator(s): ADP/ADPR analogues Inhibitor(s): Cacospongia/Scalaradial [23,64,65,66,67]
TRPM3NoGi,q-GPCRs Ca2+/CaM/MAPKsMonovalent (Na+/K+/Cs+) Divalent (Ca2+/Mg2+/Ba2+) Noxious Heat Sensation Glucose/Ca2+ HomeostasisNoActivator(s): CIM0216/Pregnenolone Inhibitor(s): ↑[Mg2+]cyt/Primidone[55,56,62]
TRPM4NoCa2+/CaMMonovalent (Na+ > K+ > Cs+ > Li+ >> Ca2+/Cl) Myogenic Tone, Cardiac Conduction, Ca2+ OscillationYes (Indirectly)Activator(s): ↑[Ca2+]cyt Inhibitor(s): AMP/ADP/ATP/DVT[8,11,63,75,77,80]
TRPM5NoCa2+/CaMMonovalent (Na+ ≥ K+ ≥ Cs+)Taste, Insulin SecretionNoActivator(s): ↑[Ca2+]cyt/PIP2/Rutamarin Inhibitor(s): TPPO[3,74,77,81]
TRPM6Yes (α-kinase)PIP2/PLCγ Mainly Mg2+/Ca2+ and other divalent (Ba2+/Zn2+/Mn2+)Mg2+ Homeostasis Embryonic DevelopmentNoActivator(s): ↓[Mg2+]cyt/EGF/Insulin Inhibitor(s): ATP/H2O2 [3,12,13,14,83,89]
TRPM7Yes (α-kinase)Phosphorylation PLCγ/Myosin IIA-C Mainly Mg2+/Ca2+ and other divalent (Ba2+/Zn2+/Mn2+)Mg2+ Homeostasis Cell Motility/DifferentiationYes (Indirectly)Activator(s): Naltriben/↓[Mg2+]cyt/PiP2 Inhibitor(s): NS8593/FTY720/WaxenicinA[6,7,25,109,117]
TRPM8Noq-GPCRs/PIP2Monovalent (Na+/K+/Cs+) Divalent (Ca2+/Mg2+/Ba2+) Cold Skin Temperatures Thermal Neuropathic Pain prostateYes (Indirectly)Activator(s): Menthol/Icilin/WS12 Inhibitor(s): AMTB/TCI2014/CPS-369 [24,152,154,159]

7. Discussion: The TRPM-SOCE Connection

The most evident connection between the TRPM channels and SOCE is represented by the TRPM7 channels. On the one hand, TRPM7, like SOCE, is implicated as a player in global Ca2+ levels and in mediating Ca2+ influx [104,141,166,167,168]. Moreover, Ca2+ influx through the channel seems to play a central role in controlling the recruitment of substrates, including annexin A1 and myosin II A heavy chain, to the TRPM7 kinase domain [105,141,143]. More importantly, recent reports highlighted the connections between TRPM7 and SOCE and their relevance to cell physiology [6,7,169]. Initial studies focused on the immune system cells because TRPM7 and SOCE play a significant role in immune cell development, activation, and the initiation of both innate and adaptive immune responses [33,116,170]. The deletion or pharmacological inhibition of TRPM7 reduces SOCE in DT 40 B-lymphocytes, which indicates a potential direct link between TRPM7 function and SOCE [6]. The observed reduction of CRAC currents (ICRAC) in the DT40 cells is neither due to membrane potential effects nor to indirect effects of K+ currents [6]. Furthermore, the authors exclude that TRPM7 channels are part of SOCE but considered these channels as SOCE regulators [6]. Rescue experiments in DT40 cells expressing kinase-dead (K1648R), or kinase-deficient mutant of TRPM7, provide evidence that regulation of SOCE takes place via its kinase domain [6]. Interestingly, TRPM7 participates in maintaining Ca2+ stores under resting conditions and contributes to ER store refilling after depletion [6].
An ensemble between SOCE and TRPM7 appears to support the intracellular Ca2+ balance under resting conditions and after activation of the Ca2+ signaling cascade [6]. In line with these findings, a second study used pharmacological (modulators) and molecular (siRNA) approaches to examine SOCE/TRPM7 connections in primary enamel forming cells (ameloblasts) and in the enamel cell line LS8 cells [7]. Especially during ameloblast differentiation, the TRPM7 kinase plays a role by phosphorylating the cAMP response element binding (CREB) protein [22]. The use of naltriben as a TRPM7 activator [7,171] enhances SOCE in rat-derived primary ameloblasts from the secretory and maturation stages [7]. Pharmacological suppression of TRPM7 pore does not decrease SOCE and excludes TRPM7 as a component of SOCE in these cells, supporting the non-critical role of TRPM7 channels on SOCE in ameloblasts [7]. Moreover, the activation of TRPM7 with naltriben in LS8 cells lacking both ORAI1 and ORAI2 (shORAI1-2) failed to increase cytosolic Ca2+ levels [7]. These findings are supported by data on primary ameloblasts of Stim1/2K14cre mice lacking Stim1 and Stim2, which had not been previously reported. Figure 2 shows that whereas naltriben enhances the SOCE peak in ameloblasts of wild type mice, the activation of TRPM7 in ameloblasts of Stim1/2K14cre mice, which show nearly abolished SOCE [28], failed to show any changes in Ca2+ influx. This supports the notion that the potentiating function of TRPM7 on SOCE likely requires the previous activation of SOCE, and that TRPM7 is not able to compensate for the lack of SOCE. Furthermore, additional rescue experiments of the TRPM7-KO-mediated phenotype with the inactive kinase mutant (K1648R) or hTRPM7 ∆-kinase could address the involvement of the channel or kinase in this potentiating function [6,7], as it has already been studied in B-lymphocytes.
Overall, additional studies seem to be required to elucidate the role of TRPM7 or its kinase in enamel cells and beyond. If the TRPM7 kinase is the regulatory portion in this context, the kinase inactivated mouse model should unmask its role in Ca2+ signaling. Inactivation of the TRPM7 α-kinase in a mouse model leads to splenomegaly with increased splenocytes numbers [172], but unaltered T cell subsets distribution in the spleen [127]. TRPM7 activity in murine splenic T cells is comparable in wild-type TRPM7 and KD mutant, highlighting the dispensability of the kinase function for ion conduction again [96,145]. Also, the SOCE magnitude tends to be larger in cells from KD mutant mice, and the initial slope is significantly increased, suggesting a potentiated Ca2+ influx through ORAI channels. The kinase deactivation causes potentiation of Ca2+ signals in resting conditions while a reduction occurs in the activated cells. The reasons can be multi-layered, including differences in metabolic stages of the cells or altered SOCE expression patterns within subpopulations of T cells [173,174,175].
Recent evidence highlighted the TRPM7 α-kinase domain as a further indirect modulatory player generating Ca2+ signals via SOCE [6]. So far, there is a lack of evidence supporting the active participation of TRPM7 in SOCE, at least in the cells studied so far [6,7]. This raises the next question: how does TRPM7 indirectly exert these regulatory abilities on the SOCE pathway? The obvious answer would be its ability to phosphorylate targets directly or indirectly involved in the SOCE pathway. The reversible phosphorylation of proteins catalyzed by kinases is central in regulatory mechanisms of signal transduction [128]. Several studies have already identified ORAI and STIM proteins as targets for kinases, leading to alterations of Ca2+ entry [176,177,178,179]. In addition, STIM and ORAI regulatory proteins including SOCE-associated regulatory factor (SARAF), CRACR2A, GOLLI proteins, caveolin and septin [180,181,182,183] may also be under the influence of kinases. Another interesting point is the involvement of TRPM7 to form the myosin II motor protein, bundle actin filaments, and build the actomyosin cytoskeleton network. The MHC-II phosphorylation by TRPM7 is Ca2+-dependent [105], and it relies on STIM1-mediated Ca2+ entry since STIM1 deletion abolished actomyosin formation [184]. These reciprocal interactions between TRPM7 and SOCE may require tight balance to maintain proper cell signaling and functioning.
Additional SOCE/TRPM links are provided by TRPM2 channels. Liu et al. reported that in irradiated salivary glands, TRPM2 channels are activated resulting in elevated mitochondrial Ca2+ and ROS which subsequently induced caspase-3 cleavage of STIM1 and loss of SOCE [73]. While these data strongly suggest a cause-effect association between SOCE and TRPM2, this is an indirect link as it does not suggest a direct interaction between TRPM2 and any of the SOCE components [73]. TRPM8 channels have also been proposed to antagonize the degree of SOCE. The downregulation of TRPM8 in pulmonary smooth muscle cells correlates with increased SOCE, and the application of icilin causes suppression of SOCE. However, the link is less clear and, thus far, likely indirect [185]. TRPM8 (and many other channels) activity can be influenced by alterations in Ca2+ and PIP2 levels [153], and are also affected by the activity of the GPCR Gαq-subunit which could disrupt the PLC-PIP2 signaling cascade [186] and therefore might indirectly affect SOCE. A general note of caution is that activation of cation selective channels with a sufficient ion flux capacity and high Na+ > Ca2+ permeability ratios will lead to membrane depolarization, which might indirectly reduce influx through STIM2 mediated pre-coupled ORAI channels, thus potentially lowering basal Ca2+.

8. Concluding Remarks

Since the identification of ORAI1-3 and STIM1/2 as key components of SOCE, several reports have suggested that members of the TRPM family are associated with SOCE. However, the available evidence at present does not support the consideration that TRPM channels are intrinsic components of SOCE. Nonetheless, at least two members of the TRPM family (TRPM2 and TRPM7) can modulate SOCE, albeit such modulation appears to be primarily indirect involving either the phosphorylation of SOCE components via the enzymatic domain of TRPM7, or via mitochondrial Ca2+ accumulation and ROS generation to degrade STIM1. Yet there are several gaps in understanding the nature of these modulatory functions, particularly evident in the case of TRPM7. Additional work is required to better discern the synergy between TRPM members and SOCE and its impact on the Ca2+ signaling cascade.

Author Contributions

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

Funding

Funding to R.S.L. was provided by the National Institute of Dental and Craniofacial Research (NIDCR) grants DE025639, DE027981 and DE027981. Funding to B.A.N was provided by Deutsche Forschungsgemeinschaft (DFG) grants SFB1027|C4, SFB894|A2 and TRR219|322900939|M04 and to A.L. by DFG grant LI 1750/4-2.

Institutional Review Board Statement

Procedures for the use of mice were conducted in accordance with the Institutional Animal Care and Use Committee (IACUC) protocol of New York University (s16-00625).

Acknowledgments

The authors would like to thank Isabelle Derler for her kind invitation to participate in this special collection “STIM and Orai Communication in Health and Disease”.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Gilabert, J.A. Cytoplasmic Calcium Buffering: An Integrative Crosstalk. Adv. Exp. Med. Biol. 2020, 1131, 163–182. [Google Scholar] [CrossRef] [PubMed]
  2. Moreau, A.; Gosselin-Badaroudine, P.; Chahine, M. Biophysics, pathophysiology, and pharmacology of ion channel gating pores. Front. Pharmacol. 2014, 5, 53. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. Huang, Y.; Fliegert, R.; Guse, A.H.; Lü, W.; Du, J. A structural overview of the ion channels of the TRPM family. Cell Calcium 2020, 85, 102111. [Google Scholar] [CrossRef] [PubMed]
  4. Saul, S.; Stanisz, H.; Backes, C.S.; Schwarz, E.C.; Hoth, M. How ORAI and TRP channels interfere with each other: Interaction models and examples from the immune system and the skin. Eur. J. Pharmacol. 2014, 739, 49–59. [Google Scholar] [CrossRef] [PubMed]
  5. Yeung, P.S.; Yamashita, M.; Prakriya, M. Molecular basis of allosteric Orai1 channel activation by STIM1. J. Physiol. 2020, 598, 1707–1723. [Google Scholar] [CrossRef]
  6. Faouzi, M.; Kilch, T.; Horgen, F.D.; Fleig, A.; Penner, R. The TRPM7 channel kinase regulates store-operated calcium entry. J. Physiol. 2017, 595, 3165–3180. [Google Scholar] [CrossRef] [Green Version]
  7. Souza Bomfim, G.H.; Costiniti, V.; Li, Y.; Idaghdour, Y.; Lacruz, R.S. TRPM7 activation potentiates SOCE in enamel cells but requires ORAI. Cell Calcium 2020, 87, 102187. [Google Scholar] [CrossRef]
  8. Jimenez, I.; Prado, Y.; Marchant, F.; Otero, C.; Eltit, F.; Cabello-Verrugio, C.; Cerda, O.; Simon, F. TRPM Channels in Human Diseases. Cells 2020, 9, 2604. [Google Scholar] [CrossRef]
  9. Nilius, B.; Owsianik, G. The transient receptor potential family of ion channels. Genome Biol. 2011, 12, 218. [Google Scholar] [CrossRef] [Green Version]
  10. Harteneck, C. Function and pharmacology of TRPM cation channels. Naunyn Schmiedebergs Arch. Pharmacol. 2005, 371, 307–314. [Google Scholar] [CrossRef] [Green Version]
  11. Winkler, P.A.; Huang, Y.; Sun, W.; Du, J.; Lü, W. Electron cryo-microscopy structure of a human TRPM4 channel. Nature 2017, 552, 200–204. [Google Scholar] [CrossRef] [PubMed]
  12. Schlingmann, K.P.; Waldegger, S.; Konrad, M.; Chubanov, V.; Gudermann, T. TRPM6 and TRPM7--Gatekeepers of human magnesium metabolism. Biochim. Biophys. Acta 2007, 1772, 813–821. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  13. Ferioli, S.; Zierler, S.; Zaißerer, J.; Schredelseker, J.; Gudermann, T.; Chubanov, V. TRPM6 and TRPM7 differentially contribute to the relief of heteromeric TRPM6/7 channels from inhibition by cytosolic Mg2+ and Mg·ATP. Sci. Rep. 2017, 7, 8806. [Google Scholar] [CrossRef] [PubMed]
  14. Chubanov, V.; Gudermann, T.; Schlingmann, K.P. Essential role for TRPM6 in epithelial magnesium transport and body magnesium homeostasis. Pflugers Arch. 2005, 451, 228–234. [Google Scholar] [CrossRef] [PubMed]
  15. Liu, Z.; Wu, H.; Wei, Z.; Wang, X.; Shen, P.; Wang, S.; Wang, A.; Chen, W.; Lu, Y. TRPM8: A potential target for cancer treatment. J. Cancer Res. Clin. Oncol. 2016, 142, 1871–1881. [Google Scholar] [CrossRef] [PubMed]
  16. Hossain, M.Z.; Bakri, M.M.; Yahya, F.; Ando, H.; Unno, S.; Kitagawa, J. The Role of Transient Receptor Potential (TRP) Channels in the Transduction of Dental Pain. Int. J. Mol. Sci. 2019, 20, 526. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  17. Peier, A.M.; Moqrich, A.; Hergarden, A.C.; Reeve, A.J.; Andersson, D.A.; Story, G.M.; Earley, T.J.; Dragoni, I.; McIntyre, P.; Bevan, S.; et al. A TRP channel that senses cold stimuli and menthol. Cell 2002, 108, 705–715. [Google Scholar] [CrossRef] [Green Version]
  18. Huang, F.; Ni, M.; Zhang, J.M.; Li, D.J.; Shen, F.M. TRPM8 downregulation by angiotensin II in vascular smooth muscle cells is involved in hypertension. Mol. Med. Rep. 2017, 15, 1900–1908. [Google Scholar] [CrossRef] [Green Version]
  19. Alves-Lopes, R.; Neves, K.B.; Anagnostopoulou, A.; Rios, F.J.; Lacchini, S.; Montezano, A.C.; Touyz, R.M. Crosstalk Between Vascular Redox and Calcium Signaling in Hypertension Involves TRPM2 (Transient Receptor Potential Melastatin 2) Cation Channel. Hypertension 2020, 75, 139–149. [Google Scholar] [CrossRef]
  20. Zsombok, A.; Derbenev, A.V. TRP Channels as Therapeutic Targets in Diabetes and Obesity. Pharmaceuticals 2016, 9, 50. [Google Scholar] [CrossRef] [Green Version]
  21. Nakano, Y.; Le, M.H.; Abduweli, D.; Ho, S.P.; Ryazanova, L.V.; Hu, Z.; Ryazanov, A.G.; Den Besten, P.K.; Zhang, Y. A Critical Role of TRPM7 As an Ion Channel Protein in Mediating the Mineralization of the Craniofacial Hard Tissues. Front. Physiol. 2016, 7, 258. [Google Scholar] [CrossRef] [PubMed]
  22. Ogata, K.; Tsumuraya, T.; Oka, K.; Shin, M.; Okamoto, F.; Kajiya, H.; Katagiri, C.; Ozaki, M.; Matsushita, M.; Okabe, K. The crucial role of the TRPM7 kinase domain in the early stage of amelogenesis. Sci. Rep. 2017, 7, 18099. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  23. Zhang, Z.; Tóth, B.; Szollosi, A.; Chen, J.; Csanády, L. Structure of a TRPM2 channel in complex with Ca2+ explains unique gating regulation. eLife 2018, 7, e36409. [Google Scholar] [CrossRef] [PubMed]
  24. Yin, Y.; Le, S.C.; Hsu, A.L.; Borgnia, M.J.; Yang, H.; Lee, S.Y. Structural basis of cooling agent and lipid sensing by the cold-activated TRPM8 channel. Science 2019, 363, eaav9334. [Google Scholar] [CrossRef] [PubMed]
  25. Abiria, S.A.; Krapivinsky, G.; Sah, R.; Santa-Cruz, A.G.; Chaudhuri, D.; Zhang, J.; Adstamongkonkul, P.; DeCaen, P.G.; Clapham, D.E. TRPM7 senses oxidative stress to release Zn2+ from unique intracellular vesicles. Proc. Natl. Acad. Sci. USA 2017, 114, e6079–e6088. [Google Scholar] [CrossRef] [Green Version]
  26. Zhang, S.L.; Yu, Y.; Roos, J.; Kozak, J.A.; Deerinck, T.J.; Ellisman, M.H.; Stauderman, K.A.; Cahalan, M.D. STIM1 is a Ca2+ sensor that activates CRAC channels and migrates from the Ca2+ store to the plasma membrane. Nature 2005, 437, 902–905. [Google Scholar] [CrossRef]
  27. Liou, J.; Kim, M.L.; Heo, W.D.; Jones, J.T.; Myers, J.W.; Ferrell, J.E., Jr.; Meyer, T. STIM is a Ca2+ sensor essential for Ca2+-store-depletion-triggered Ca2+ influx. Curr. Biol. 2005, 15, 1235–1241. [Google Scholar] [CrossRef] [Green Version]
  28. Feske, S.; Gwack, Y.; Prakriya, M.; Srikanth, S.; Puppel, S.H.; Tanasa, B.; Hogan, P.G.; Lewis, R.S.; Daly, M.; Rao, A. A mutation in Orai1 causes immune deficiency by abrogating CRAC channel function. Nature 2006, 441, 179–185. [Google Scholar] [CrossRef]
  29. 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]
  30. Prakriya, M.; Lewis, R.S. Store-Operated Calcium Channels. Physiol. Rev. 2015, 95, 1383–1436. [Google Scholar] [CrossRef] [Green Version]
  31. Putney, J.W., Jr. New molecular players in capacitative Ca2+ entry. J. Cell Sci. 2007, 120, 1959–1965. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  32. Bakowski, D.; Murray, F.; Parekh, A.B. Store-Operated Ca2+ Channels: Mechanism, Function, Pharmacology, and Therapeutic Targets. Annu. Rev. Pharmacol. Toxicol. 2021, 61, 629–654. [Google Scholar] [CrossRef]
  33. Yoast, R.E.; Emrich, S.M.; Zhang, X.; Xin, P.; Johnson, M.T.; Fike, A.J.; Walter, V.; Hempel, N.; Yule, D.I.; Sneyd, J.; et al. The native ORAI channel trio underlies the diversity of Ca2+ signaling events. Nat. Commun. 2020, 11, 2444. [Google Scholar] [CrossRef] [PubMed]
  34. Hogan, P.G.; Rao, A. Store-operated calcium entry: Mechanisms and modulation. Biochem. Biophys. Res. Commun. 2015, 460, 40–49. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  35. Peinelt, C.; Lis, A.; Beck, A.; Fleig, A.; Penner, R. 2-Aminoethoxydiphenyl borate directly facilitates and indirectly inhibits STIM1-dependent gating of CRAC channels. J. Physiol. 2008, 586, 3061–3073. [Google Scholar] [CrossRef] [PubMed]
  36. Kozak, J.A.; Kerschbaum, H.H.; Cahalan, M.D. Distinct properties of CRAC and MIC channels in RBL cells. J. Gen. Physiol. 2002, 120, 221–235. [Google Scholar] [CrossRef] [Green Version]
  37. DeHaven, W.I.; Smyth, J.T.; Boyles, R.R.; Putney, J.W., Jr. Calcium inhibition and calcium potentiation of Orai1, Orai2, and Orai3 calcium release-activated calcium channels. J. Biol. Chem. 2007, 282, 17548–17556. [Google Scholar] [CrossRef] [Green Version]
  38. Derler, I.; Fahrner, M.; Muik, M.; Lackner, B.; Schindl, R.; Groschner, K.; Romanin, C. A Ca2(+ )release-activated Ca2(+) (CRAC) modulatory domain (CMD) within STIM1 mediates fast Ca2(+)-dependent inactivation of ORAI1 channels. J. Biol. Chem. 2009, 284, 24933–24938. [Google Scholar] [CrossRef] [Green Version]
  39. Lis, A.; Peinelt, C.; Beck, A.; Parvez, S.; Monteilh-Zoller, M.; Fleig, A.; Penner, R. CRACM1, CRACM2, and CRACM3 are store-operated Ca2+ channels with distinct functional properties. Curr. Biol. 2007, 17, 794–800. [Google Scholar] [CrossRef] [Green Version]
  40. Hoth, M.; Niemeyer, B.A. The neglected CRAC proteins: Orai2, Orai3, and STIM2. Curr. Top. Membr. 2013, 71, 237–271. [Google Scholar] [CrossRef]
  41. Hou, X.; Outhwaite, I.R.; Pedi, L.; Long, S.B. Cryo-EM structure of the calcium release-activated calcium channel Orai in an open conformation. eLife 2020, 9, e62772. [Google Scholar] [CrossRef] [PubMed]
  42. Vaeth, M.; Kahlfuss, S.; Feske, S. CRAC Channels and Calcium Signaling in T Cell-Mediated Immunity. Trends Immunol. 2020, 41, 878–901. [Google Scholar] [CrossRef] [PubMed]
  43. Eckstein, M.; Vaeth, M.; Aulestia, F.J.; Costiniti, V.; Kassam, S.N.; Bromage, T.G.; Pedersen, P.; Issekutz, T.; Idaghdour, Y.; Moursi, A.M.; et al. Differential regulation of Ca2+ influx by ORAI channels mediates enamel mineralization. Sci. Signal. 2019, 12, eaav4663. [Google Scholar] [CrossRef] [PubMed]
  44. Alansary, D.; Bogeski, I.; Niemeyer, B.A. Facilitation of Orai3 targeting and store-operated function by Orai1. Biochim. Biophys. Acta 2015, 1853, 1541–1550. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  45. Stauderman, K.A. CRAC channels as targets for drug discovery and development. Cell Calcium 2018, 74, 147–159. [Google Scholar] [CrossRef] [PubMed]
  46. Bird, G.S.; Putney, J.W., Jr. Pharmacology of Store-Operated Calcium Entry Channels. In Calcium Entry Channels in Non-Excitable Cells; Kozak, J.A., Putney, J.W., Jr., Eds.; CRC Press: Boca Raton, FL, USA, 2018; pp. 311–324. [Google Scholar]
  47. Riva, B.; Griglio, A.; Serafini, M.; Cordero-Sanchez, C.; Aprile, S.; Di Paola, R.; Gugliandolo, E.; Alansary, D.; Biocotino, I.; Lim, D.; et al. Pyrtriazoles, a Novel Class of Store-Operated Calcium Entry Modulators: Discovery, Biological Profiling, and in Vivo Proof-of-Concept Efficacy in Acute Pancreatitis. J. Med. Chem. 2018, 61, 9756–9783. [Google Scholar] [CrossRef]
  48. Chen, G.; Panicker, S.; Lau, K.Y.; Apparsundaram, S.; Patel, V.A.; Chen, S.L.; Soto, R.; Jung, J.K.; Ravindran, P.; Okuhara, D.; et al. Characterization of a novel CRAC inhibitor that potently blocks human T cell activation and effector functions. Mol. Immunol. 2013, 54, 355–367. [Google Scholar] [CrossRef]
  49. Zhang, B.; Naik, J.S.; Jernigan, N.L.; Walker, B.R.; Resta, T.C. Reduced membrane cholesterol after chronic hypoxia limits Orai1-mediated pulmonary endothelial Ca2+ entry. Am. J. Physiol. Heart Circ. Physiol. 2018, 314, h359–h369. [Google Scholar] [CrossRef]
  50. Rahman, S.; Rahman, T. Unveiling some FDA-approved drugs as inhibitors of the store-operated Ca2+ entry pathway. Sci. Rep. 2017, 7, 12881. [Google Scholar] [CrossRef] [Green Version]
  51. Waldron, R.T.; Chen, Y.; Pham, H.; Go, A.; Su, H.Y.; Hu, C.; Wen, L.; Husain, S.Z.; Sugar, C.A.; Roos, J.; et al. The Orai Ca2+ channel inhibitor CM4620 targets both parenchymal and immune cells to reduce inflammation in experimental acute pancreatitis. J. Physiol. 2019, 597, 3085–3105. [Google Scholar] [CrossRef]
  52. Duncan, L.M.; Deeds, J.; Hunter, J.; Shao, J.; Holmgren, L.M.; Woolf, E.A.; Tepper, R.I.; Shyjan, A.W. Down-regulation of the novel gene melastatin correlates with potential for melanoma metastasis. Cancer Res. 1998, 58, 1515–1520. [Google Scholar] [PubMed]
  53. Koike, C.; Obara, T.; Uriu, Y.; Numata, T.; Sanuki, R.; Miyata, K.; Koyasu, T.; Ueno, S.; Funabiki, K.; Tani, A.; et al. TRPM1 is a component of the retinal ON bipolar cell transduction channel in the mGluR6 cascade. Proc. Natl. Acad. Sci. USA 2010, 107, 332–337. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  54. Irie, S.; Furukawa, T. TRPM1. Handb. Exp. Pharmacol. 2014, 222, 387–402. [Google Scholar] [CrossRef]
  55. Bellone, R.R.; Brooks, S.A.; Sandmeyer, L.; Murphy, B.A.; Forsyth, G.; Archer, S.; Bailey, E.; Grahn, B. Differential gene expression of TRPM1, the potential cause of congenital stationary night blindness and coat spotting patterns (LP) in the Appaloosa horse (Equus caballus). Genetics 2008, 179, 1861–1870. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  56. Oberwinkler, J.; Philipp, S.E. TRPM3. Handb. Exp. Pharmacol. 2014, 222, 427–459. [Google Scholar] [CrossRef] [PubMed]
  57. Grimm, C.; Kraft, R.; Sauerbruch, S.; Schultz, G.; Harteneck, C. Molecular and functional characterization of the melastatin-related cation channel TRPM3. J. Biol. Chem. 2003, 278, 21493–21501. [Google Scholar] [CrossRef] [Green Version]
  58. Nilius, B.; Owsianik, G.; Voets, T.; Peters, J.A. Transient receptor potential cation channels in disease. Physiol. Rev. 2007, 87, 165–217. [Google Scholar] [CrossRef] [Green Version]
  59. Shen, Y.; Rampino, M.A.; Carroll, R.C.; Nawy, S. G-protein-mediated inhibition of the Trp channel TRPM1 requires the Gβγ dimer. Proc. Natl. Acad. Sci. USA 2012, 109, 8752–8757. [Google Scholar] [CrossRef] [Green Version]
  60. Rampino, M.A.; Nawy, S.A. Relief of Mg²⁺-dependent inhibition of TRPM1 by PKCα at the rod bipolar cell synapse. J. Neurosci. 2011, 31, 13596–13603. [Google Scholar] [CrossRef] [Green Version]
  61. Schmidt, T.M. Role of melastatin-related transient receptor potential channel TRPM1 in the retina: Clues from horses and mice. J. Neurosci. 2009, 29, 11720–11722. [Google Scholar] [CrossRef]
  62. Held, K.; Aloi, V.D.; Freitas, A.C.N.; Janssens, A.; Segal, A.; Przibilla, J.; Philipp, S.E.; Wang, Y.T.; Voets, T.; Vriens, J. Pharmacological properties of TRPM3 isoforms are determined by the length of the pore loop. Br. J. Pharmacol. 2020, 1–16. [Google Scholar] [CrossRef] [PubMed]
  63. Straub, I.; Krügel, U.; Mohr, F.; Teichert, J.; Rizun, O.; Konrad, M.; Oberwinkler, J.; Schaefer, M. Flavanones that selectively inhibit TRPM3 attenuate thermal nociception in vivo. Mol. Pharmacol. 2013, 84, 736–750. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  64. Bouron, A.; Kiselyov, K.; Oberwinkler, J. Permeation, regulation and control of expression of TRP channels by trace metal ions. Pflugers Arch. 2015, 467, 1143–1164. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  65. Faouzi, M.; Penner, R. TRPM2. Handb. Exp. Pharmacol. 2014, 222, 403–426. [Google Scholar] [CrossRef]
  66. Robledo-Avila, F.H.; Ruiz-Rosado, J.D.; Brockman, K.L.; Partida-Sánchez, S. The TRPM2 Ion Channel Regulates Inflammatory Functions of Neutrophils During Listeria monocytogenes Infection. Front. Immunol. 2020, 11, 97. [Google Scholar] [CrossRef]
  67. Sano, Y.; Inamura, K.; Miyake, A.; Mochizuki, S.; Yokoi, H.; Matsushime, H.; Furuichi, K. Immunocyte Ca2+ influx system mediated by LTRPC2. Science 2001, 293, 1327–1330. [Google Scholar] [CrossRef]
  68. Turlova, E.; Feng, Z.P.; Sun, H.S. The role of TRPM2 channels in neurons, glial cells and the blood-brain barrier in cerebral ischemia and hypoxia. Acta Pharmacol. Sin. 2018, 39, 713–721. [Google Scholar] [CrossRef]
  69. Tóth, B.; Iordanov, I.; Csanády, L. Selective profiling of N- and C-terminal nucleotide-binding sites in a TRPM2 channel. J. Gen. Physiol. 2020, 152, e201912533. [Google Scholar] [CrossRef] [Green Version]
  70. Huang, Y.; Winkler, P.A.; Sun, W.; Lü, W.; Du, J. Architecture of the TRPM2 channel and its activation mechanism by ADP-ribose and calcium. Nature 2018, 562, 145–149. [Google Scholar] [CrossRef]
  71. Huang, Y.; Roth, B.; Lü, W.; Du, J. Ligand recognition and gating mechanism through three ligand-binding sites of human TRPM2 channel. eLife 2019, 8, e50175. [Google Scholar] [CrossRef]
  72. Iordanov, I.; Tóth, B.; Szollosi, A.; Csanády, L. Enzyme activity and selectivity filter stability of ancient TRPM2 channels were simultaneously lost in early vertebrates. eLife 2019, 8, e44556. [Google Scholar] [CrossRef] [PubMed]
  73. Liu, X.; Gong, B.; de Souza, L.B.; Ong, H.L.; Subedi, K.P.; Cheng, K.T.; Swaim, W.; Zheng, C.; Mori, Y.; Ambudkar, I.S. Radiation inhibits salivary gland function by promoting STIM1 cleavage by caspase-3 and loss of SOCE through a TRPM2-dependent pathway. Sci. Signal. 2017, 10, eaal4064. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  74. Ullrich, N.D.; Voets, T.; Prenen, J.; Vennekens, R.; Talavera, K.; Droogmans, G.; Nilius, B. Comparison of functional properties of the Ca2+-activated cation channels TRPM4 and TRPM5 from mice. Cell Calcium 2005, 37, 267–278. [Google Scholar] [CrossRef] [PubMed]
  75. Earley, S.; Brayden, J.E. Transient receptor potential channels in the vasculature. Physiol. Rev. 2015, 95, 645–690. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  76. Nilius, B.; Prenen, J.; Tang, J.; Wang, C.; Owsianik, G.; Janssens, A.; Voets, T.; Zhu, M.X. Regulation of the Ca2+ sensitivity of the nonselective cation channel TRPM4. J. Biol. Chem. 2005, 280, 6423–6433. [Google Scholar] [CrossRef] [Green Version]
  77. Bousova, K.; Jirku, M.; Bumba, L.; Bednarova, L.; Sulc, M.; Franek, M.; Vyklicky, L.; Vondrasek, J.; Teisinger, J. PIP2 and PIP3 interact with N-terminus region of TRPM4 channel. Biophys. Chem. 2015, 205, 24–32. [Google Scholar] [CrossRef]
  78. Prawitt, D.; Enklaar, T.; Klemm, G.; Gärtner, B.; Spangenberg, C.; Winterpacht, A.; Higgins, M.; Pelletier, J.; Zabel, B. Identification and characterization of MTR1, a novel gene with homology to melastatin (MLSN1) and the trp gene family located in the BWS-WT2 critical region on chromosome 11p15.5 and showing allele-specific expression. Hum. Mol. Genet. 2000, 9, 203–216. [Google Scholar] [CrossRef] [Green Version]
  79. Pérez, C.A.; Huang, L.; Rong, M.; Kozak, J.A.; Preuss, A.K.; Zhang, H.; Max, M.; Margolskee, R.F. A transient receptor potential channel expressed in taste receptor cells. Nat. Neurosci. 2002, 5, 1169–1176. [Google Scholar] [CrossRef]
  80. Launay, P.; Fleig, A.; Perraud, A.L.; Scharenberg, A.M.; Penner, R.; Kinet, J.P. TRPM4 is a Ca2+-activated nonselective cation channel mediating cell membrane depolarization. Cell 2002, 109, 397–407. [Google Scholar] [CrossRef] [Green Version]
  81. Xu, X.Z.; Moebius, F.; Gill, D.L.; Montell, C. Regulation of melastatin, a TRP-related protein, through interaction with a cytoplasmic isoform. Proc. Natl. Acad. Sci. USA 2001, 98, 10692–10697. [Google Scholar] [CrossRef] [Green Version]
  82. Earley, S. TRPM4 channels in smooth muscle function. Pflugers Arch. 2013, 465, 1223–1231. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  83. Autzen, H.E.; Myasnikov, A.G.; Campbell, M.G.; Asarnow, D.; Julius, D.; Cheng, Y. Structure of the human TRPM4 ion channel in a lipid nanodisc. Science 2018, 359, 228–232. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  84. Ruan, Z.; Haley, E.; Orozco, I.J.; Sabat, M.; Myers, R.; Roth, R.; Du, J.; Lü, W. Structures of the TRPM5 channel elucidate mechanisms of activation and inhibition. Nat. Struct. Mol. Biol. 2021, 28, 604–613. [Google Scholar] [CrossRef] [PubMed]
  85. Holzmann, C.; Kappel, S.; Kilch, T.; Jochum, M.M.; Urban, S.K.; Jung, V.; Stöckle, M.; Rother, K.; Greiner, M.; Peinelt, C. Transient receptor potential melastatin 4 channel contributes to migration of androgen-insensitive prostate cancer cells. Oncotarget 2015, 6, 41783–41793. [Google Scholar] [CrossRef]
  86. Kappel, S.; Stokłosa, P.; Hauert, B.; Ross-Kaschitza, D.; Borgström, A.; Baur, R.; Galván, J.A.; Zlobec, I.; Peinelt, C. TRPM4 is highly expressed in human colorectal tumor buds and contributes to proliferation, cell cycle, and invasion of colorectal cancer cells. Mol. Oncol. 2019, 13, 2393–2405. [Google Scholar] [CrossRef]
  87. Cáceres, M.; Ortiz, L.; Recabarren, T.; Romero, A.; Colombo, A.; Leiva-Salcedo, E.; Varela, D.; Rivas, J.; Silva, I.; Morales, D.; et al. TRPM4 Is a Novel Component of the Adhesome Required for Focal Adhesion Disassembly, Migration and Contractility. PLoS ONE 2015, 10, e0130540. [Google Scholar] [CrossRef]
  88. Earley, S.; Waldron, B.J.; Brayden, J.E. Critical role for transient receptor potential channel TRPM4 in myogenic constriction of cerebral arteries. Circ. Res. 2004, 95, 922–929. [Google Scholar] [CrossRef] [Green Version]
  89. Vennekens, R.; Olausson, J.; Meissner, M.; Bloch, W.; Mathar, I.; Philipp, S.E.; Schmitz, F.; Weissgerber, P.; Nilius, B.; Flockerzi, V.; et al. Increased IgE-dependent mast cell activation and anaphylactic responses in mice lacking the calcium-activated nonselective cation channel TRPM4. Nat. Immunol. 2007, 8, 312–320. [Google Scholar] [CrossRef] [Green Version]
  90. Launay, P.; Cheng, H.; Srivatsan, S.; Penner, R.; Fleig, A.; Kinet, J.P. TRPM4 regulates calcium oscillations after T cell activation. Science 2004, 306, 1374–1377. [Google Scholar] [CrossRef] [Green Version]
  91. Weber, K.S.; Hildner, K.; Murphy, K.M.; Allen, P.M. Trpm4 differentially regulates Th1 and Th2 function by altering calcium signaling and NFAT localization. J. Immunol. 2010, 185, 2836–2846. [Google Scholar] [CrossRef]
  92. Zhang, Z.; Yu, H.; Huang, J.; Faouzi, M.; Schmitz, C.; Penner, R.; Fleig, A. The TRPM6 kinase domain determines the Mg·ATP sensitivity of TRPM7/M6 heteromeric ion channels. J. Biol. Chem. 2014, 289, 5217–5227. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  93. Gees, M.; Colsoul, B.; Nilius, B. The role of transient receptor potential cation channels in Ca2+ signaling. Cold Spring Harb. Perspect. Biol. 2010, 2, a003962. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  94. Blanchard, M.G.; Kittikulsuth, W.; Nair, A.V.; de Baaij, J.H.; Latta, F.; Genzen, J.R.; Kohan, D.E.; Bindels, R.J.; Hoenderop, J.G. Regulation of Mg2+ Reabsorption and Transient Receptor Potential Melastatin Type 6 Activity by cAMP Signaling. J. Am. Soc. Nephrol. 2016, 27, 804–813. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  95. Chubanov, V.; Waldegger, S.; Mederos y Schnitzler, M.; Vitzthum, H.; Sassen, M.C.; Seyberth, H.W.; Konrad, M.; Gudermann, T. Disruption of TRPM6/TRPM7 complex formation by a mutation in the TRPM6 gene causes hypomagnesemia with secondary hypocalcemia. Proc. Natl. Acad. Sci. USA 2004, 101, 2894–2899. [Google Scholar] [CrossRef] [Green Version]
  96. Kaitsuka, T.; Katagiri, C.; Beesetty, P.; Nakamura, K.; Hourani, S.; Tomizawa, K.; Kozak, J.A.; Matsushita, M. Inactivation of TRPM7 kinase activity does not impair its channel function in mice. Sci. Rep. 2014, 4, 5718. [Google Scholar] [CrossRef]
  97. Ryazanova, L.V.; Rondon, L.J.; Zierler, S.; Hu, Z.; Galli, J.; Yamaguchi, T.P.; Mazur, A.; Fleig, A.; Ryazanov, A.G. TRPM7 is essential for Mg2+ homeostasis in mammals. Nat. Commun. 2010, 1, 109. [Google Scholar] [CrossRef] [Green Version]
  98. Schmitz, C.; Perraud, A.L.; Johnson, C.O.; Inabe, K.; Smith, M.K.; Penner, R.; Kurosaki, T.; Fleig, A.; Scharenberg, A.M. Regulation of vertebrate cellular Mg2+ homeostasis by TRPM7. Cell 2003, 114, 191–200. [Google Scholar] [CrossRef] [Green Version]
  99. He, Y.; Yao, G.; Savoia, C.; Touyz, R.M. Transient receptor potential melastatin 7 ion channels regulate magnesium homeostasis in vascular smooth muscle cells: Role of angiotensin II. Circ. Res. 2005, 96, 207–215. [Google Scholar] [CrossRef] [Green Version]
  100. Castiglioni, S.; Cazzaniga, A.; Trapani, V.; Cappadone, C.; Farruggia, G.; Merolle, L.; Wolf, F.I.; Iotti, S.; Maier, J.A.M. Magnesium homeostasis in colon carcinoma LoVo cells sensitive or resistant to doxorubicin. Sci. Rep. 2015, 5, 16538. [Google Scholar] [CrossRef]
  101. Stritt, S.; Nurden, P.; Favier, R.; Favier, M.; Ferioli, S.; Gotru, S.K.; van Eeuwijk, J.M.; Schulze, H.; Nurden, A.T.; Lambert, M.P.; et al. Defects in TRPM7 channel function deregulate thrombopoiesis through altered cellular Mg2+ homeostasis and cytoskeletal architecture. Nat. Commun. 2016, 7, 11097. [Google Scholar] [CrossRef]
  102. Abed, E.; Moreau, R. Importance of melastatin-like transient receptor potential 7 and cations (magnesium, calcium) in human osteoblast-like cell proliferation. Cell Prolif. 2007, 40, 849–865. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  103. Nadler, M.J.; Hermosura, M.C.; Inabe, K.; Perraud, A.L.; Zhu, Q.; Stokes, A.J.; Kurosaki, T.; Kinet, J.P.; Penner, R.; Scharenberg, A.M.; et al. LTRPC7 is a Mg.ATP-regulated divalent cation channel required for cell viability. Nature 2001, 411, 590–595. [Google Scholar] [CrossRef] [PubMed]
  104. Wei, C.; Wang, X.; Chen, M.; Ouyang, K.; Song, L.S.; Cheng, H. Calcium flickers steer cell migration. Nature 2009, 457, 901–905. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  105. Clark, K.; Langeslag, M.; van Leeuwen, B.; Ran, L.; Ryazanov, A.G.; Figdor, C.G.; Moolenaar, W.H.; Jalink, K.; van Leeuwen, F.N. TRPM7, a novel regulator of actomyosin contractility and cell adhesion. EMBO J. 2006, 25, 290–301. [Google Scholar] [CrossRef]
  106. Meng, X.; Cai, C.; Wu, J.; Cai, S.; Ye, C.; Chen, H.; Yang, Z.; Zeng, H.; Shen, Q.; Zou, F. TRPM7 mediates breast cancer cell migration and invasion through the MAPK pathway. Cancer Lett. 2013, 333, 96–102. [Google Scholar] [CrossRef]
  107. Su, L.T.; Liu, W.; Chen, H.C.; González-Pagán, O.; Habas, R.; Runnels, L.W. TRPM7 regulates polarized cell movements. Biochem. J. 2011, 434, 513–521. [Google Scholar] [CrossRef] [Green Version]
  108. Siddiqui, T.A.; Lively, S.; Vincent, C.; Schlichter, L.C. Regulation of podosome formation, microglial migration and invasion by Ca2+-signaling molecules expressed in podosomes. J. Neuroinflamm. 2012, 9, 250. [Google Scholar] [CrossRef] [Green Version]
  109. Su, L.T.; Agapito, M.A.; Li, M.; Simonson, W.T.; Huttenlocher, A.; Habas, R.; Yue, L.; Runnels, L.W. TRPM7 regulates cell adhesion by controlling the calcium-dependent protease calpain. J. Biol. Chem. 2006, 281, 11260–11270. [Google Scholar] [CrossRef] [Green Version]
  110. Liu, H.; Li, J.; Huang, Y.; Huang, C. Inhibition of transient receptor potential melastain 7 channel increases HSCs apoptosis induced by TRAIL. Life Sci. 2012, 90, 612–618. [Google Scholar] [CrossRef]
  111. Ng, N.M.; Jiang, S.P.; Lv, Z.Q. Retrovirus-mediated siRNA targeting TRPM7 gene induces apoptosis in RBL-2H3 cells. Eur. Rev. Med. Pharmacol. Sci. 2012, 16, 1172–1178. [Google Scholar]
  112. Zhang, Z.; Wang, M.; Fan, X.H.; Chen, J.H.; Guan, Y.Y.; Tang, Y.B. Upregulation of TRPM7 channels by angiotensin II triggers phenotypic switching of vascular smooth muscle cells of ascending aorta. Circ. Res. 2012, 111, 1137–1146. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  113. Abed, E.; Martineau, C.; Moreau, R. Role of melastatin transient receptor potential 7 channels in the osteoblastic differentiation of murine MC3T3 cells. Calcif. Tissue Int. 2011, 88, 246–253. [Google Scholar] [CrossRef] [PubMed]
  114. Numata, T.; Shimizu, T.; Okada, Y. TRPM7 is a stretch- and swelling-activated cation channel involved in volume regulation in human epithelial cells. Am. J. Physiol. Cell Physiol. 2007, 292, C460–C467. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  115. Runnels, L.W.; Yue, L.; Clapham, D.E. TRP-PLIK, a bifunctional protein with kinase and ion channel activities. Science 2001, 291, 1043–1047. [Google Scholar] [CrossRef] [PubMed]
  116. Sahni, J.; Scharenberg, A.M. TRPM7 ion channels are required for sustained phosphoinositide 3-kinase signaling in lymphocytes. Cell Metab. 2008, 8, 84–93. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  117. Bessac, B.F.; Fleig, A. TRPM7 channel is sensitive to osmotic gradients in human kidney cells. J. Physiol. 2007, 582, 1073–1086. [Google Scholar] [CrossRef]
  118. Zierler, S.; Yao, G.; Zhang, Z.; Kuo, W.C.; Pörzgen, P.; Penner, R.; Horgen, F.D.; Fleig, A. Waixenicin A inhibits cell proliferation through magnesium-dependent block of transient receptor potential melastatin 7 (TRPM7) channels. J. Biol. Chem. 2011, 286, 39328–39335. [Google Scholar] [CrossRef] [Green Version]
  119. Desai, B.N.; Krapivinsky, G.; Navarro, B.; Krapivinsky, L.; Carter, B.C.; Febvay, S.; Delling, M.; Penumaka, A.; Ramsey, I.S.; Manasian, Y.; et al. Cleavage of TRPM7 releases the kinase domain from the ion channel and regulates its participation in Fas-induced apoptosis. Dev. Cell 2012, 22, 1149–1162. [Google Scholar] [CrossRef] [Green Version]
  120. Hermosura, M.C.; Nayakanti, H.; Dorovkov, M.V.; Calderon, F.R.; Ryazanov, A.G.; Haymer, D.S.; Garruto, R.M. A TRPM7 variant shows altered sensitivity to magnesium that may contribute to the pathogenesis of two Guamanian neurodegenerative disorders. Proc. Natl. Acad. Sci. USA 2005, 102, 11510–11515. [Google Scholar] [CrossRef] [Green Version]
  121. Antunes, T.T.; Callera, G.E.; He, Y.; Yogi, A.; Ryazanov, A.G.; Ryazanova, L.V.; Zhai, A.; Stewart, D.J.; Shrier, A.; Touyz, R.M. Transient Receptor Potential Melastatin 7 Cation Channel Kinase: New Player in Angiotensin II-Induced Hypertension. Hypertension 2016, 67, 763–773. [Google Scholar] [CrossRef] [Green Version]
  122. Yogi, A.; Callera, G.E.; Antunes, T.T.; Tostes, R.C.; Touyz, R.M. Transient receptor potential melastatin 7 (TRPM7) cation channels, magnesium and the vascular system in hypertension. Circ. J. 2011, 75, 237–245. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  123. Suzuki, S.; Penner, R.; Fleig, A. TRPM7 contributes to progressive nephropathy. Sci. Rep. 2020, 10, 2333. [Google Scholar] [CrossRef] [Green Version]
  124. Sontia, B.; Montezano, A.C.; Paravicini, T.; Tabet, F.; Touyz, R.M. Downregulation of renal TRPM7 and increased inflammation and fibrosis in aldosterone-infused mice: Effects of magnesium. Hypertension 2008, 51, 915–921. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  125. Rios, F.J.; Zou, Z.G.; Harvey, A.P.; Harvey, K.Y.; Nosalski, R.; Anyfanti, P.; Camargo, L.L.; Lacchini, S.; Ryazanov, A.G.; Ryazanova, L.; et al. Chanzyme TRPM7 protects against cardiovascular inflammation and fibrosis. Cardiovasc. Res. 2020, 116, 721–735. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  126. Du, J.; Xie, J.; Zhang, Z.; Tsujikawa, H.; Fusco, D.; Silverman, D.; Liang, B.; Yue, L. TRPM7-mediated Ca2+ signals confer fibrogenesis in human atrial fibrillation. Circ. Res. 2010, 106, 992–1003. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  127. Romagnani, A.; Vettore, V.; Rezzonico-Jost, T.; Hampe, S.; Rottoli, E.; Nadolni, W.; Perotti, M.; Meier, M.A.; Hermanns, C.; Geiger, S.; et al. TRPM7 kinase activity is essential for T cell colonization and alloreactivity in the gut. Nat. Commun. 2017, 8, 1917. [Google Scholar] [CrossRef]
  128. Demeuse, P.; Penner, R.; Fleig, A. TRPM7 channel is regulated by magnesium nucleotides via its kinase domain. J. Gen. Physiol. 2006, 127, 421–434. [Google Scholar] [CrossRef] [Green Version]
  129. Kozak, J.A.; Matsushita, M.; Nairn, A.C.; Cahalan, M.D. Charge screening by internal pH and polyvalent cations as a mechanism for activation, inhibition, and rundown of TRPM7/MIC channels. J. Gen. Physiol. 2005, 126, 499–514. [Google Scholar] [CrossRef] [Green Version]
  130. Runnels, L.W.; Yue, L.; Clapham, D.E. The TRPM7 channel is inactivated by PIP(2) hydrolysis. Nat. Cell Biol. 2002, 4, 329–336. [Google Scholar] [CrossRef]
  131. Yu, H.; Zhang, Z.; Lis, A.; Penner, R.; Fleig, A. TRPM7 is regulated by halides through its kinase domain. Cell. Mol. Life Sci. 2013, 70, 2757–2771. [Google Scholar] [CrossRef] [Green Version]
  132. Jiang, J.; Li, M.; Yue, L. Potentiation of TRPM7 inward currents by protons. J. Gen. Physiol. 2005, 126, 137–150. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  133. Numata, T.; Okada, Y. Proton conductivity through the human TRPM7 channel and its molecular determinants. J. Biol. Chem. 2008, 283, 15097–15103. [Google Scholar] [CrossRef] [Green Version]
  134. Monteilh-Zoller, M.K.; Hermosura, M.C.; Nadler, M.J.; Scharenberg, A.M.; Penner, R.; Fleig, A. TRPM7 provides an ion channel mechanism for cellular entry of trace metal ions. J. Gen. Physiol. 2003, 121, 49–60. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  135. Hermosura, M.C.; Monteilh-Zoller, M.K.; Scharenberg, A.M.; Penner, R.; Fleig, A. Dissociation of the store-operated calcium current I(CRAC) and the Mg-nucleotide-regulated metal ion current MagNuM. J. Physiol. 2002, 539, 445–458. [Google Scholar] [CrossRef]
  136. Inoue, K.; Branigan, D.; Xiong, Z.G. Zinc-induced neurotoxicity mediated by transient receptor potential melastatin 7 channels. J. Biol. Chem. 2010, 285, 7430–7439. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  137. Doboszewska, U.; Młyniec, K.; Wlaź, A.; Poleszak, E.; Nowak, G.; Wlaź, P. Zinc signaling and epilepsy. Pharmacol. Ther. 2019, 193, 156–177. [Google Scholar] [CrossRef] [PubMed]
  138. Aarts, M.; Iihara, K.; Wei, W.L.; Xiong, Z.G.; Arundine, M.; Cerwinski, W.; MacDonald, J.F.; Tymianski, M. A key role for TRPM7 channels in anoxic neuronal death. Cell 2003, 115, 863–877. [Google Scholar] [CrossRef] [Green Version]
  139. Ryazanov, A.G.; Pavur, K.S.; Dorovkov, M.V. Alpha-kinases: A new class of protein kinases with a novel catalytic domain. Curr. Biol. 1999, 9, R43–R45. [Google Scholar] [CrossRef] [Green Version]
  140. Middelbeek, J.; Clark, K.; Venselaar, H.; Huynen, M.A.; van Leeuwen, F.N. The alpha-kinase family: An exceptional branch on the protein kinase tree. Cell. Mol. Life Sci. 2010, 67, 875–890. [Google Scholar] [CrossRef] [Green Version]
  141. Dorovkov, M.V.; Ryazanov, A.G. Phosphorylation of annexin I by TRPM7 channel-kinase. J. Biol. Chem. 2004, 279, 50643–50646. [Google Scholar] [CrossRef] [Green Version]
  142. Deason-Towne, F.; Perraud, A.L.; Schmitz, C. Identification of Ser/Thr phosphorylation sites in the C2-domain of phospholipase C γ2 (PLCγ2) using TRPM7-kinase. Cell. Signal. 2012, 24, 2070–2075. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  143. Clark, K.; Middelbeek, J.; Dorovkov, M.V.; Figdor, C.G.; Ryazanov, A.G.; Lasonder, E.; van Leeuwen, F.N. The alpha-kinases TRPM6 and TRPM7, but not eEF-2 kinase, phosphorylate the assembly domain of myosin IIA, IIB and IIC. FEBS Lett. 2008, 582, 2993–2997. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  144. Gotru, S.K.; Chen, W.; Kraft, P.; Becker, I.C.; Wolf, K.; Stritt, S.; Zierler, S.; Hermanns, H.M.; Rao, D.; Perraud, A.L.; et al. TRPM7 Kinase Controls Calcium Responses in Arterial Thrombosis and Stroke in Mice. Arterioscler. Thromb. Vasc. Biol. 2018, 38, 344–352. [Google Scholar] [CrossRef] [PubMed]
  145. Matsushita, M.; Kozak, J.A.; Shimizu, Y.; McLachlin, D.T.; Yamaguchi, H.; Wei, F.Y.; Tomizawa, K.; Matsui, H.; Chait, B.T.; Cahalan, M.D.; et al. Channel function is dissociated from the intrinsic kinase activity and autophosphorylation of TRPM7/ChaK1. J. Biol. Chem. 2005, 280, 20793–20803. [Google Scholar] [CrossRef] [Green Version]
  146. Ryazanova, L.V.; Dorovkov, M.V.; Ansari, A.; Ryazanov, A.G. Characterization of the protein kinase activity of TRPM7/ChaK1, a protein kinase fused to the transient receptor potential ion channel. J. Biol. Chem. 2004, 279, 3708–3716. [Google Scholar] [CrossRef] [Green Version]
  147. Clark, K.; Middelbeek, J.; Morrice, N.A.; Figdor, C.G.; Lasonder, E.; van Leeuwen, F.N. Massive autophosphorylation of the Ser/Thr-rich domain controls protein kinase activity of TRPM6 and TRPM7. PLoS ONE 2008, 3, e1876. [Google Scholar] [CrossRef] [Green Version]
  148. Krapivinsky, G.; Krapivinsky, L.; Manasian, Y.; Clapham, D.E. The TRPM7 chanzyme is cleaved to release a chromatin-modifying kinase. Cell 2014, 157, 1061–1072. [Google Scholar] [CrossRef] [Green Version]
  149. Asagiri, M.; Sato, K.; Usami, T.; Ochi, S.; Nishina, H.; Yoshida, H.; Morita, I.; Wagner, E.F.; Mak, T.W.; Serfling, E.; et al. Autoamplification of NFATc1 expression determines its essential role in bone homeostasis. J. Exp. Med. 2005, 202, 1261–1269. [Google Scholar] [CrossRef] [Green Version]
  150. Caterina, M.J.; Rosen, T.A.; Tominaga, M.; Brake, A.J.; Julius, D. A capsaicin-receptor homologue with a high threshold for noxious heat. Nature 1999, 398, 436–441. [Google Scholar] [CrossRef]
  151. Caterina, M.J.; Leffler, A.; Malmberg, A.B.; Martin, W.J.; Trafton, J.; Petersen-Zeitz, K.R.; Koltzenburg, M.; Basbaum, A.I.; Julius, D. Impaired nociception and pain sensation in mice lacking the capsaicin receptor. Science 2000, 288, 306–313. [Google Scholar] [CrossRef]
  152. McKemy, D.D.; Neuhausser, W.M.; Julius, D. Identification of a cold receptor reveals a general role for TRP channels in thermosensation. Nature 2002, 416, 52–58. [Google Scholar] [CrossRef] [PubMed]
  153. Diver, M.M.; Cheng, Y.; Julius, D. Structural insights into TRPM8 inhibition and desensitization. Science 2019, 365, 1434–1440. [Google Scholar] [CrossRef] [PubMed]
  154. Coste, B.; Xiao, B.; Santos, J.S.; Syeda, R.; Grandl, J.; Spencer, K.S.; Kim, S.E.; Schmidt, M.; Mathur, J.; Dubin, A.E.; et al. Piezo proteins are pore-forming subunits of mechanically activated channels. Nature 2012, 483, 176–181. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  155. Bautista, D.M.; Siemens, J.; Glazer, J.M.; Tsuruda, P.R.; Basbaum, A.I.; Stucky, C.L.; Jordt, S.E.; Julius, D. The menthol receptor TRPM8 is the principal detector of environmental cold. Nature 2007, 448, 204–208. [Google Scholar] [CrossRef] [PubMed]
  156. Kühn, F.J.; Knop, G.; Lückhoff, A. The transmembrane segment S6 determines cation versus anion selectivity of TRPM2 and TRPM8. J. Biol. Chem. 2007, 282, 27598–27609. [Google Scholar] [CrossRef] [Green Version]
  157. Brauchi, S.; Orta, G.; Salazar, M.; Rosenmann, E.; Latorre, R. A hot-sensing cold receptor: C-terminal domain determines thermosensation in transient receptor potential channels. J. Neurosci. 2006, 26, 4835–4840. [Google Scholar] [CrossRef] [Green Version]
  158. Brauchi, S.; Orta, G.; Mascayano, C.; Salazar, M.; Raddatz, N.; Urbina, H.; Rosenmann, E.; Gonzalez-Nilo, F.; Latorre, R. Dissection of the components for PIP2 activation and thermosensation in TRP channels. Proc. Natl. Acad. Sci. USA 2007, 104, 10246–10251. [Google Scholar] [CrossRef] [Green Version]
  159. González-Muñiz, R.; Bonache, M.A.; Martín-Escura, C.; Gómez-Monterrey, I. Recent Progress in TRPM8 Modulation: An Update. Int. J. Mol. Sci. 2019, 20, 2618. [Google Scholar] [CrossRef] [Green Version]
  160. Yin, Y.; Lee, S.Y. Current View of Ligand and Lipid Recognition by the Menthol Receptor TRPM8. Trends Biochem. Sci. 2020, 45, 806–819. [Google Scholar] [CrossRef]
  161. Yin, Y.; Wu, M.; Zubcevic, L.; Borschel, W.F.; Lander, G.C.; Lee, S.Y. Structure of the cold- and menthol-sensing ion channel TRPM8. Science 2018, 359, 237–241. [Google Scholar] [CrossRef] [Green Version]
  162. Chuang, H.H.; Neuhausser, W.M.; Julius, D. The super-cooling agent icilin reveals a mechanism of coincidence detection by a temperature-sensitive TRP channel. Neuron 2004, 43, 859–869. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  163. Prescott, E.D.; Julius, D. A modular PIP2 binding site as a determinant of capsaicin receptor sensitivity. Science 2003, 300, 1284–1288. [Google Scholar] [CrossRef]
  164. Daniels, R.L.; Takashima, Y.; McKemy, D.D. Activity of the neuronal cold sensor TRPM8 is regulated by phospholipase C via the phospholipid phosphoinositol 4,5-bisphosphate. J. Biol. Chem. 2009, 284, 1570–1582. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  165. Liu, Y.; Mikrani, R.; He, Y.; Faran Ashraf Baig, M.M.; Abbas, M.; Naveed, M.; Tang, M.; Zhang, Q.; Li, C.; Zhou, X. TRPM8 channels: A review of distribution and clinical role. Eur. J. Pharmacol. 2020, 882, 173312. [Google Scholar] [CrossRef] [PubMed]
  166. Langeslag, M.; Clark, K.; Moolenaar, W.H.; van Leeuwen, F.N.; Jalink, K. Activation of TRPM7 channels by phospholipase C-coupled receptor agonists. J. Biol. Chem. 2007, 282, 232–239. [Google Scholar] [CrossRef] [Green Version]
  167. Guilbert, A.; Gautier, M.; Dhennin-Duthille, I.; Haren, N.; Sevestre, H.; Ouadid-Ahidouch, H. Evidence that TRPM7 is required for breast cancer cell proliferation. Am. J. Physiol. Cell Physiol. 2009, 297, C493–C502. [Google Scholar] [CrossRef]
  168. Yang, Y.M.; Jung, H.H.; Lee, S.J.; Choi, H.J.; Kim, M.S.; Shin, D.M. TRPM7 Is Essential for RANKL-Induced Osteoclastogenesis. Korean J. Physiol. Pharmacol. 2013, 17, 65–71. [Google Scholar] [CrossRef] [Green Version]
  169. Zhao, Y.; McVeigh, B.M.; Moiseenkova-Bell, V.Y. Structural Pharmacology of TRP Channels. J. Mol. Biol. 2021, 433, 166914. [Google Scholar] [CrossRef]
  170. Vaeth, M.; Yang, J.; Yamashita, M.; Zee, I.; Eckstein, M.; Knosp, C.; Kaufmann, U.; Karoly Jani, P.; Lacruz, R.S.; Flockerzi, V.; et al. ORAI2 modulates store-operated calcium entry and T cell-mediated immunity. Nat. Commun. 2017, 8, 14714. [Google Scholar] [CrossRef] [Green Version]
  171. Chubanov, V.; Ferioli, S.; Gudermann, T. Assessment of TRPM7 functions by drug-like small molecules. Cell Calcium 2017, 67, 166–173. [Google Scholar] [CrossRef]
  172. Beesetty, P.; Wieczerzak, K.B.; Gibson, J.N.; Kaitsuka, T.; Luu, C.T.; Matsushita, M.; Kozak, J.A. Inactivation of TRPM7 kinase in mice results in enlarged spleens, reduced T-cell proliferation and diminished store-operated calcium entry. Sci. Rep. 2018, 8, 3023. [Google Scholar] [CrossRef] [PubMed]
  173. Vaeth, M.; Maus, M.; Klein-Hessling, S.; Freinkman, E.; Yang, J.; Eckstein, M.; Cameron, S.; Turvey, S.E.; Serfling, E.; Berberich-Siebelt, F.; et al. Store-Operated Ca2+ Entry Controls Clonal Expansion of T Cells through Metabolic Reprogramming. Immunity 2017, 47, 664–679.e6. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  174. Kircher, S.; Merino-Wong, M.; Niemeyer, B.A.; Alansary, D. Profiling calcium signals of in vitro polarized human effector CD4(+) T cells. Biochim. Biophys. Acta Mol. Cell Res. 2018, 1865, 932–943. [Google Scholar] [CrossRef] [PubMed]
  175. Angenendt, A.; Steiner, R.; Knörck, A.; Schwär, G.; Konrad, M.; Krause, E.; Lis, A. Orai, STIM, and PMCA contribute to reduced calcium signal generation in CD8(+) T cells of elderly mice. Aging 2020, 12, 3266–3286. [Google Scholar] [CrossRef] [PubMed]
  176. Smyth, J.T.; Petranka, J.G.; Boyles, R.R.; DeHaven, W.I.; Fukushima, M.; Johnson, K.L.; Williams, J.G.; Putney, J.W., Jr. Phosphorylation of STIM1 underlies suppression of store-operated calcium entry during mitosis. Nat. Cell. Biol. 2009, 11, 1465–1472. [Google Scholar] [CrossRef] [PubMed]
  177. Kawasaki, T.; Ueyama, T.; Lange, I.; Feske, S.; Saito, N. Protein kinase C-induced phosphorylation of Orai1 regulates the intracellular Ca2+ level via the store-operated Ca2+ channel. J. Biol. Chem. 2010, 285, 25720–25730. [Google Scholar] [CrossRef] [Green Version]
  178. Eylenstein, A.; Gehring, E.M.; Heise, N.; Shumilina, E.; Schmidt, S.; Szteyn, K.; Münzer, P.; Nurbaeva, M.K.; Eichenmüller, M.; Tyan, L.; et al. Stimulation of Ca2+-channel Orai1/STIM1 by serum- and glucocorticoid-inducible kinase 1 (SGK1). FASEB J. 2011, 25, 2012–2021. [Google Scholar] [CrossRef] [Green Version]
  179. Pozo-Guisado, E.; Casas-Rua, V.; Tomas-Martin, P.; Lopez-Guerrero, A.M.; Alvarez-Barrientos, A.; Martin-Romero, F.J. Phosphorylation of STIM1 at ERK1/2 target sites regulates interaction with the microtubule plus-end binding protein EB1. J. Cell Sci. 2013, 126, 3170–3180. [Google Scholar] [CrossRef] [Green Version]
  180. Srikanth, S.; Jung, H.J.; Kim, K.D.; Souda, P.; Whitelegge, J.; Gwack, Y. A novel EF-hand protein, CRACR2A, is a cytosolic Ca2+ sensor that stabilizes CRAC channels in T cells. Nat. Cell Biol. 2010, 12, 436–446. [Google Scholar] [CrossRef] [Green Version]
  181. Palty, R.; Raveh, A.; Kaminsky, I.; Meller, R.; Reuveny, E. SARAF inactivates the store operated calcium entry machinery to prevent excess calcium refilling. Cell 2012, 149, 425–438. [Google Scholar] [CrossRef] [Green Version]
  182. Walsh, C.M.; Doherty, M.K.; Tepikin, A.V.; Burgoyne, R.D. Evidence for an interaction between Golli and STIM1 in store-operated calcium entry. Biochem. J. 2010, 430, 453–460. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  183. Sharma, S.; Quintana, A.; Findlay, G.M.; Mettlen, M.; Baust, B.; Jain, M.; Nilsson, R.; Rao, A.; Hogan, P.G. An siRNA screen for NFAT activation identifies septins as coordinators of store-operated Ca2+ entry. Nature 2013, 499, 238–242. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  184. Chen, Y.T.; Chen, Y.F.; Chiu, W.T.; Wang, Y.K.; Chang, H.C.; Shen, M.R. The ER Ca²⁺ sensor STIM1 regulates actomyosin contractility of migratory cells. J. Cell Sci. 2013, 126, 1260–1267. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  185. Mu, Y.P.; Lin, D.C.; Zheng, S.Y.; Jiao, H.X.; Sham, J.S.K.; Lin, M.J. Transient Receptor Potential Melastatin-8 Activation Induces Relaxation of Pulmonary Artery by Inhibition of Store-Operated Calcium Entry in Normoxic and Chronic Hypoxic Pulmonary Hypertensive Rats. J. Pharmacol. Exp. Ther. 2018, 365, 544–555. [Google Scholar] [CrossRef] [Green Version]
  186. Zhang, X. Direct Gα(q) Gating Is the Sole Mechanism for TRPM8 Inhibition Caused by Bradykinin Receptor Activation. Cell Rep. 2019, 27, 3672–3683.e4. [Google Scholar] [CrossRef] [Green Version]
Figure 1. Structural architecture of TRPM7. TRPM7 channels are formed by six helical transmembrane domains (TM1-TM6). The channel pore (P) of TRPM7 is located between TM5 and TM6. TM1 contains the N-terminal region and TM6 harbors the serine/threonine kinase domain. The N’ and C’- terminal regions are in the cytosol.
Figure 1. Structural architecture of TRPM7. TRPM7 channels are formed by six helical transmembrane domains (TM1-TM6). The channel pore (P) of TRPM7 is located between TM5 and TM6. TM1 contains the N-terminal region and TM6 harbors the serine/threonine kinase domain. The N’ and C’- terminal regions are in the cytosol.
Cells 11 01190 g001
Figure 2. TRPM7 stimulation does not elicit Ca2+ influx in SOCE-deficient ameloblasts. (A) Representative original traces of [Ca2+]cyt transients in ameloblasts of Stim1/2K14cre mice (Stim1/2cKO) and controls (WT) ameloblasts. The ameloblasts of Stim1/2K14cre mice were isolated as reported (28). SOCE was measured after pre-incubation with thapsigargin (20 min, 1 μM), followed by perfusion with a Ca2+-free Ringer’s solution (60 s) before simultaneous re-addition of 2.0 mM extracellular Ca2+ or with 2 mM Ca2+ and the TRPM7 agonist naltriben (NAL,100 μM). (B) Quantification of the SOCE peak. Data were analyzed by one-way ANOVA followed by Tukey’s multiple comparison post-hoc test. * p< 0.05, *** p < 0.001, n.s., non-significant.
Figure 2. TRPM7 stimulation does not elicit Ca2+ influx in SOCE-deficient ameloblasts. (A) Representative original traces of [Ca2+]cyt transients in ameloblasts of Stim1/2K14cre mice (Stim1/2cKO) and controls (WT) ameloblasts. The ameloblasts of Stim1/2K14cre mice were isolated as reported (28). SOCE was measured after pre-incubation with thapsigargin (20 min, 1 μM), followed by perfusion with a Ca2+-free Ringer’s solution (60 s) before simultaneous re-addition of 2.0 mM extracellular Ca2+ or with 2 mM Ca2+ and the TRPM7 agonist naltriben (NAL,100 μM). (B) Quantification of the SOCE peak. Data were analyzed by one-way ANOVA followed by Tukey’s multiple comparison post-hoc test. * p< 0.05, *** p < 0.001, n.s., non-significant.
Cells 11 01190 g002
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Souza Bomfim, G.H.; Niemeyer, B.A.; Lacruz, R.S.; Lis, A. On the Connections between TRPM Channels and SOCE. Cells 2022, 11, 1190. https://doi.org/10.3390/cells11071190

AMA Style

Souza Bomfim GH, Niemeyer BA, Lacruz RS, Lis A. On the Connections between TRPM Channels and SOCE. Cells. 2022; 11(7):1190. https://doi.org/10.3390/cells11071190

Chicago/Turabian Style

Souza Bomfim, Guilherme H., Barbara A. Niemeyer, Rodrigo S. Lacruz, and Annette Lis. 2022. "On the Connections between TRPM Channels and SOCE" Cells 11, no. 7: 1190. https://doi.org/10.3390/cells11071190

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