International Union of Basic and Clinical Pharmacology XXX: Calcium-sensing receptor Nomenclature, Pharmacology, and Function
. Author manuscript; available in PMC: 2021 Jan 1.
Published in final edited form as: Pharmacol Rev. 2020 Jul 1;72(3):558–604. doi: 10.1124/pr.119.018531
Abstract
The calcium-sensing receptor (CaSR) is a class C G protein-coupled receptor (GPCR) that responds to multiple endogenous agonists and allosteric modulators, including di- and tri-valent cations, L-amino acids, γ-glutamyl peptides, polyamines, polycationic peptides and protons. The CaSR plays a critical role in extracellular calcium (Ca2+ o) homeostasis, demonstrated by the many naturally occurring mutations in the CaSR or its signaling partners that cause Ca2+ o homeostasis disorders. However, CaSR tissue expression in mammals is broad and includes tissues unrelated to Ca2+ o homeostasis, where it e.g. regulates the secretion of digestive hormones, airway constriction, cardiovascular effects, cellular differentiation and proliferation. Thus, while the CaSR is targeted clinically by the positive allosteric modulators (PAMs), cinacalcet, evocalcet and etelcalcetide in hyperparathyroidism, it is also a putative therapeutic target in diabetes, asthma, cardiovascular disease and cancer. The CaSR is somewhat unique in possessing multiple ligand binding sites, including at least five putative sites for the “orthosteric” agonist Ca2+ o, an allosteric site for endogenous L-amino acids, two further allosteric sites for small molecules and the peptide PAM, etelcalcetide, and additional sites for other cations and anions. The CaSR is promiscuous in its G protein coupling preferences, and signals via Gq/11, Gi/o, potentially G12/13 and even Gs in some cell types. Not surprisingly, the CaSR is subject to biased agonism, where distinct ligands preferentially stimulate a subset of the CaSR’s possible signaling responses, to the exclusion of others. The CaSR thus serves as a model receptor to study natural bias and allostery.
I. Introduction
A. Identification and cloning of the calcium-sensing receptor (CaSR)
Ca2+ is an essential ion, both intra- and extracellularly in mammals. Intracellular Ca2+ (Ca2+ i) is maintained at approximately 100 nM, but rises to low micromolar concentrations upon membrane or endoplasmic reticulum Ca2+ channel opening, thus serving as an important second messenger (Brini et al., 2013). Ca2+ also functions as a key first messenger via activation of the CaSR (Alexander, et al., 2017; Bikle et al., 2019), which plays a pivotal role in tightly regulating ionized (free) extracellular calcium (Ca2+ o). In human plasma, total calcium (referred to herein as calcium to signify ionized and non-ionized calcium) levels are maintained between 2.1 and 2.6 mM of which roughly half is in an ionized form (Brini et al., 2013).
In the mid 1980s there was significant interest in the mechanisms regulating parathyroid hormone (PTH) release from the parathyroid glands. It was consequently shown that elevated Ca2+ o increased Ca2+ i levels and decreased PTH release (LeBoff et al., 1985; Nemeth et al., 1986). In the following years, elevated Ca2+ o was demonstrated to increase inositol phosphate (IP) and decrease cAMP levels, which led to the suggestion of a cell surface calcium-sensing G protein-coupled receptor (GPCR) (Brown et al., 1987a; Chen et al., 1989; Nemeth and Scarpa, 1986; 1987). Further evidence for the receptor was provided via activation of Ca2+-sensitive Cl- channels in Xenopus oocytes injected with mRNA isolated from bovine parathyroid cells (Racke et al., 1993), which subsequently led to expression cloning of the bovine CaSR (Brown et al., 1993). In isolated parathyroid cells, the cloned bovine CaSR was activated by gadolinium > neomycin > Ca2+ o > Mg2+ (in rank order of potency) and signaled through elevation of Ca2+ i, providing strong evidence of the cloned receptor being the long sought CaSR (Brown et al., 1993).
Analyses of the cloned receptor sequence revealed a 1,085 amino acid long protein consisting of a large amino-terminal extracellular domain (ECD) of 613 amino acids comprised of a “venus flytrap” (VFT) domain, which closes upon activation much like the VFT plant, and cysteine-rich domain, a 7-transmembrane (7TM) domain of 250 amino acids and an intracellular carboxy-terminus of 222 amino acids (Brown et al., 1993). The analyses also revealed that the CaSR was homologous to the metabotropic glutamate receptors, which were later shown to form the class C GPCRs together with GABAB, taste type 1, GPRC6A and a handful of orphan receptors (Wellendorph and Bräuner-Osborne, 2009). The structurally conserved class C GPCR VFT domain is homologous to bacterial periplasmic binding proteins, thus it has been predicted that class C GPCRs arose from fusion of the GPCR 7TM with a periplasmic binding protein (O'Hara et al., 1993). Nucleic acid hybridization techniques quickly led to cloning of the human (Garrett et al., 1995a), rat (Riccardi et al., 1995; Ruat et al., 1995), rabbit (Butters et al., 1997), chicken (Diaz et al., 1997) and shark (Nearing et al., 2002) CaSR orthologs, and genome database mining subsequently suggested that the CaSR is evolutionarily conserved to flies and worms (Bjarnadóttir et al., 2005).
B. General gene structure
The human CASR gene has been mapped to chromosome 3q13.3-21 by fluorescence in situ hybridization (FISH) (Janicic et al., 1995) and linkage analyses (Chou et al., 1992). The human CaSR is encoded by 7 exons, of which exons 2-6 encode the ECD, and exon 7 encodes the 7TM and intracellular carboxy terminus (Pearce et al., 1995; Pollak et al., 1993). Two different 5'-untranslated promotor regions, termed exon 1A and 1B, have been identified in humans (Chikatsu et al., 2000), which both splice with the same site in exon 2. As recently reviewed (Hendy and Canaff, 2016), the promotors and thus CaSR expression are regulated by cis-elements responding to 1,25-dihydroxyvitamin D (1,25(OH)2D), proinflammatory cytokines, and the transcription factor glial cells missing-2 (GCM2).
Tissue specific splice variants lacking exon 3 (Bradbury et al., 1998) and exon 5 (Oda et al., 1998) have been reported but their function (if any) remains elusive. The exon 5 splice variant is of particular interest as it is functional in growth plate chondrocytes (Rodriguez et al., 2005) despite being non-functional when recombinantly expressed in HEK293 and CHO cells. These latter findings led to an initial underestimation of the role of the CaSR in bone development, as the original exon 5 knockout mouse (Ho et al., 1995) displayed a mild bone phenotype compared to a more severe phenotype in the exon 7 knockout mouse model (Chang et al., 2008).
C. Tissue distribution
mRNA probes and antibodies have revealed that the CaSR is widely expressed both in tissues directly involved in controlling systemic Ca2+ o homeostasis as well as in tissues with other functions. As detailed in Section V, the plasma calcium level is mainly regulated via actions on the parathyroid gland (PTH release), thyroid gland (calcitonin release, although calcitonin in humans is less important than in rodents) and the kidney (production of vitamin 1,25(OH)2D3 and regulation of ion excretion), but other tissues such as the bone (release of skeletal Ca2+ o) and small intestine (Ca2+ o absorption) also play a role both via direct CaSR activation and via the PTH, calcitonin and vitamin 1,25(OH)2D3 hormones (Brown, 2013; Brown and MacLeod, 2001; Lee et al., 2019). In addition, the CaSR is expressed in a range of tissues not involved in systemic Ca2+ o homeostasis such as the keratinocytes of the skin (Section VE), mammary glands (Section VF), pancreas (Section VG), colon (Section VH), airway smooth muscle and epithelium (Section VI), vascular smooth muscle and endothelium (Section VJ) and the brain (Section VK) where the CaSR regulates a range of (patho)physiological functions.
D. Signal transduction pathways
Some of the key CaSR signaling pathways are shown in Figure 1. The CaSR primarily elicits its functions by coupling to the Gi/o and Gq/11 families of heterotrimeric G proteins to activate intracellular signaling pathways that inhibit PTH synthesis and release from parathyroid cells (Section VA). CaSR activation of Gi/o proteins leads to inhibition of the cAMP synthesizing enzyme, adenylate cyclase, causing a decrease in intracellular cAMP levels (Chang et al., 1998; Kifor et al., 2001). CaSR coupling to Gq/11 is usually considered the primary signaling pathway, which activates phospholipase C (PLC)-β to hydrolyze phosphatidylinositol 4,5-bisphosphate to the second messengers IP3 and diacylglycerol (DAG) (Brown et al., 1993; Chang et al., 1998). IP3 triggers release of Ca2+ i from intracellular stores, such as the endoplasmic reticulum, and DAG alone or in combination with Ca2+ i activates protein kinase C (PKC). Cytosolic phospholipase A2, which is the rate-limiting enzyme in arachidonic acid metabolism, can also be activated by the CaSR-mediated Gq/11 pathway through calmodulin and the Ca2+/calmodulin-dependent protein kinase II (Handlogten et al., 2001).
Figure 1. Key CaSR signaling pathways.
The CaSR primarily couples to Gq/11 and Gi/o proteins to mediate many of its physiological responses including PTH release. The CaSR may also couple to G12/13 but the physiological relevance of this is unknown, therefore G12/13 is semi-transparent in the figure. CaM, calmodulin; EGFR, epidermal growth factor receptor; ER, endoplasmic reticulum; PI3K, phosphatidylinositol 3-kinase; PI4K, phosphatidylinositol 4-kinase; all other abbreviations are as described in the text.
The importance of the Gq/11 pathway in CaSR physiology has been demonstrated by the similarities between selective parathyroid knockout of the genes encoding Gα q (Gnaq) and Gα 11 (Gna11) in mice, which results in a phenotype with almost all the features of Casr germline knockout mice (Wettschureck et al., 2007). Similarly, human CASR and GNA11 loss- or gain-of-function mutations cause familial hypocalciuric hypercalcemia (FHH) types 1 (CASR) and 2 (GNA11) or autosomal dominant hypocalcemia (ADH) types 1 (CASR) or 2 (GNA11), respectively (Nesbit et al., 2013a; Pollak et al., 1993; Pollak et al., 1994) (Section VI).
Studies of CaSR coupling to G12/13 are limited due to a lack of inhibitors and suitable functional readouts. However, the CaSR activates phospholipase D (PLD) in Madin-Darby canine kidney cells through a Gq/11- and Gi/o-independent pathway involving activation of the Rho family of small GTPases, most likely via G12/13 coupling (Huang et al., 2004). The G12/13 pathway is also likely to be the Gq/11- and Gi/o-independent pathway that activates the phosphatidylinositol 4-kinase responsible for the first step in inositol biosynthesis through Rho (Huang et al., 2002). However, CaSR can activate RhoA by a Gq/11 pathway in HEK293 cells (Pi et al., 2002) and PLD by a PKC-dependent mechanism likely mediated by Gq/11 in HEK293 cells and parathyroid cells (Kifor et al., 1997), so it remains unclear whether CaSR also couples to G12/13 in these cells.
CaSR coupling to Gs and the consequent increase in intracellular cAMP levels activates PKA and stimulates PTH related protein (PTHrP) release in immortalized and malignant breast cells and in the AtT-20 pituitary-derived cell line (Mamillapalli et al., 2008; Mamillapalli and Wysolmerski, 2010) (Section VH). Stimulation of cAMP production is not observed in HEK293 cells recombinantly expressing the CaSR (Thomsen et al., 2012a) and the molecular mechanism for the switch in G protein preference in breast cancer and AtT-20 cells remains unknown.
The CaSR activates several mitogen-activated protein kinase (MAPK) cascades, including extracellular signal-regulated kinase 1/2 (ERK1/2), p38 MAPK and c-Jun N-terminal kinase (JNK) to regulate PTHrP release, proliferation and other functions (Chattopadhyay et al., 2004; MacLeod et al., 2003; Tfelt-Hansen et al., 2003). ERK1/2 is activated by phosphorylation (pERK1/2) through multiple CaSR-mediated pathways, including parallel G protein-dependent pathways involving either Gq/11 and PKC, or Gi/o and epidermal growth factor receptor transactivation (Kifor et al., 2001; MacLeod et al., 2004; Thomsen et al., 2012a). Ras and phosphatidylinositol 3-kinase are also involved in ERK1/2 activation by the CaSR (Hobson et al., 2003), but it is unclear if this pathway overlaps with the Gq/11 or Gi/o dependent pathways. The CaSR can also activate ERK1/2 through a β-arrestin-dependent and G protein-independent pathway (Thomsen et al., 2012a). Furthermore, an Arg6803.32Gly (numbering shown in superscript after residue numbers throughout this manuscript is based on Ballesteros-Weinstein numbering assigned in (Ballesteros and Weinstein, 1995) for class A GPCRs and in (Dore et al., 2014) for class C GPCRs) CaSR mutation associated with ADH1 selectively increases β-arrestin dependent ERK1/2 activation, where the mutation is predicted to disrupt an extracellular salt bridge between Arg6803.32 and Glu767 in the second extracellular loop (ECL) (Gorvin et al., 2018a).
In some cell types, the CaSR stimulates opening of L-type voltage-gated Ca2+ channels (Fajtova et al., 1991; McGehee et al., 1997; Muff et al., 1988), and nonselective cation channels including transient receptor potential cation (TRPC) channels (El Hiani et al., 2006; Meng et al., 2014; Ye et al., 1996), although the pathways that couple the CaSR to ion channels are poorly defined.
II. Agonists and allosteric modulators
A. Endogenous and exogenous agonists
Polyvalent cations
The CaSR is now well-known for its ability to sense fluctuations in Ca2+ o. CaSR radioligand binding assays to quantify the affinity of Ca2+ o and other agonists have to date not been possible due to low agonist affinity, a lack of suitable radioligands, and complexities in quantifying agonist binding to multiple binding sites. However, spectroscopic studies indicate Ca2+ o binds to the VFT with an affinity in the range of 3.0 – 5.0 mM (Zhang et al., 2014b). These findings are supported by the use of an operational model of agonism for receptors with multiple agonist binding sites, where Ca2+ o affinity at the full-length CaSR was 1.1 – 1.3 mM (Gregory et al., 2020). The low millimolar Ca2+ o affinity is consistent with Ca2+ o potency in healthy human subjects, where Ca2+ o suppresses PTH secretion with an approximate IC50 of 1.2 mM (which is also the approximate free Ca2+ o concentration in human serum) (Brown, 1991; Ramirez et al., 1993), while in cultured parathyroid cells, the Ca2+ o IC50 for PTH release is closer to 1 mM (Brown, 1983; 1991). The Ca2+ o-PTH relationship is characterized by a Hill coefficient greater than unity (Brown, 1983; 1991; Ramirez et al., 1993). This is because multiple Ca2+ o ions bind to the CaSR in a positively cooperative manner, allowing the CaSR to respond to minute changes in Ca2+ o concentrations that span less than 100 μM (Brown, 1983; 1991; Ramirez et al., 1993). Thus, although Ca2+ o is considered the primary endogenous and therefore orthosteric agonist of the CaSR, strictly speaking it is an allosteric modulator of its own activity.
In addition to Ca2+ o, the CaSR is activated by many other polyvalent cations, including magnesium, zinc, manganese, ferrous iron, strontium, barium, cadmium, cobalt, nickel, lead, terbium, gadolinium, europium and yttrium (Brown et al., 1990; Handlogten et al., 2000; Ruat et al., 1996). Trivalent cations are generally more potent than divalent cations, of which Ca2+ o and magnesium are the most physiologically relevant. The role of non-Ca2+ o cations in CaSR-mediated (patho)physiology is unknown. Agonists that mimic the actions of Ca2+ o at the CaSR have traditionally been called type I calcimimetics.
Although much larger and structurally more complex than the small cations described above, polyamines are CaSR agonists. Polyamines are found in all eukaryotes, with spermine, spermidine and their diamine precursor putrescine the most abundant in mammals. Polyamines are synthesized ubiquitously in the body, and are also ingested in the diet and secreted by intestinal bacteria. Although polyamines activate the CaSR in the absence of Ca2+ o, there is some evidence they also potentiate the potency of Ca2+ o (Quinn et al., 1997). Spermine is the most potent CaSR agonist, followed by spermidine, then putrescine (Quinn et al., 1997). Spermine IC50 for suppression of PTH release from cultured bovine parathyroid cells is ~200 μM (Quinn et al., 1997). Blood polyamine concentrations in healthy humans are ~5-10 μM (Casti et al., 1982; Soda et al., 2009), concentrations that are likely sufficient to activate the CaSR in tissues where receptor density is high. In the lung, polyamines and other polycations stimulate CaSR-mediated airway contraction (Yarova et al., 2015) (described in Section VI). Intriguingly, other overlapping functions of the CaSR and polyamines exist, including promotion of osteoblast, keratinocyte, vascular smooth muscle cell and gastrointestinal epithelial cell differentiation and proliferation (Leach et al., 2014; Miller-Fleming et al., 2015; Riccardi and Kemp, 2012). Thus, polyamines may contribute to multiple (patho)physiological processes mediated by the CaSR.
Not surprisingly, additional positively charged molecules activate the CaSR, including poly-L-arginine, protamine, and aminoglycoside antibiotics including neomycin, tobramycin and gentamicin (McLarnon and Riccardi, 2002). Poly-L-arginine is a mimetic of eosinophil major basic protein (MBP), released to activate mast cells, neutrophils, basophils and macrophages in asthma and other allergic diseases. Ca2+ i mobilization in CaSR-HEK293 cells stimulated by the related eosinophil cationic protein (ECP) was completely absent in untransfected HEK293 cells and was blocked by structurally distinct CaSR inhibitors, demonstrating a CaSR-dependent signaling mechanism (Yarova et al., 2015).
B. Endogenous and exogenous allosteric modulators
Allosteric modulators bind to sites that are topographically distinct from the orthosteric binding site and act to either potentiate (positive allosteric modulators, or PAMs), inhibit (negative allosteric modulators, or NAMs) or have no effect on (neutral allosteric ligands, or NALs) the binding or efficacy of the orthosteric agonist. Allosteric modulators may also be agonists (or inverse agonists) in the absence of orthosteric agonists, and can simultaneously act as agonists and PAMs (PAM-agonists). CaSR PAMs have been termed type II calcimimetics, and CaSR NAMs calcilytics.
L-amino acids
L-amino acids are endogenous CaSR activators that are generally recognized as PAMs. Thus, L-amino acids have no activity in the absence of Ca2+ o or another cationic activator such as gadolinium or spermine, but potentiate CaSR-mediated responses in the presence of submaximal concentrations of cationic activators (Conigrave et al., 2000). In a Ca2+ i mobilization assay performed in CaSR-HEK293 cells, the magnitude of Ca2+ o potentiation mediated by 10 mM amino acids followed the rank order L-phenylalanine (L-Phe), L-tryptophan (L-Trp), L-histidine > L-alanine > L-serine, L-proline, L-glutamic acid > L-aspartic acid (but not L-lysine, L-arginine, L-leucine and L-isoleucine (Conigrave et al., 2000). Similarly, in human parathyroid cells in culture, aromatic amino acids such as L-Trp and L-Phe were the most potent L-amino acid CaSR activators in Ca2+ i mobilization assays (Conigrave et al., 2004). Thus, the CaSR, like a number of other class C GPCRs, is a promiscuous sensor of L-amino acids (Conigrave and Hampson, 2006; 2010; Smajilovic et al., 2014).
As would be expected for a positive binding interaction, L-amino acids and Ca2+ o markedly enhance the CaSR’s sensitivity to one another in a reciprocal manner (Conigrave et al., 2000). Based on observations of Ca2+ i mobilization and PTH secretion assays in vitro, amino acids support normal physiological Ca2+ o sensitivity and thus underpin the physiological Ca2+ o concentration set-point for the parathyroid at around 1.1 - 1.2 mM (Conigrave et al., 2004).
Recent crystal structures of the CaSR’s VFT (Zhang et al., 2016) and entire extracellular (Geng et al., 2016) domains as well as mutational studies suggest that L-amino acids and analogs might be better viewed as co-agonists of the receptor rather than PAMs (see Section III). As detailed later, L-amino acids display pronounced biased signaling properties (Section IIC) and L-amino acid signaling appears to be attenuated by PKC mediated phosphorylation of Thr888 in the C-terminal tail of CaSR (Section IVA).
γ-glutamyl peptides
Wang et al. demonstrated that the γ-glutamyl peptide, glutathione (GSH), is a potent activator of the CaSR and of another class C GPCR, the fish 5.24 receptor (Wang et al., 2006). Subsequently, various natural and synthetic analogs of glutathione were found to activate the CaSR in the presence of threshold Ca2+ o concentrations in a similar manner to L-amino acids. A receptor double mutant (Thr145Ala + Ser170Thr) exhibits similar impairments of function when exposed to either L-amino acids (Mun et al., 2005) or the GSH analog, S-methylglutathione (Broadhead et al., 2011). Interestingly, γ-glutamyl peptides active at the CaSR are also potent activators of kokumi taste (Amino et al., 2016; Ohsu et al., 2010).
pH
Large supraphysiological changes in buffer pH alter the potencies of Ca2+ o and Mg2+ o at the CaSR. In the blood, pH rarely varies by more than 0.2 units, however this represents a change in H+ concentration of ~58%. Such acidosis can occur in advanced chronic kidney disease (CKD), which has relevance to the CaSR (see Section V). Interestingly, altering buffer pH from 7.4 to just 7.2 or 7.6 elicits significant attenuation or enhancement of CaSR signaling, respectively, as observed in both HEK293 cells and bovine parathyroid cells (Campion et al., 2015). The site of H+ action is unknown though not apparently mediated via the CaSR’s extracellular histidine residues (Campion et al., 2015). Crucially, pathophysiologic changes in pH elicit significant changes in PTH secretion from isolated human parathyroid cells (Campion et al., 2015). This indicates the potential clinical relevance of altered acid or base balance in CaSR-modulated mineral metabolism.
Phosphate
Crystallization of the CaSR ECD has revealed up to four anion-binding sites (Geng et al., 2016) (see Section III) and a recent study has revealed that phosphate inhibits the CaSR directly and in a non-competitive manner (Centeno et al., 2019). This phosphate effect is more substantial than can be explained by buffering of free Ca2+ o ions, and mutation of Arg62 inhibits the phosphate action. Exposure of human and murine parathyroid cells to pathophysiologic phosphate concentrations induces rapid and reversible PTH secretion indicative of a receptor-mediated action (Centeno et al., 2019). Similarly, other anions, such as sulfate (SO4 2-), act as inhibitors of the CaSR (Geng et al., 2016) potentially also acting via Arg62 (Centeno et al., 2019).
Osmolarity
High sodium chloride (NaCl) concentrations are inhibitory for the CaSR such that concomitant Ca2+ o concentration-response curves are right-shifted, whereas lowering the NaCl concentration raises the potency of Ca2+ o for the CaSR (Quinn et al., 1998). Accordingly, in dispersed bovine parathyroid cells, raising extracellular osmolarity with either NaCl or sucrose elicits rapid (within minutes) and substantial PTH secretion, an effect that cannot be suppressed by raising Ca2+ o concentrations (Chen et al., 1987). While this means that the CaSR could represent an ionic strength sensor where it is expressed in, for example, the renal tubules or the subfornical organ of the brain, there is little evidence to date that the CaSR is a substantive contributor to mammalian osmoregulation. Indeed, Na+ is a well-known negative allosteric modulator of multiple class A GPCRs, where it binds in a conserved 7TM domain pocket. Therefore, allosteric modulation of GPCRs at least by Na+ is likely a general phenomenon. Nonetheless, some severe gain-of-function clinical CaSR mutations (see Section VI) can elicit a Bartter’s-like salt-wasting syndrome, while loss-of-function CaSR mutations can enhance the natriuretic response to loop diuretics indicative of mild Na+ retention (Huang and Miller, 2010; Miller, 2013).
Small molecule allosteric modulators
A detailed review on the discovery and development of CaSR small molecule drugs has recently been published (Nemeth et al., 2018). Therefore, for the purposes of this review, the focus will be on small molecules for which pharmacological or clinical data is available. To date, all CaSR small molecule binding sites have been localized to the 7TM domain and/or ECLs (Bu et al., 2008; Leach et al., 2016; Miedlich et al., 2004; Petrel et al., 2004; Petrel et al., 2003). These sites are distinct from the predominant Ca2+ o, L-amino acid, or γ-glutamyl binding sites in the ECD (see Section III), thus all small molecule CaSR drugs identified so far are allosteric.
For the majority of small molecule PAMs and NAMs, pharmacological characterization has been based on their ability to potentiate or inhibit a single concentration of Ca2+ o, usually in a Ca2+ i mobilization or IP accumulation assay (see Table 1). This approach provides a measure of modulator potency, which is a composite value of affinity, cooperativity (the magnitude and direction of modulator potentiation or inhibition of the orthosteric agonist) and efficacy (i.e. agonism or inverse agonism). While potency measurements facilitate drug comparisons in a series when in vitro assays are performed under identical conditions, they can be misleading when different assay conditions are employed (e.g. different orthosteric agonist concentrations, different signaling outputs) (Gregory et al., 2018). Therefore, more recent work has quantified PAM and NAM affinity, cooperativity and efficacy values as separate parameters using an operational model of allosterism or an allosteric ternary complex model (Cook et al., 2015; Davey et al., 2012; Diepenhorst et al., 2018; Gregory et al., 2018; Leach et al., 2016; Leach et al., 2013).
Table 1. Representative CaSR agonists or endogenous and small molecule allosteric modulators, and their pharmacological properties.
Ligand | Structure | (Cell type or model, assay) | Potencya or affinityb | Cooperativityc with Ca2+ o | References |
---|---|---|---|---|---|
Agonists | |||||
Ca2+ o | Ca2+ | Human, PTH release | pEC50 2.9 | NA | (Brown, 1983; 1991; Gregory et al., 2018, 2020; Quinn et al., 1997; Ramirez et al., 1993) |
Parathyroid cell, PTH release | pEC50 3.0 | NA | |||
HEK293, Ca2+ i | pEC50 2.5 – 3.5 | NA | |||
HEK293, Caa2+ i | pKB 3.0 | NA | |||
Spermine |
![]() |
HEK293, Ca2+ i | pEC50 3.3 – 4.4 | NA | (Gregory et al., 2018; Quinn et al., 1997) |
Neomycin |
![]() |
HEK293, Ca2+ i | pEC50 4.4 | NA | (McLarnon et al., 2002) |
PAMs | |||||
L-Trp |
![]() |
HEK293, Ca2+ i | pEC50 2.6 | ND | (Conigrave et al., 2000) |
Cinacalcet |
![]() |
HEK293, Ca2+ i | pKB 5.9 – 6.7 | 2.6 – 4.7 | (Cook et al., 2015; Davey et al., 2012; Diepenhorst et al., 2018; Leach et al., 2016; Leach et al., 2013) |
HEK293, IP1 | pKB 6.1 | 2.6 – 4.8 | |||
HEK293, pERK1/2 | pKB 5.9 – 6.5 | 1.3 – 2.9 | |||
HEK293, membrane ruffling | pKB 8.1 | 2.6 | |||
HEK293, SRF-RE lucd | pKB 7.1 | 4.5 | |||
NPS R-568 |
![]() |
HEK293, Ca2+ i | pKB 6.0 – 6.6 | 3.0 - 3.9 | (Cook et al., 2015; Davey et al., 2012; Gregory et al., 2018; Keller et al., 2018; Lu et al., 2009) |
CHO, Aequorin | pKB 6.2 | 2.7 | |||
CHO/HEK293, IP1 | pKB 6.2 – 6.8 | 4.3 - 4.5 | |||
HEK293, pERK1/2 | pKB 5.6 – 6.6 | 2.0 – 5.1 | |||
HEK293, membrane ruffling | pKB 9.4 | 1.7 | |||
Calindol |
![]() |
HEK293, Ca2+ i | pKB 6.3 | 5.4 | (Cook et al., 2015) |
HEK293, IP1 | pKB 6.4 | 4.7 | |||
HEK293, pERK1/2 | pKB 5.2 | 8.1 | |||
Evocalcet |
![]() |
HEK293, Ca2+ i | pEC50 7.0 | ND | (Kawata et al., 2018) |
R,R-calcimimetic B |
![]() |
HEK293, Ca2+ i | pKB 7.2 | 1.9 | (Cook et al., 2015) |
HEK293, IP1 | pKB 7.0 | 3.2 | |||
HEK293, pERK1/2 | pKB 7.1 | 3.0 | |||
AC265347 |
![]() |
HEK293, Ca2+ i | pKB 6.2 – 6.4 | 2.5 – 4.3 | (Cook et al., 2015; Diepenhorst et al., 2018) (Leach et al., 2016) |
HEK293, IP1 | pKB 7.3 - 8.0 | 4.0 – 4.7 | |||
HEK293, pERK1/2 | pKB 6.1 - 6.7 | 4.5 – 10 | |||
HEK293, SRF-RE luc | pKB 6.2 | 13 | |||
BTU compound 13 |
![]() |
HEK293, Ca2+ i | pKB 6.7 | 3.2 | (Diepenhorst et al., 2018) |
HEK293, IP1 | pKB 7.2 | 2.9 | |||
HEK293, pERK1/2 | pKB 6.2 | 1.2 | |||
HEK293, SRF-RE luc | pKB 6.5 | 17 | |||
Etelcalcetide |
![]() |
HEK293, IP1 | pEC50 4.6 | ND | (Walter et al., 2013) |
NAMs | |||||
NPS 2143 |
![]() |
HEK293, Ca2+ i | pKB 6.2 - 6.7 | 0.3 – 0.5 | (Davey et al., 2012; Leach et al., 2016) |
HEK293, pERK1/2 | pKB 6.2 - 6.6 | 0.3 - 0.6 | |||
HEK293, membrane ruffling | pKB 7.8 | 0.3 | |||
NPSP795 |
![]() |
HEK293, assay not disclosed | pIC50 7.1 | ND | (Kumar et al., 2010) |
Ronacaleret |
![]() |
HEK293, Ca2+ i | pKB 6.4 | 0.03 | (Josephs et al., 2019) |
JTT-305/MK-5442 |
![]() |
PC12h, zif luce | pIC50 7.9 | ND | (Shinagawa et al., 2011) |
ATF936 |
![]() |
HEK293, Ca2+ i | pKB 7.6 | 0.005 | (Josephs et al., 2019) |
BMS compound 1 |
![]() |
HEK293, Ca2+ i | pKB 7.0 | 0.03 | (Josephs et al., 2019) |
3H-pyrimidine-4-one compound (R)-2h |
![]() |
HEK293, Ca2+ i | pKB 7.8 | 0.009 | (Josephs et al., 2019) |
Benzimidazole compound 40 |
![]() |
Hamster fibroblasts, Ca2+ i | pIC50 8.4 | ND | (Gerspacher et al., 2010) |
Mixed PAM/NAM | |||||
Calhex 231 |
![]() |
HEK293, Ca2+ i | pKB 6.5 | ND | (Gregory et al., 2018) |
Small molecule PAMs
The structural and chemical diversity of small molecule CaSR PAMs is relatively limited, with few distinct series discovered. Two chemically and structurally related small molecule PAMs, cinacalcet and evocalcet (Table 1), are clinically approved. Cinacalcet is FDA approved for the treatment of primary hyperparathyroidism in patients who cannot undergo parathyroidectomy, and for hypercalcemia in adults with parathyroid carcinoma. Cinacalcet is also FDA approved for secondary hyperparathyroidism in patients on renal replacement therapy, and has been used off-label to treat naturally occurring loss-of-function mutations in the CaSR or its signaling partners that cause disorders of Ca2+ o and PTH homeostasis (described in Section VI). Cinacalcet was the first GPCR allosteric modulator to be approved for clinical use in 2004. Evocalcet was approved in Japan in 2018 for the treatment of secondary hyperparathyroidism in patients on dialysis. Cinacalcet and evocalcet potentiate Ca2+ o activity at the CaSR, thus left-shifting the Ca2+ o-PTH concentration-response relationship in the body. This means lower Ca2+ o concentrations are required to suppress PTH release, thus normalizing elevated serum PTH levels. However, both cinacalcet and evocalcet carry a risk of hypocalcemia in patients that limits their clinical utility (Fukagawa et al., 2018), presumably in part, from potentiation of the CaSR in the kidney and enhanced CaSR-mediated calcitonin secretion from thyroid parafollicular C cells (see Section V). Further, cinacalcet and evocalcet are associated with adverse gastrointestinal side effects, including nausea and vomiting, which may occur via the CaSR expressed in the gastrointestinal tract. In rats and humans, however, evocalcet appears to have reduced actions in the gastrointestinal tract in comparison to cinacalcet (Fukagawa et al., 2018; Kawata et al., 2018).
Cinacalcet and evocalcet belong to the arylalkylamine family of PAMs, derived from the nonselective calcium channel blocker, fendiline. A number of structurally related arylalkylamine PAMs have been identified, including NPS R-467 and NPS R-568 (the precursors to the discovery of cinacalcet), calindol and calcimimetic B (Table 1). The activity of these PAMs is highly dependent upon their stereoselectivity, where the R-configuration of the methyl between the aromatic and secondary nitrogen is more active than the S-configuration (Nemeth et al., 2018). While NPS R-568, cinacalcet and calindol exhibit similar affinity and cooperativity values when measured in a Ca2+ i mobilization assay (Cook et al., 2015; Davey et al., 2012; Diepenhorst et al., 2018; Keller et al., 2018; Leach et al., 2016), R,R-calcimimetic B has a roughly 10-fold higher affinity but comparable cooperativity (Cook et al., 2015). Although concentrations of cinacalcet that exceed 1 μM weakly activate the CaSR in the absence of divalent cations (Nemeth et al., 2018), suggesting it is a “PAM-agonist”, arylalkylamine PAMs demonstrate negligible agonism at concentrations that robustly potentiate CaSR activity (Cook et al., 2015; Keller et al., 2018). In contrast, R,R-calcimimetic B is a PAM and a partial agonist at micromolar concentrations (Cook et al., 2015). Arylalkylamine PAMs also exhibit pronounced positive interactions with L-amino acids (Zhang et al., 2002a) and glutathione (Broadhead et al., 2011).
A benzothiazole series of CaSR PAMs that is structurally and chemically distinct from the arylalkylamines has been discovered. These PAMs include the small benzothiazole, AC265347 (Table 1), which has been characterized in detail. AC265347 has comparable affinity and cooperativity to cinacalcet when measured in a Ca2+ i mobilization assay (Cook et al., 2015; Diepenhorst et al., 2018; Leach et al., 2016), and similar to the arylalkylamine PAMs, AC265347 is a PAM-agonist, although as an agonist AC265347 is more potent and efficacious than the arylalkylamines (Cook et al., 2015). Although AC265347 has not been tested in humans, in healthy rats AC265347 suppressed serum PTH levels with greater potency than cinacalcet and a lower propensity to cause hypocalcemia (Ma et al., 2011).
Trisubstituted urea compounds have been identified as another potent class of CaSR PAMs (Temal et al., 2013) (Table 1). Benzothiazole trisubstituted urea (BTU) compound 13 (Deprez et al., 2013) is the best characterized of this series. BTU compound 13 has similar affinity and cooperativity to cinacalcet at the CaSR in a Ca2+ i mobilization assay (Cook et al., 2015; Diepenhorst et al., 2018). Much like AC265347, BTU compound 13 suppressed PTH levels in a rat model of CKD, while avoiding significant hypocalcemia (Deprez et al., 2013).
Peptide PAM, etelcalcetide
In 2017, a novel CaSR PAM, etelcalcetide (chemical name N-acetyl-D-cysteinyl-D-alanyl-D-arginyl-D-arginyl-D-arginyl-D-alanyl-D-argininamide disulfide with L-cysteine hydrochloride), was FDA approved for the treatment of secondary hyperparathyroidism in patients with CKD on dialysis. Etelcalcetide is administered intravenously at the end of dialysis. Similar to cinacalcet and evocalcet, etelcalcetide is associated with adverse gastrointestinal side effects and hypocalcemia (Hamano et al., 2017).
Etelcalcetide is comprised of 7 D-amino acids linked via a disulfide bond to L-cysteine. Not surprisingly given it is the only peptide CaSR PAM identified, etelcalcetide has a unique mode of PAM action in comparison to small molecule PAMs, which involves binding to the CaSR via disulfide bond formation (Alexander et al., 2015) (see Section III). Although etelcalcetide has been classified as a PAM-agonist, assays used to discern agonism contained 0.5 mM MgCl2, therefore it is currently uncertain whether observed etelcalcetide efficacy for stimulation of IP1 accumulation in the absence of Ca2+ o is true agonism, or potentiation of Mg2+ o (Walter et al., 2013). The affinity and cooperativity of etelcalcetide at the CaSR has not been quantified, but its potency for potentiation of 1.2 mM Ca2+ o in a HEK293 IP1 accumulation assay was 25 μM (Walter et al., 2013).
Small molecule NAMs
Due to the role of the CaSR in regulation of PTH secretion, there was significant interest in the development of CaSR NAMs that could stimulate PTH release. Intermittent and transient increases in serum PTH levels enhance the formation of new bone via the differentiation and proliferation of bone forming osteoblasts. This is evidenced by clinical use of recombinant PTH1-34 injections to promote bone formation in osteoporosis. However, if PTH levels remain elevated, PTH stimulates the differentiation and proliferation of bone resorbing osteoclasts, resulting in bone breakdown (Dobnig and Turner, 1997).
Although several pharmaceutical companies have embarked on CaSR NAM discovery programs, similar to CaSR PAMs, there is fairly limited structural and chemical diversity in the NAM scaffolds identified to date. NPS 2143 (Table 1) was one of the first CaSR NAMs to be discovered (Gowen et al., 2000), and is structurally and chemically related to cinacalcet and other arylalkylamines. Like CaSR PAMs, NAMs have generally been evaluated for their potency to inhibit a single Ca2+ o (usually EC80) concentration. Nevertheless, more recent studies have employed an operational model of allosterism to quantify NPS 2143 activity, and have indicated that NPS 2143 binds at the CaSR with μM to sub-μM affinity depending on the assay (Table 1) (Davey et al., 2012; Gregory et al., 2020; Leach et al., 2016; Leach et al., 2013). Importantly, NPS 2143 is a partial NAM at the CaSR, meaning that it does not fully inhibit Ca2+ o-mediated signaling (Cook et al., 2015; Gregory et al., 2018; Leach et al., 2016).
In rats, NPS 2143 stimulated the release of PTH, resulting in an increase in bone turnover markers, but it did not promote the formation of new bone (Gowen et al., 2000). The lack of new bone formation was hypothesized to be due to the prolonged, rather than transient, PTH release in response to NPS 2143, resulting in both bone formation and resorption. Efforts to develop shorter acting CaSR NAMs based on the structure of NPS 2143 led to the discovery of ronacaleret (Fitzpatrick et al., 2012; Fitzpatrick et al., 2011a; Fitzpatrick et al., 2011b) and JTT-305/MK-5442 (Shinagawa et al., 2011) (Table 1). However, in rats, JTT-305/MK-5442 did not increase bone mass and density (Fisher et al., 2012), while in human clinical trials, both ronacaleret and JTT-305/MK-5442 lacked efficacy in treating postmenopausal osteoporosis (Fitzpatrick et al., 2012; Fitzpatrick et al., 2011a; Fitzpatrick et al., 2011b; Halse et al., 2014).
Further efforts to identify additional CaSR NAMs that may prove successful in treating osteoporosis lead to the discovery of four chemically distinct NAM series, exemplified by the quinazolinones ATF936 and AXT914 (Gerspacher et al., 2010), the pyridine BMS compound 1 (Arey et al., 2005), a series of 3H-quinazoline-4-ones and 3H-pyrimidine-4-ones (Didiuk et al., 2009; Shcherbakova et al., 2005), and benzimidazoles (Gerspacher et al., 2010). A recent study revealed that the affinity of ATF936 was 17-fold higher than that of ronacaleret, while ATF936 also demonstrated higher negative cooperativity (Josephs et al., 2019) (Table 1). However, despite findings that the quinazolinone NAMs may be superior to ronacaleret in terms of desirable drug properties, when AXT914 was evaluated for its effects on bone turnover in humans, the trial was terminated early due to a lack of effect on bone turnover markers and a propensity to cause hypercalcemia (John et al., 2014).
Following the failure of three different NAMs in human clinical trials of osteoporosis, the development of CaSR NAMs diminished. However, there has been recent interest in repurposing these NAMs for the treatment of Ca2+ o homeostasis disorders caused by gain-of-function mutations in the CaSR or its interactors (described in Section VI). Indeed, the NAM, NPSP795 (SHP635) has recently undergone clinical testing for its therapeutic potential in the treatment of ADH1 (Roberts et al., 2019).
Calhex231: a mixed PAM and NAM
Although the arylalkylamine, calhex 231, was originally classified as a NAM based on its ability to inhibit an EC100 Ca2+ o concentration (Kessler et al., 2006), a recent study has revealed that calhex 231 is both a PAM and a NAM (Gregory et al., 2018). This novel mode-switching mechanism appears to be due to allostery across the CaSR dimer, where calhex 231 acts as a PAM when it occupies a single protomer in the dimer, and a NAM when bound to both protomers. Mixed PAM and NAM activity was observed in HEK293 cells stably expressing the CaSR, and in primary cultures of human parathyroid cells, demonstrating that mode-switching may occur under physiological conditions (Gregory et al., 2018). As several CaSR NAMs have been characterized based on their ability to modulate only a single Ca2+ o concentration, it is unclear at present whether other CaSR allosteric modulators also exhibit mixed PAM and NAM activity. However, whereas other CaSR NAMs were identified from high throughput screens of large compound libraries, calhex231 originated from a PAM scaffold (Kessler et al., 2006; Kessler et al., 2004), which likely contributes to its mixed PAM and NAM activity.
C. Biased agonism and biased allosteric modulation
Given that the CaSR responds to a diverse array of different ligands, it is unsurprising that the CaSR is subject to biased agonism and biased modulation. Biased agonism is the phenomenon by which distinct ligands stabilize preferred GPCR signaling states, with each state having the potential to stimulate or inhibit discrete subsets of the full repertoire of intracellular signaling pathways that couple to a given receptor (Kenakin and Christopoulos, 2013). This is in contrast to the earlier dogma that all agonists activate the same subsets of GPCR signaling pathways to greater (e.g. full agonists) or lesser (e.g. partial agonists) extents. Similarly, biased modulation arises when an allosteric ligand differentially modulates different agonist-mediated signaling pathways.
For instance, in CaSR-HEK293 cells, Ca2+ o preferentially mediates stimulation of Ca2+ i mobilization over pERK1/2, while spermine preferentially activates pERK1/2 (Thomsen et al., 2012a). Similarly, L-amino acids activate Ca2+ i mobilization and ERK phosphorylation (Lee et al., 2007), and also inhibit cAMP synthesis. However, they are inactive in stimulating PI-PLC and various other signaling events including Rho-dependent actin stress fibre formation (Davies et al., 2006) and CREB phosphorylation (Avlani et al., 2013), and appear to promote Ca2+ i mobilization via a G12/13 / TRPC1-dependent Ca2+ o-influx pathway (Rey et al., 2006; Rey et al., 2005).
Evidence from FHH patients suggests CaSR bias may arise, in part, from spatial and temporal CaSR signaling patterns. Loss-of-function germline mutations of the AP2S1 gene, which encodes the sigma subunit of the heterotetrameric adaptor-related protein complex-2 (AP2σ), cause FHH type 3 (FHH3) (Hannan et al., 2015a; Nesbit et al., 2013b). AP2σ forms part of the heterotetrameric AP2 that plays a critical role in clathrin-mediated endocytosis. AP2σ mutations increase CaSR cell surface expression yet reduce CaSR signaling, because CaSR residency time in clathrin-coated pits is increased, consequently impairing CaSR Gq/11 signaling from endosomes (Gorvin et al., 2018c). In contrast, Gi/o-mediated signaling is less sensitive to AP2σ mutations. Thus, whereas the plasma-membrane localized CaSR signals via Gq/11 and Gi/o, endosomal CaSRs signal predominantly via Gq/11 (Gorvin et al., 2018c).
It must be noted that many of the studies reporting differential CaSR-mediated pathway activation have not been performed in a systematic manner using identical conditions across assays (e.g. buffers, duration of agonist stimulation etc.), or the same cellular background. Further, bias has not been quantified in these studies. Therefore, it remains to be definitively proven whether biased agonism is truly operative at the CaSR, or whether previous observations were due to observational bias (e.g. different assay conditions, different cell types) or system bias (e.g. the relative efficiency with which the receptor couples to different pathways).
Nonetheless, small molecule allosteric modulators do appear to exhibit true biased modulation at the CaSR. Evidence of biased modulation comes from reversals in the magnitude of cooperativity in different pathways between distinct PAMs or NAMs, or from differences in PAM or NAM affinity for receptor states that couple to different signal transducers. For instance, while cinacalcet and NPS 2143 preferentially potentiate or inhibit, respectively, Ca2+ o-mediated Ca2+ i mobilization over pERK1/2, AC265347 and R,R-calcimimetic B show reversed bias for CaSR-mediated pERK1/2 over Ca2+ i mobilization (Cook et al., 2015; Diepenhorst et al., 2018; Leach et al., 2016). Similarly, AC265347, NPS R-568 and calindol, but not cinacalcet or R,R-calcimimetic B, have a higher functional affinity (i.e. an affinity quantified in a functional assay using an operational model of allosterism (Leach et al., 2007)) for the CaSR state that signals to IP1 accumulation versus Ca2+ i mobilization (Cook et al., 2015; Diepenhorst et al., 2018), while cinacalcet, NPS R-568 and NPS 2143 all have a higher functional affinity for the CaSR state that couples to membrane ruffling (Davey et al., 2012).
Evidence for small molecule PAM and NAM bias also comes from pharmacochaperone studies, which reveal that while cinacalcet, AC265347 and BTU compound 13 are all PAMs in multiple CaSR-mediated signaling assays, only cinacalcet positively modulates the trafficking of an endosomally-trapped naturally occurring mutant CaSR, rescuing its cell surface expression back to levels comparable to wild type CaSR (Cook et al., 2015; Diepenhorst et al., 2018; Leach et al., 2013). In contrast, while NPS 2143 is a NAM of CaSR signaling, it is a PAM of loss-of-expression mutant receptor trafficking (Leach et al., 2013). This is in contrast to the actions of NPS 2143 at the wild type CaSR, where it reduces CaSR surface expression (Huang and Breitwieser, 2007), suggesting naturally occurring mutations (which cause Ca2+ o homeostasis disorders; see Section VI) may engender bias in CaSR function. Indeed, Ca2+ o-mediated bias towards Ca2+ i mobilization is abolished by some naturally occurring mutations (Leach et al., 2012).
Although the physiological relevance of biased agonism and biased modulation at the CaSR is not at present known, differences in the propensity of CaSR PAMs to cause hypocalcemia could be linked to this phenomenon. For instance, as already mentioned, R,R-calcimimetic B and AC265347 are effective suppressors of PTH release. However, in comparison to cinacalcet, R,R-calcimimetic B and AC265347 demonstrate reduced propensity to cause hypocalcemia in rats successfully treated for severe hyperparathyroidism induced by CKD (R,R-calcimimetic B) or in normal rats (AC265347). The reduced incidence of hypocalcemia with R,R-calcimimetic B and AC265347 is presumably linked, in part, to their lower potency and efficacy for the stimulation of calcitonin secretion versus suppression of PTH release (Henley et al., 2011; Ma et al., 2011). Importantly, although suppression of PTH release has been associated with pERK1/2, calcitonin release is independent of pERK1/2 in rat medullary thyroid carcinoma (MTC) cells (Thomsen et al., 2012b). This highlights differences in the coupling specificity of the CaSR in distinct tissues and is consistent with observations that when compared to cinacalcet, AC265347 and R,R-calcimimetic B show reversed bias for CaSR-mediated pERK1/2 over Ca2+ i mobilization.
Another apparent difference between CaSR PAMs points towards putative clinical advantages for cinacalcet. The CaSR agonist, Sr2+, reduces the differentiation of bone resorbing osteoclasts (Bonnelye et al., 2008) and stimulates osteoclast apoptosis (Hurtel-Lemaire et al., 2009) (described in Section V). In cultured osteoclasts differentiated from human CD14+ monocytes, while cinacalcet potentiated Sr2+-mediated tartrate-resistant acid phosphatase (TRAP) expression (a marker of osteoclast activity), and robustly inhibited osteoclast-mediated hydroxyapartite artificial bone resorption, AC265347 and BTU compound 13 were without effect in these two assays (Diepenhorst et al., 2018). Although it is not clear whether differences in the biased profile of AC265347 and BTU compound 13 versus cinacalcet are responsible for their distinct PAM activities in osteoclasts, it is interesting that only cinacalcet, and not AC265347 or BTU compound 13 can pharmacochaperone loss-of-function mutant CaSRs potentially via differential stabilization of different conformations of the CaSR. A more detailed understanding of the signaling and trafficking pathways that couple the CaSR to its many physiological responses will aid our understanding of why the CaSR responds to so many endogenous activators, and may facilitate the development of biased compounds with improved tissue-specific effects.
In addition to bias engendered by small molecule allosteric modulators, CaSR autoantibodies that cause acquired hypocalciuric hypercalcemia (AHH) can act as biased allosteric modulators. Biased autoantibodies directed against the CaSR VFT can potentiate IP accumulation, while inhibiting pERK1/2 generation (Makita and Iiri, 2014; Makita et al., 2007), while others inhibit pERK1/2 generation but have no effect on IP accumulation (Pallais et al., 2011). Importantly, cinacalcet corrected the severe hypercalcemia associated with AHH caused by a biased autoantibody (Makita et al., 2019). Taken together, these findings once again highlight how bias and allostery are key features of CaSR (patho)physiology and drug actions.
III. Receptor structure
To date, the complete structure of the CaSR has not been determined. Current CaSR structural knowledge comes from the inactive (Geng et al., 2016) and active (Geng et al., 2016; Zhang et al., 2016) crystal structures of the CaSR ECD in isolation, from mutagenesis studies and homology modeling of the 7TM based on the crystal structures of the metabotropic glutamate receptors (mGluR) 1 and 5 7TMs (Christopher et al., 2015; Christopher et al., 2019; Dore et al., 2014), and from comparisons with the low resolution cryogenic electron microscopy (cryo-EM) structure of mGluR5 (Koehl et al., 2019).
The CaSR is an obligate homodimer (Bai et al., 1998a; Pidasheva et al., 2006; Ray et al., 1999; Romano et al., 1996; Ward et al., 1998; Zhang et al., 2001), with each protomer comprised of an extracellular VFT domain (amino acids 20 – 542) and a cysteine-rich (CR) domain (9 Cys residues within amino acids 542 – 612), which links the VFT to the prototypical GPCR 7TM domain (amino acids 613 - 862) (Figure 2). The 7TM domain is followed by a long intracellular tail (amino acids 863 – 1078), which is predicted to be largely unstructured, but is important for trafficking and phosphorylation (Bai et al., 1998b; Chang et al., 2001; Stepanchick et al., 2010; Zhuang et al., 2012).
Figure 2. Structural conformation of the CaSR.
(A). Model of the CaSR based on homology with full length mGluR5 (PDB 6N51). The CaSR (cartoon ribbon) comprises an extracellular VFT domain, composed of lobe 1 (LB1, dark blue) and lobe 2 (LB2, teal), and a cysteine rich (CR) domain (yellow) anchored to the 7TM (orange). (B). Inactive ECD monomer (PDB 5K5T). The bilobed VFT adopts an open conformation revealing a conserved binding cleft between the two lobes. (C). Inactive ECD dimer (left, front view; right, side view). The CR domains of the inactive ECDs are separated. (D). Active ECD monomer (PDB 5K5S). Upon activation, the bilobed VFT closes the amino acid binding site, narrowing the cleft. E). Active ECD dimer (left, front view; right, side view). Upon activation, each protomer (orange and yellow) is drawn closer together.
A. CaSR extracellular domain
Structural overview of the CaSR extracellular domain
The VFT extends outside the cell and is comprised of two lobe sub-domains (lobe 1 & 2; Figure 2), with each lobe forming part of a ligand binding cleft. In other class C GPCRs, this cleft forms the orthosteric binding pocket (Kunishima et al., 2000; Muto et al., 2007; Tsuchiya et al., 2002). However, in the CaSR it is an allosteric or co-agonist binding site for L-amino acids, with Ca2+ o and other cations binding elsewhere.
Two recent VFT crystal structures confirm that the CaSR VFT forms a dimer, with each CaSR protomer orientated next to each other as mirror images (Figure 2). The dimer orientation of the extracellular domain is similar to that reported for other Class C GPCRs, including mGluRs (Kunishima et al., 2000; Muto et al., 2007; Tsuchiya et al., 2002) and the γ-aminobutyric acid subtype B (GABAB) receptor (Geng et al., 2013; Geng et al., 2012). In the inactive state, the two VFT lobes adopt an open conformation (buried surface area of 740 Å2, calculated using methods described in (Krissinel and Henrick, 2007)) and the interdomain cleft is empty. In contrast, the active state structures adopt a closed conformation and a resulting increase in the buried surface area to just over 1000 Å2 between the VFT lobes (Figure 2). Upon VFT closure, the interdomain cleft interface rotates 29°, mediated by interactions between the two lobes of the VFT (Geng et al., 2016).
The crystal structure of the CaSR VFT plus the CR domains shows an 83 Å distance between the CR domains when the CaSR VFT is in the open (inactive) conformation, which is reduced to 23 Å once the VFT is closed (active; Figure 2D) (Geng et al., 2016). This change is consistent with other X-ray structures of class C ECDs (Chappell et al., 2016; Muto et al., 2007), likely driving a similar reorientation of the 7TM domains as seen in the mGlu5 cryo-EM structure to produce a transition-state receptor (Koehl et al., 2019). This reorientation is sustained by the rigid CR domain and its 9 Cys residues, which form 5 covalent disulfide bonds; four within the CR domain and one that anchors the CR domain with lobe 2 of the VFT. Consequently, mutation of the Cys residues compromises this rigidity, impacting significantly on receptor function (Fan et al., 1998).
Amino acid and γ-glutamyl peptide binding site
Although Ca2+ o has long been considered the orthosteric agonist for the CaSR, Ca2+ o does not occupy the conserved cleft that forms the orthosteric binding site in other class C GPCRs. Both mutagenesis (Mun et al., 2005; Mun et al., 2004; Zhang et al., 2014a; Zhang et al., 2002b), and more recently the crystal structures of the VFT, have revealed that L-amino acids bind the conserved cleft (between lobe 1 and 2) similar to L-Glu binding in the mGluRs (Wellendorph and Bräuner-Osborne, 2009). Thus, L-amino acids and analogs might be better viewed as co-agonists rather than PAMs.
The binding of L-Trp (Geng et al., 2016) or the tryptophan derivative, L-1,2,3,4-tetrahydronorharman-3-carboxylic acid (TNCA) (Zhang et al., 2016), stabilizes closing of the bilobed domains through hydrogen bonding and hydrophobic interactions with the receptor. Mutational analysis of residues in the conserved interdomain cleft support binding of L-Trp here (Mun et al., 2005; Mun et al., 2004; Zhang et al., 2014a; Zhang et al., 2002b). Interestingly, residues in the conserved cleft are also important for Ca2+ o activation of the CaSR (Bräuner-Osborne et al., 1999; Geng et al., 2013; Jacobsen et al., 2017; Kunishima et al., 2000; Muto et al., 2007; Tsuchiya et al., 2002), suggesting that L-amino acids are required for Ca2+ o activation in line with a classification as co-agonists. However, these mutational studies have not accounted for mutation-induced changes in receptor expression, therefore the mutation-induced signaling impairments may be due to reduced receptor expression and consequent reductions in apparent agonist efficacy.
Receptor contacts with L-Trp or TNCA are predominantly through backbone interactions, and the fact that these interactions are largely not L-Trp or TNCA specific, means other amino acids could be accommodated within this pocket, explaining the L-amino acid promiscuity of the CaSR (see Section IIB). Interestingly, TNCA was not included as a constituent of the crystallization conditions and was hypothesized to have been acquired during protein purification. This highlights not only the diversity of ligands that can bind and activate the CaSR, but also suggests that TNCA has such high affinity for the CaSR that it is difficult to remove during the purification process.
γ-glutamyl peptides are also potent CaSR PAMs that can promote Ca2+ o dependent Ca2+ i mobilization, suppress intracellular cAMP levels and inhibit PTH secretion from normal parathyroid cells (see Section IIB) (Broadhead et al., 2011). This activity is lost when Thr145 and Ser170 located in the interdomain cleft are mutated to Ala, indicating that the γ-glutamyl peptides likely share the same binding site as the amino acids (Broadhead et al., 2011; Mun et al., 2005).
Cation binding sites
In both crystal structures of the CaSR VFT domain, cation binding sites were identified, but these sites differed in their number, location (with the exception of cation binding site 1) and the cation that was bound to each site (Figure 3).
Figure 3. Binding sites within the CaSR crystal structures.
ECD conformations of the (A) active (PDB 5K5S), and (B) inactive (PDB 5K5S) calcium bound structures; and the VFT conformations of the (C) active magnesium bound (PDB 5FBK) and (D) active magnesium and gadolinium bound (PDB 5FBH) structures. Crystal structures are shown as cartoon ribbon within the transparent molecular surface and colored as in Figure 2. Hydrogen bond interactions (dashed lines) of calcium (red spheres), magnesium (green spheres), and gadolinium (yellow spheres) with key residues or water molecules (black spheres), are shown for each proposed binding site.
Anomalous difference mapping indicated four Ca2+ o binding sites in the VFT structure solved by Geng et al. (Figure 3) (Geng et al., 2016). In lobe 1 of the active VFT conformation, backbone carbonyl oxygen atoms of Ile81, Ser84, Leu87, and Leu88 coordinate Ca2+ o binding at cation binding site 1 (PDB: 5K5S). There was no Ca2+ o coordinated at cation binding site 1 in the inactive structure (PDB: 5K5T), even though this site is not significantly different in the active versus inactive structures (Geng et al., 2016). As such, it is possible that Ca2+ o, which was used at a lower concentration in the crystallization conditions for the inactive structure, could bind to this site without the need for the VFT domain being closed.
Cation binding site 2 is located adjacent to the L-Trp binding site above the interdomain cleft in lobe 1 of the VFT. Cation binding site 2 is occupied by Ca2+ o in both the inactive and active structures, where Ca2+ o is coordinated by the hydroxyl group of Thr100 in both states and by the carbonyl of Asn102 via a water molecule in the active structure. Thr145 also lines cation binding site 2 and forms part of the L-Trp binding cleft in the active state (Geng et al., 2016).
The hydroxyl groups of Ser302 and Ser303 coordinate cation binding site 3, either directly or indirectly through water molecules, at the edge of the interdomain cleft of lobe 2. The closing of lobe 1 and lobe 2 of the VFT is facilitated by Ca2+ o stabilization of a conformation that permits an interdomain hydrogen bond interaction between lobe 1 residue Arg66 and lobe 2 residue Ser301 (Geng et al., 2016).
Finally, upon agonist binding, cation binding site 4 forms part of the homodimer interface bridging the lobe 2 domain of one subunit and the CR domain of the second subunit. Three interfacial residues, the carboxylate group of Asp234 and carbonyl oxygen of Glu231 and Gly557, coordinate Ca2+ o binding to site 4 (Geng et al., 2016).
The anomalous difference map intensities varied at each of the Ca2+ o binding sites where intensity was ranked as Ca2+ binding site 1 = 2 > 3 > 4. The lower anomalous signal for Ca2+ o in sites 3 and 4 indicates incomplete occupancy or higher flexibility at these positions in the crystal lattice. The authors suggested the lower signal reflects a lower affinity for Ca2+ o at these sites. In support of a lower Ca2+ o binding affinity for cation binding site 4, the authors proposed that Ca2+ o binding at site 4 stabilizes the active homodimer conformation and thus the site is occupied only at elevated concentrations required for receptor activation (Geng et al., 2016).
In contrast to the structures by Geng et al. (Geng et al., 2016), Zhang et al. (Zhang et al., 2016) identified two cation binding sites in their active VFT structures. Electron density and geometric restraints were used to identify Mg2+ occupying these cation binding sites, one of which overlapped with cation binding site 1 in the structure by Geng et al. However, in contrast to the unoccupied cation binding site 1 in the inactive structure by Geng et al., cation binding site 1 was occupied by Mg2+ in the inactive structure by Zhang et al. The Mg2+ is coordinated by Ser84 and backbone interactions with Ile81, Ile87 and Leu88, in addition to two water molecules. This site is similarly occupied by a Mg2+ cation in the rat mGluR1 VFT structure (Kunishima et al., 2000).
The second Mg2+ binding site (cation binding site 5) is located at the dimerization interface of lobe 2 and is coordinated through Ser240 and four water molecules (Zhang et al., 2016). The highly conserved residues Glu228 and Glu231 from one protomer and Glu241 from the other protomer surround this site.
Anomalous difference maps identified a Gd3+ binding site (cation binding site 6), coordinated by Glu232, Glu228 and Glu229 adjacent to cation binding site 5 on the lobe 2 dimerization interface (PDB: 5FBN) (Zhang et al., 2016). The Glu228Ile and the double mutant Glu228Ile/Glu229Ile, have previously been shown to reduce Mg2+ o induced Ca2+ i mobilization, therefore other cations could bind here (Huang et al., 2009).
The crystal structures of the ECD suggest that Ca2+ o and other cations play a role in 1) local stabilization of the CaSR ECD; and 2) activation of the receptor via stabilization of the homodimer through cation binding at sites 4 – 6 (Geng et al., 2016; Jensen et al., 2002; Zhang et al., 2016). It is unknown whether Ca2+ o alone can activate the receptor, or whether it requires the presence of the cleft-binding ligands. Although Geng et al. obtained an active (closed) structure in the absence of amino acids, an unidentified continuous stretch of density in the conserved interdomain cleft was observed, which could be attributed to an endogenous ligand or a ligand acquired during the crystallization process. If ligands that bind the conserved interdomain cleft are difficult to remove during crystallography studies, it is likely that these same ligands are present during in vitro assays that measure CaSR activation. Further, cations identified in the crystal structures could be artefacts of the crystallization conditions and merely stabilize the crystal contacts required for structure determination. Although mutagenesis was used to corroborate the observed cation binding sites (Geng et al., 2016; Zhang et al., 2016), these mutational studies neither accounted for mutation-induced changes in receptor expression, nor quantified changes in cation affinity and efficacy. Therefore, a reduction in cation binding upon mutation of these sites has not been validated. Further, analysis of Ca2+ o binding proteins to predict the CaSR’s Ca2+ o sites, coupled with mutagenesis and spectroscopic techniques to validate the predictions, confirmed multiple VFT Ca2+ o binding sites, but they differed to those identified in the VFT crystal structures (Huang et al., 2009; Huang et al., 2007; Kirberger et al., 2008; Wang et al., 2009; Wang et al., 2010; Zhao et al., 2012). Moreover, analyses of a "headless" CaSR where the ECD has been removed, has shown that Ca2+ o can also activate the CaSR via binding sites in the 7TM domain (see Section IIIB) (Leach et al., 2016; Ray and Northup, 2002). Accordingly, the cooperative binding of Ca2+ o at multiple binding sites likely maximizes the CaSR’s ability to respond to Ca2+ o over a narrow physiological range. Additional active state structures, biophysical studies and mutagenesis work are required to fully understand how these sites interact.
Site directed mutagenesis and functional studies show that Ca2+ o and L-amino acids potentiate one another’s activity in a positively cooperative manner (Conigrave et al., 2000; Mun et al., 2005; Zhang et al., 2014a; Zhang et al., 2014b; Zhang et al., 2002b). Under physiological conditions, L-amino acids potentiate Ca2+ o potency for evoking intracellular responses (Conigrave et al., 2007), and mutating residues important for L-amino acid binding eliminated L-Phe potentiation of Ca2+ i mobilization (Conigrave et al., 2000; Zhang et al., 2002a). The ability of Ca2+ o and L-amino acids to cooperatively activate the CaSR was further demonstrated using saturation transfer difference (STD) nuclear magnetic resonance (NMR) (Zhang et al., 2014b). Using STD NMR, L-Phe was estimated to bind to the CaSR with an affinity of ~10 mM in the absence of Ca2+ o, while in the presence of Ca2+ o, L-Phe affinity was increased. Similarly, and as expected for reciprocal cooperativity, the binding affinity of Ca2+ o in the presence of 10 mM L-Phe was increased. Therefore, dual binding of Ca2+ o and amino acids enhances the sensitivity of the CaSR to changes in concentrations of these ligands.
Anion binding sites
A total of 4 anion binding sites in the inactive and active extracellular domain structures were identified based on electron density and crystallization conditions (Geng et al., 2016). Anion binding sites 1-3 are located above the interdomain cleft in lobe 1 and anion site 4 is located in lobe 2. Although SO4 2- and PO4 3- anions were modelled into these structures, it is possible other anions may be present. These anions act to stabilize the local conformation of the receptor in the crystal structure, because in the absence of PO4 3- in the inactive crystal structure, several binding site residue side chains are disordered. In the inactive structure, anions were bound at sites 1-3 whereas in the active structure, only sites 2 and 4 were occupied. In the crystal structures, anions may have stabilized the CaSR to aid crystallization. However, like all GPCRs, the CaSR can sample multiple conformations not captured in these crystal structures. Thus, under physiological conditions, anions may act to stabilize intermediate CaSR states.
Etelcalcetide binding site
The polypeptide allosteric modulator etelcalcetide binds to a distinct site in the CaSR’s VFT domain, and requires a covalent S-S bond formed directly with the CaSR VFT to retain activity (Alexander et al., 2015). This interaction occurs when the free Cys482, which is located at the back of VFT lobe 1 near the hinge loops, exchanges with a L-Cys disulphide bound to a D-Cys in the etelcalcetide D-amino acid peptide sequence. Despite this covalent linkage, the interaction appears transient and the effect of etelcalcetide on plasma PTH levels rapidly diminishes immediately after withdrawal of intravenous injection (Alexander et al., 2015). It is not known how etelcalcetide binding potentiates CaSR activity at a structural level, therefore further structural and mutagenesis studies are needed to determine the conformational changes stabilized by etelcalcetide that mediate its PAM activity.
B. CaSR 7TM domain
Structural basis of CaSR 7TM activation
The only full-length class C GPCR structure is of mGluR5, which was determined using cryo-EM. The full length mGluR5 structure shows how the inactive (or open) VFT receptor complex disrupts the interface between the 7TM domains, whereas the activated (closed) complex forces a reorientation of the 7TM domains, fostering an interface between the top of TM6 and TM7 (Koehl et al., 2019). Without a comparable structure available for the CaSR, similar conformational changes driving CaSR activation can only be hypothesized. Nevertheless, there is significant structural and functional data that is available for the CaSR 7TM that is important for understanding its activity.
Like all GPCRs, the CaSR’s 7TM helices are joined by intracellular loops 1-3 (ICL1, ICL2 and ICL3), which are important for effector coupling, and ECLs 1-3 (ECL1, ECL2 and ECL3), where ECL2 and ECL3 contain a number of residues important for receptor activation (Goolam et al., 2014; Leach et al., 2012). Structural and biochemical data for other GPCR classes show that receptor activation involves an outward movement of TM5 and TM6 to permit G protein coupling and signal transduction. 7TM movements are driven by a number of conserved amino acid sequences important for receptor activation, known as switch motifs. How this process may happen in the CaSR is discussed in this section.
Although the CaSR responds to a diverse array of stimuli through its VFT, the VFT is not required for the receptor to respond to Ca2+ o. The CaSR 7TM domain alone signals in response to Ca2+ o, albeit with lower potency and a significant reduction in the Ca2+ o Hill coefficient (Leach et al., 2016; Ray and Northup, 2002). This indicates that the CaSR 7TM also contains one or more orthosteric binding sites. Regrettably, no structures of the CaSR 7TM have been determined experimentally. However, sequence comparisons between the CaSR and mGluR1 or mGluR5 reveal that putative switch motifs important for receptor activation are shared throughout the 7TMs of Class C GPCRs, guiding our understanding of CaSR activation.
With the lack of a CaSR 7TM domain structure, the CaSR 7TM has been the subject of extensive mutagenesis and structure-function studies in an attempt to understand this domain. Guided by naturally occurring and engineered mutations and sequence homology with other GPCRs, residues important for Ca2+ o activity, allosteric modulation, biased agonism and biased modulation have been identified (Cook et al., 2015; Goolam et al., 2014; Leach et al., 2014; Leach et al., 2013). Indeed, the putative Ca2+ o binding site within the 7TM has been predicted using this approach, where Ca2+ o is hypothesized to mediate an interaction network between Glu767ECL2 and Glu8377.32 (Leach et al., 2016).
The mGluR1 and mGluR5 X-ray structures revealed an ionic lock, formed between Lys3.50 (Lys6983.50 in the CaSR), and Glu6.35 (Glu8036.35 in the CaSR) (Christopher et al., 2015; Christopher et al., 2019; Dore et al., 2014). These ionic lock residues are conserved across class C GPCRs and this “switch motif” is believed to stabilize the inactive conformation of the class C 7TM domain in the absence of agonist (Dore et al., 2014). Furthermore, a conserved sequence in class A GPCRs important for their activation called the “toggle” switch motif (protein sequence: FxxCWxP6.50), appears to be replaced by a “wl switch motif” (protein sequence: W6.50L6.51) in class C GPCRs (Trzaskowski et al., 2012). While the wl switch motif differs markedly in sequence from the class A toggle switch motif, most notably by its lack of Pro6.50 to induce a characteristic kink in TM6 (Lagerström and Schiöth, 2008), Trp6.50 in the Class C GPCR wl motif (Trp8186.50 in the CaSR) is in an identical position to Trp6.48 in the Class A GPCR FxxCWxP6.50 motif (Dore et al., 2014; Trzaskowski et al., 2012). Rotation of the Trp6.48 sidechain is a central feature of the toggle switch motif during class A GPCR activation. Molecular dynamic simulations suggest a similar rotation of Trp6.50 may occur in mGluR2 upon activation (Perez-Benito et al., 2017), while the mGluR5 crystal structures demonstrate that Trp6.50 can alternate between two distinct rotomers when bound to different NAMs, indicating it differentially orientates upon binding of different ligands (Christopher et al., 2015; Christopher et al., 2019; Dore et al., 2014). Thus, it is hypothesized that Trp6.50 in class C GPCRs fulfils an equivalent toggle switch function to Trp6.48 in class A GPCRs (Dore et al., 2014; Trzaskowski et al., 2012).
Finally, the CaSR and other class C GPCRs contain a P7.56KxY motif, which is believed to perform an analogous role to the NP7.50xxY(x)5/6F motif (where F sits 5 or 6 residues away from the Y) in Class A GPCRs. The NP7.50xxY(x)5/6F motif undergoes significant rearrangement during activation (Dore et al., 2014; Fritze et al., 2003; Katritch et al., 2013). Nevertheless, without high resolution structures of the CaSR and with only inactive mGlu1 and mGlu5 7TM structures available, it is difficult to confidently determine any importance of these motifs to CaSR activation and effector coupling.
Small molecule allosteric modulator binding sites
The CaSR 7TM contains allosteric binding sites for small molecule allosteric modulators (Bu et al., 2008; Gerspacher et al., 2010; Gregory et al., 2018; Hu et al., 2006; Josephs et al., 2019; Keller et al., 2018; Leach et al., 2016; Miedlich et al., 2004; Petrel et al., 2004; Petrel et al., 2003; Ray and Northup, 2002). These sites have been established by mutagenesis studies that examined changes in modulator potency or affinity, coupled with homology modeling to understand the context of this mutagenesis data. Initial homology modeling was based on the solved X-ray crystallography structures of class A GPCRs (Bu et al., 2008; Gerspacher et al., 2010; Hu et al., 2006; Miedlich et al., 2004), but this was later extended to modeling based on the NAM-bound 7TM structures of mGluR1 and mGluR5 (Gregory et al., 2018; Josephs et al., 2019; Keller et al., 2018; Leach et al., 2016).
Mutagenesis and homology modeling has established that the CaSR 7TM domain contains an extended allosteric binding pocket, formed by Phe6682.56, Arg6803.32, Phe6843.36, Phe6883.40, Glu767ECL2, Leu7765.43, Trp8186.50, Phe8216.53, Tyr8256.56, Val833ECL3, Ser834ECL3, Glu8377.32, Ala8407.35, Ile8417.36 and Ala8447.39 (Leach et al., 2016). This extended pocket overlaps with the allosteric and orthosteric binding sites in biogenic amine class A GPCRs (Kruse et al., 2013), and contains multiple binding sites. For instance, arylalkylamine PAMs and NAMs, such as cinacalcet and NPS 2143, are predicted to form direct salt bridge interactions with Glu8377.32 at the top of the extended binding pocket, supported by substitutions of Glu8377.32 with uncharged or positively charged amino acids, which abolish or significantly reduce arylalkylamine activity (Bu et al., 2008; Gregory et al., 2018; Jacobsen et al., 2017; Josephs et al., 2019; Keller et al., 2018; Leach et al., 2016; Miedlich et al., 2004). AC265347 is believed to bind lower in the allosteric pocket, lacking the capacity to interact with Glu8377.32 (Leach et al., 2016). Although ATF936 is predicted to bind in a comparable position to the arylalkylamines, mutation of Glu8377.32 has no effect on ATF936 potency or affinity, therefore some of its binding interactions with the CaSR differ to the arylalkylamines (Gerspacher et al., 2010; Josephs et al., 2019).
Excitingly, the established 7TM allosteric pocket is unlikely to be the only binding site for small molecule allosteric modulators. The CaSR NAM, BMS compound 1, does not appear to utilize this binding site, because it interacts in a non-competitive manner with NPS 2143 and is largely unaffected by many of the 7TM mutations that reduce the affinity of other CaSR NAMs (Arey et al., 2005; Josephs et al., 2019). Thus, there remains scope for yet uncharacterized allosteric binding to the CaSR 7TM.
Structural basis of small molecule allosteric modulator cooperativity, efficacy and bias
Fitting an operational model of agonism or allosterism to functional CaSR data has revealed structural features important for allosteric cooperativity, agonism and bias. For the PAMs cinacalcet, NPS R-568 and AC265347, mutations Glu767ECL2Ala, Val8176.49Ala or Ala8447.37Val all reduced the cooperativity of these PAMs (Keller et al., 2018; Leach et al., 2016). However, substantial differences between PAMs have also been described. For instance, while mutation of Phe6883.40Ala, Tyr8256.57Ala or Leu8487.43Ala reduced the cooperativity of the two arylalkylamine PAMs, cinacalcet and NPS R-568, mutation of Ala6151.42Val or Lys831ECL3Ala only reduced the cooperativity of cinacalcet. Furthermore, mutation of Trp8186.50Ala, which is part of the wl motif discussed above, increased cooperativity of cinacalcet, but had no significant effect on NPS R-568 cooperativity. Although structurally and pharmacologically similar, the divergent residues mediating cinacalcet or NPS R-568 cooperativity demonstrate how subtle differences in chemical scaffolds can stabilize distinct structural conformations of the CaSR 7TM domain (Keller et al., 2018; Leach et al., 2016).
The PAM agonist, AC265347, demonstrated further differences from cinacalcet and NPS R-568. For instance, unlike cinacalcet and NPS R-568, mutations Tyr8256.57Ala or Leu8487.43Ala had no effect on AC265347 cooperativity, while mutation of Phe6883.40Ala altered AC265347 cooperativity (Keller et al., 2018; Leach et al., 2016). Interestingly, AC265347 biased modulation of pERK1/2 versus Ca2+ i mobilization was altered by the mutations Leu7765.43Ala or Trp8186.50Ala. Here, these two mutations increased or decreased AC265347 cooperativity in pERK1/2 assays without altering cooperativity in Ca2+ i mobilization assays, providing some insight into 7TM residues that specifically mediated CaSR signaling towards a specific signaling pathway (Cook et al., 2015; Leach et al., 2016). Further, allosteric agonism mediated by AC265347 has different requirements to Ca2+ o agonism. While mutation of Leu7765.43Ala or V8176.49Ala reduced efficacy of both AC265347 and Ca2+ o, mutations Phe6843.36Ala or Phe6883.40Ala decreased AC265347 efficacy, without altering the efficacy or affinity of Ca2+ o (Keller et al., 2018; Leach et al., 2016).
Similar to residues that transmit cooperativity mediated by PAMs, distinct amino acids appear to transmit negative cooperativity mediate by different NAMs. For instance, of the residues analyzed to date, only the mutation Leu7765.43Ala significantly altered NPS 2143 cooperativity (Leach et al., 2016). In contrast, a number of mutations that had no effect on NPS 2143 cooperativity increased or decreased ATF936 cooperativity, including Glu767ECL2Ala, Trp8186.50Ala and Ile8417.36Ala (Josephs et al., 2019). Other NAMs were sensitive to different mutations (Josephs et al., 2019). Further analysis of additional 7TM mutations will help to unravel cooperativity networks that drive global and ligand-specific allosteric effects.
C. CaSR dimerization
Like all class C GPCRs, CaSR dimerization is a key feature governing receptor function. The dominant interaction underpinning the CaSR dimer is two covalent disulfide bonds formed at the top of lobe 1 of the VFT domains, between Cys129 and Cys131 (Ray et al., 1999). However, the CaSR is not dependent on the disulfide links for activity, evidenced by mutation of these residues to Ser, which does not alter surface expression or Ca2+ o potency in vitro (Fan et al., 1998; Zhang et al., 2001).
Dimerization influences allosteric modulation at the CaSR. For instance, negative allosteric modulators must bind both protomers to block signaling, whereas PAMs only need occupy one protomer to exert their full modulatory effect (Gregory et al., 2018; Hauache et al., 2000; Jacobsen et al., 2017). This feature likely reflects agonist-mediated signal transmission through the CaSR, which occurs across the dimer, rather than propagating through a single protomer (Hauache et al., 2000). Consequently, transactivation across the dimer can result in unique pharmacology for CaSR allosteric modulators. An example is calhex 231, which shows positive allosteric activity when bound to the allosteric site in only one protomer, but shows negative allosteric activity when occupying both the allosteric sites of the dimer (Gregory et al., 2018).
Immunoprecipitation data has demonstrated that the CaSR forms heterodimers in vitro with mGluR1/5 or the GABAB receptor, with heterodimers detected in bovine and mouse brain lysates, respectively (Chang et al., 2007; Gama et al., 2001). On the other hand, FRET studies have revealed that the CaSR does not heterodimerize with its closest receptor homolog, the GPRC6A receptor (Hauache et al., 2000; Jacobsen et al., 2017). Heterodimerization may facilitate the varied functional roles of the CaSR in different tissues, particularly in the brain where the expression of the GABAB receptor regulates CaSR expression and vice versa (discussed in Section VK).
D. CaSR glycosylation
The CaSR VFT domain contains 11 potential N-linked glycosylation sites, however not all of these sites have been experimentally verified. The CaSR is glycosylated in the endoplasmic reticulum with mannose (immature) carbohydrate prior to mature complex glycosylation processing in the Golgi. Disruption of at least three glycosylation sites can impair receptor processing and cell surface expression (Ray et al., 1998). Eight glycosylation sites (Asn90, -130, -261, -287, -446, -468, -488, and -541) have been experimentally validated; while questions remain over the three remaining sites (Asn386, -400, and -594). Notably, Asn594 was glycosylated in the solved X-ray crystal structure, while Asn386 was mutated to Gln to prevent glycosylation and aid crystallization. However, it is unclear if this observation at a truncated CaSR sample reflects the glycosylation arrangement of the full length CaSR. Importantly, the functional role of glycosylation for the CaSR’s function, beyond a function in controlling surface expression, needs further investigation.
IV. CaSR regulation
A. Phosphorylation and dephosphorylation
PKC-mediated phosphorylation of the CaSR provides a rapid and quickly reversible mechanism for inhibiting receptor activity. Indeed, treatment of parathyroid cells with PKC-activating phorbol esters overcomes the inhibitory effect of Ca2+ o on PTH release (Brown et al., 1984; Nemeth et al., 1986). When first cloned, the CaSR was predicted to contain 5 PKC consensus motifs, although 54 serine and threonine residues reside in the receptor’s C-terminal tail and ICLs (Garrett et al., 1995a). However, the key inhibitory phosphorylation site is Thr888 in the C-terminal tail (Bai et al., 1998b). Thr888 is most likely phosphorylated by PKCα (Young et al., 2014) and dephosphorylated by a calyculin A-sensitive protein phosphatase (McCormick et al., 2010).
The functional importance of inhibitory Thr888 phosphorylation is most apparent with the clinical mutant, Thr888Met, which cannot be phosphorylated. In vitro Thr888Met is a gain-of-function mutant, while it suppresses PTH secretion in vivo resulting clinically in ADH (see Section VIC) (Lazarus et al., 2011). Therefore, CaSR phosphorylation contributes significantly to CaSR activity in vivo and thus to the overall control of PTH secretion and Ca2+ o homeostasis (reviewed in (Conigrave and Ward, 2013)).
The non-phosphorylatable mutation, Thr888Val, also produced a significant gain-of-function, which was not further enhanced by co-mutation of the other 4 predicted PKC sites (Bai et al., 1998b). However, PKC inhibition at the wild-type CaSR resulted in a greater gain-of-function than produced at the Thr888Val mutant, thus it appeared likely that another unknown site may be phosphorylated in tandem with Thr888. However, the identity of this site has remained elusive. Interestingly, in mGlu5 the key PKC phosphorylation site, Ser839 (Kim et al., 2005), aligns not with Thr888 in the CaSR but with Ser875, a residue not originally predicted to be phosphorylated by PKC (Garrett et al., 1995a). Intriguingly, current data indicates removal of this putative phosphorylation site from the CaSR (Ser875Ala) also produces a gain-of-function, similar to that of Thr888Ala, while a phosphomimetic mutation at this site (Ser875Asp) produces a loss-of-function (Binmahfouz et al., 2019b). The double Ser875Ala plus Thr888Ala mutant exhibits a greater gain-of-function than Ser875Ala alone and concomitant PKC inhibition exerts no further signal enhancement. Thus, Ser875 is most likely the second major inhibitory PKC site in the CaSR (Binmahfouz et al., 2019).
Ca2+ o induces biphasic concentration-dependent phosphorylation of Thr888 in CaSR-HEK cells, with 0.5 – 2.5 mM Ca2+ o eliciting increased Thr888 phosphorylation after 10-mins, whereas 2.5 – 5 mM Ca2+ o decreases phosphorylation apparently by activating a calyculin A-sensitive protein phosphatase (McCormick et al., 2010). The decrease in Thr888 phosphorylation mediated by 2.5 – 5 mM Ca2+ o occurs at the same Ca2+ o concentrations that elicit sustained as opposed to oscillatory Ca2+ i mobilization. Consistent with this, the Thr888Ala mutant is not only gain-of-function but also exhibits less oscillatory and more sustained Ca2+ i mobilization, as does the wild-type CaSR when co-treated with a PKC inhibitor (Davies et al., 2007). Furthermore, PKC-dependent phosphorylation of Thr888 attenuates L-amino acid dependent signaling in a manner similar to its effect on Ca2+ o (Bai et al., 1998b; McCormick et al., 2010). Since PKC-mediated Thr888 phosphorylation is thus a critical regulator of CaSR function, differential CaSR phosphorylation could provide a mechanism to permit biased signaling in different cells or in response to various agonists.
CaSR signaling can also be modulated by the GPCR kinases (GRKs). Specifically, overexpression of GRK 2 and 3 decreases CaSR-induced IP formation in a HEK-derived cell-line by >70% (Lorenz et al., 2007). Mutating GRK2 so that it could no longer bind Gq overcame the inhibitory effect of GRK2 on CaSR signaling, indicating that GRK2 inhibition of CaSR signaling might be caused by sequestering of Gq rather than by phosphorylation of the CaSR. Overexpression of either β-arrestin 1 or β-arrestin 2 partly inhibits CaSR-induced IP production and this effect was abolished by deleting all 5 of the predicted PKC sites as identified by Bai et al. (1998b) (Lorenz et al., 2007).
B. Internalization and agonist-driven insertional signaling (ADIS)
In heterologous expression systems, the CaSR undergoes constitutive internalization (Gorvin et al., 2018c; Mos et al., 2019b; Reyes-Ibarra et al., 2007) and agonist-induced internalization (Lorenz et al., 2007; Nesbit et al., 2013b; Zhuang et al., 2012). Furthermore, CaSR internalization is increased by the CaSR PAM, NPS R-568, and agonist-induced but not constitutive internalization is inhibited by the NAM, NPS 2143 (Mos et al., 2019a). GPCR internalization usually involves desensitization by kinase phosphorylation and subsequent β-arrestin binding followed by recruitment to clathrin-coated pits by β-arrestin and the clathrin binding AP2 heterotetramer (Hanyaloglu and von Zastrow, 2008). As described above, phosphorylation by PKC and GRKs, or β-arrestin recruitment, is involved in CaSR desensitization (Lorenz et al., 2007; Pi et al., 2005b). Similarly, agonist-induced CaSR internalization requires β-arrestin, which is in contrast to the GABAB and mGlu receptors, which function independently of β-arrestin (Pin and Bettler, 2016). However, constitutive and agonist-induced CaSR internalization is largely independent of Gq/11 and Gi/o in HEK293 cells (Mos et al., 2019a), thus indicating that G protein-mediated activation of PKC is not involved in this cell line. Studies of the internalization mechanisms of endogenously expressed CaSRs in non-recombinant cells are still lacking due to the technical difficulty of performing such experiments.
The CaSR is also predicted to couple directly to AP2s through a dileucine motif in the CaSR C-terminal tail (Nesbit et al., 2013b). Similar to loss-of-function mutations in the CASR or GNA11 genes (Nesbit et al., 2013a; Pollak et al., 1993), mutations that disrupt the CaSR interaction with AP2s reduce the sensitivity of CaSR expressing cells to Ca2+ o (Nesbit et al., 2013b). Similarly, germline mutations of the AP2S1 gene that lead to alteration of Arg15 in AP2s cause FHH3 (Nesbit et al., 2013b), which is clinically the most severe of the three FHH types (Hannan et al., 2015a) (Section VI). AP2s Arg-15 mutations inhibit CaSR internalization (Gorvin et al., 2018c; Nesbit et al., 2013b) and the functional similarity between loss-of-function mutations in the CASR, GNA11 and AP2S1 genes shows a close relationship between internalization and CaSR signaling. This relationship could be explained by reduced resensitization and/or intracellular signaling when internalization is inhibited (Gorvin et al., 2018c; Reyes-Ibarra et al., 2007; Zhuang et al., 2012).
After internalization, GPCRs are either resensitized and recycled to the cell surface or degraded (Hanyaloglu and von Zastrow, 2008). Cell surface expression of CaSR is constant under basal conditions (Reyes-Ibarra et al., 2007; Zhuang et al., 2012), which means constitutively internalized receptors are replaced. In heterologous cells, internalized CaSR is recycled through Rab11a-dependent slow-recycling endosomes (Reyes-Ibarra et al., 2007), to be sorted to lysosomes for degradation (Grant et al., 2011; Zhuang et al., 2012). The CaSR’s C-terminal tail is involved in post-endocytic sorting, as deletion of residues 920-970 increased cell surface expression and reduced colocalization with a lysosomal marker (Zhuang et al., 2012). Similarly, overexpression of associated molecule with the SH3 domain of STAM (AMSH), which interacts with the CaSR C-terminal tail, downregulated cell surface CaSR (Herrera-Vigenor et al., 2006; Reyes-Ibarra et al., 2007).
The interaction of 14-3-3 proteins with an arginine-rich motif in the CaSR C-terminal tail partly retains an intracellular CaSR pool (regulated by Ser899 phosphorylation) (Grant et al., 2015; Grant et al., 2011; Stepanchick et al., 2010), but the CaSR is upregulated at the cell surface upon agonist stimulation via a Gq/11-dependent mechanism called agonist-driven insertional signaling (ADIS) (Gorvin et al., 2018c; Grant et al., 2011). This process involves rapid mobilization of the intracellular pool of receptors to the cell surface and initiation of receptor synthesis to support prolonged upregulation (Grant et al., 2015; Grant et al., 2011). The rapid increase in cell surface receptors is proposed to support the high sensitivity of the CaSR to increases in Ca2+ o.
V. (Patho)physiology of the CaSR and its ligands
A. CaSR in the parathyroid glands
CaSR expression appears during parathyroid development in response to key parathyroid determining genes including GCM2 (encoding Gcmb) and SHH (encoding the inhibitory controller, Sonic Hedgehog) (Grevellec et al., 2011). Consistent with a direct connection between Gcm2 and CaSR expression, GCM2 control elements have been identified in the CASR promoters (Canaff et al., 2009) and shRNA directed against GCM2 in parathyroid cell cultures suppressed the protein levels of Gcm2 and the CaSR (Mizobuchi et al., 2009).
The CaSR’s non-redundant roles in Ca2+ o metabolism have been clearly established by the hypercalcemic disorders, neonatal severe hyperparathyroidism (NSHPT) and FHH, and the hypocalcemic disorders, ADH and Bartter’s syndrome type 5. These are discussed together with animal models of loss- and gain-of-function mutations of the CaSR in Section VI.
The CaSR negatively controls parathyroid function by suppressing acute PTH secretion primarily from chief cells (review: (Conigrave, 2016)), inhibiting cell proliferation and thus cell number and gland size (Fan et al., 2018), and reducing PTH gene transcription (review: (Chen and Goodman, 2004)). It also activates the local synthesis, particularly in parathyroid oxyphil cells, of 1,25-dihydroxyvitamin D3 (Ritter et al., 2012), a recognized inhibitor of PTH synthesis. The CaSR’s effects on cell proliferation are particularly noticeable in the context of primary hyperparathyroidism (e.g., due to adenomatous disease) or hyperplasia in the context of CKD. Interestingly, in parathyroid adenoma and CKD, cellular CaSR expression is reduced (Kifor et al., 1996). Nonetheless, in CKD patients and in rat models of secondary hyperparathyroidism, sustained treatment with cinacalcet suppresses parathyroid gland size as well as serum PTH levels (Colloton et al., 2005; Yamada et al., 2015). Similarly, exposure of parathyroid cells to cinacalcet in vitro suppresses proliferation and promotes apoptosis (Tatsumi et al., 2013).
The parathyroid CaSR continuously monitors the Ca2+ o concentration as well as various other stimuli that affect CaSR function, including the plasma levels of L-amino acids (Conigrave et al., 2004), pH (Campion et al., 2015), ionic strength (Quinn et al., 1998), and, perhaps, locally generated polyamines (Quinn et al., 1997) (see Section II). CaSR activity in the parathyroid glands is resistant to desensitization, in part due to efficient receptor recycling, as well as a large intracellular receptor pool that undergoes a high rate of trafficking from the endoplasmic reticulum and Golgi to the plasma membrane (reviews: (Breitwieser, 2013; Ray, 2015)). Whether ADIS operates in parathyroid glands is unknown, but the CaSR interacts with a signaling assembly dependent on caveolin-1 (Kifor et al., 1998) and undergoes AP2s-regulated endocytosis (Gorvin et al., 2018c; Nesbit et al., 2013b).
PTH secretion control
The set-point for the CaSR’s half-maximal inhibitory effect on PTH secretion lies at the lower limit of the normal free Ca2+ o concentration range (1.1-1.3 mM). In this way, the disinhibited parathyroid provides the body’s primary defense against hypocalcemia (Ca2+ o < 1.1 mM). However, the parathyroid CaSR does not provide the primary defense against hypercalcemia, which is mediated by CaSRs in the renal cortical thick ascending limbs (cTAL) of Henle’s loop, which accelerate Ca2+ o excretion (Kantham et al., 2009; Loupy et al., 2012). Further, as Ca2+ o increases, its inhibitory effect on PTH secretion suppresses bone resorption.
Two distinct paradigms for CaSR-mediated inhibition of PTH secretion have been identified at the cellular level: (i) Stimulation of Gi/o proteins, which oppose cAMP-dependent increases in PTH secretion mediated by Gs-coupled receptors, e.g., for adrenaline (β-adrenergic receptors 1 and 2), dopamine, histamine, and prostanoids (review: (Conigrave, 2016)); (ii) Inhibition of endogenous PTH secretion mechanisms that occur in the absence of exogenous activators, at least in part via stimulation of Gq/11 proteins. Endogenous PTH secretion mechanisms may depend upon the intrinsic production of activators for parathyroid secretion (review: (Conigrave, 2016)), or may represent true constitutive secretion (Muresan and MacGregor, 1994).
CaSR structure and function in the parathyroid
The primary protein form adopted by the CaSR in parathyroid cells appears to be disulphide-linked homodimers (Kifor et al., 2003) similar to those observed in HEK293 cells when it is expressed heterologously (Bai et al., 1998a). However, it may also form heterodimers with some other receptors e.g., GABAB receptors as reported for growth plate chondrocytes (Cheng et al., 2007) with unknown consequences for parathyroid cell signaling and function (see Section IIIC).
The CaSR in the parathyroid couples to various signaling pathways as it does when expressed heterologously in HEK293 cells, and in other cell types (review: (Conigrave and Ward, 2013)). In parathyroid cells the CaSR couples to multiple heterotrimeric G proteins including most notably Gi/o and Gq/11. Of these, Gi/o supports Ca2+ o-mediated suppression of PTH secretion stimulated by Gs-coupled receptors (e.g., for dopamine (Brown et al., 1990)) but not intrinsic PTH secretion (Brown et al., 1992). Of apparently greater importance, Gq/11 signaling is absolutely required for CaSR-mediated control of PTH secretion. Thus, mice that are global null for Gα 11 or have conditional deletion of Gα q in the parathyroid exhibit mild-moderate hyperparathyroidism. Interestingly, however, cross-breeding to generate mice that are both global null for Gα 11 and parathyroid null for Gα q results in severe neonatal hyperparathyroidism (Wettschureck et al., 2007) that is comparable to that seen in human neonates with homozygous or compound heterozygous loss-of-function CaSR mutations (Pollak et al., 1993). Consistent with these observations, loss-of-function mutations of Gα 11, which only partially impair signaling, have been linked to a variant form of FHH in humans, now known as FHH2, and gain-of-function mutations of Gα 11 have been linked to a variant form of ADH, now known as ADH2 (Nesbit et al., 2013a). The Arg60Cys and Ile62Val gain-of-function mutations in Gα 11 also induce ADH2 in mice, in which treatment with the NAM, NPS 2143, or the specific Gq/11 inhibitor, YM-254890, increases PTH and Ca2+ o concentrations (Gorvin et al., 2017; Roszko et al., 2017). These findings demonstrate the critical importance of Gq/11 in control of PTH synthesis and/or secretion. CaSR signaling in the parathyroid is also negatively regulated by the GTPase activator RGS5, and overexpression of RGS5 in the parathyroid induces hyperparathyroidism in mice (Koh et al., 2011). Whether RGS5 has a preference for either Gq or G11 in parathyroid cells is unknown. Interestingly, studies in other tissues suggest that RGS5 preferentially suppresses the function of Gq with little or no effect on G11 (Ladds et al., 2007). Whether this might support a parathyroid-based preference for CaSR-mediated activation of G11 rather than Gq is unknown.
The mechanism by which the CaSR controls PTH secretion downstream of Gq/11 is surprisingly ill-defined. Contributing factors appear to include PI-PLC, which is robustly activated by Ca2+ o stimulation in parathyroid cells (Brown et al., 1987a; Shoback et al., 1988), Ca2+ i signals whose frequency and amplitude are dependent on the phosphorylation status of Thr888 (Lazarus et al., 2011; McCormick et al., 2010), and the MAPK, ERK1/2 (Corbetta et al., 2002). Evidence has also been presented for a convergent signaling pathway mediated by α-klotho and the CaSR on PTH synthesis and parathyroid hyperplasia downstream of FGF receptors (Fan et al., 2018). In other work, parathyroid Na+/K+-ATPase activity has been implicated in the control mechanism (Brown et al., 1987b; Imura et al., 2007). Whether this might operate via changes in cell volume or intracellular ion concentrations is unclear; changes in Ca2+ i concentration appear to have been excluded (Brown et al., 1987b).
B. CaSR in the thyroid gland
In the thyroid the CaSR is expressed at high levels in a relatively small subpopulation of cells, the parafollicular C cells (Garrett et al., 1995b). In C cells, the CaSR acts to promote secretion of the peptide hormone calcitonin. Evidence that the CaSR stimulates calcitonin secretion is supported by studies in CaSR knockout mice in which plasma calcitonin levels were suppressed (Fudge and Kovacs, 2004; Kantham et al., 2009). Thus, elevated Ca2+ o stimulates calcitonin release, which, in turn, lowers the plasma calcium level, primarily by suppressing bone resorption. Both the CaSR and calcitonin (or calcitonin gene related peptide) genes are under inhibitory regulation by thyroid transcription factor-1 (TTF-1) in C cells, and CaSR activation promotes calcitonin synthesis, at least in part, by suppressing the levels of TTF-1 (Suzuki et al., 1998). CaSR coupling to G proteins in C cells is similar to its coupling in parathyroid cells and various other cell-types, and the C cell CaSR thereby activates plasma membrane phospholipases and cellular protein kinases (McGehee et al., 1997). Activation of the CaSR also stimulates acute elevations in Ca2+ i in various C cell models. In some cell types, this occurs via Ca2+ o entry through plasma membrane L-type voltage-gated Ca2+ channels (Fajtova et al., 1991; McGehee et al., 1997; Muff et al., 1988) and, in others, via Ca2+ i mobilization (Freichel et al., 1996). Recently, the amino acid activated CaSR was shown to stimulate calcitonin release from human C-cells (Mun et al., 2019). Despite these insights, the molecular mechanisms by which the CaSR stimulates calcitonin secretion are largely unknown.
C. CaSR in the kidney
CaSR expression in the kidney is one of the highest in the body and the renal CaSR plays a major role in the regulation of renal function in both a hormone-dependent and independent fashion (see Riccardi and Valenti, 2016 and references therein). A large body of functional, molecular and genetic evidence indicates that the kidney CaSR plays a crucial role in mineral ion homeostasis. Indeed, the CaSR is widely expressed along the nephron, at both the apical and basolateral sides of kidney cells, thereby being uniquely poised to monitor both urine and plasma and alter the final ultrafiltrate composition accordingly (Graca et al., 2016; Riccardi et al., 1998). Urinary calcium excretion mirrors serum calcium levels and is directly proportional to the filtered calcium load (Brown, 1991). Within the kidney, the thick ascending limb (TAL) of Henle’s loop is the main site for active divalent cation movement, mostly via the paracellular route, and is coupled to NaCl reabsorption (Friedman, 1998). The latter occurs through a concerted action of an apical Na+:K+:2Cl- cotransporter, NKCC2, followed by basolateral exit via the voltage-gated Cl- channel, CLCNKB, and the Na+:K+:ATPase. Overall, NaCl movement generates a favorable transepithelial electrochemical gradient for positively charged ions to move from the urine towards the basolateral side. In concert, the tight junctional proteins, claudins 14, 16 and 19, establish a divalent cation-selective permeable route thereby allowing Ca2+ o (and magnesium) reabsorption (Gong and Hou, 2014). The TAL has the highest CaSR expression, and here the CaSR is expressed basolaterally (Riccardi et al., 1998). In the event of hypercalcemia, CaSR activation dampens Ca2+ o reabsorption in two ways: firstly, by inhibiting NaCl reabsorption, hence the driving force for divalent cation movement; secondly, by directly reducing Ca2+ o and magnesium junctional permeability through its actions on claudin 14 by activating microRNA-9 and -374 (Gong and Hou, 2014). If the hypercalcemic stimulus persists hypercalciuria can occur, with excess urinary calcium excretion in the terminal collecting duct.
In the presence of hypovolemia, the anti-diuretic hormone, vasopressin, promotes water reabsorption through the insertion of aquaporin-2 water channels into the lumen of inner medullary collecting duct cells. However, excessive water reabsorption could lead to supra-saturating urinary calcium concentrations and attendant pathological kidney stone formation, which could severely impair renal function. The CaSR is expressed at the luminal side of inner medullary collecting duct cells where it monitors Ca2+ o concentration in the urine (Sands et al., 1997). Thus, CaSR activation inhibits the tubular response to vasopressin by limiting the number of apical aquaporin 2 water channel insertions (Procino et al., 2012). In addition, CaSR activation stimulates the activity of the proton pump, V-ATPase, therefore evoking urine acidification and reducing the risk of precipitation (Renkema et al., 2009).
Further, the kidney proximal tubule is a major site of PTH action, which promotes a phosphaturia by inhibiting the activity of the Na+:Pi cotransporters, Npt2a and Npt2c (Murer et al., 2001). Excess phosphate in the urine could also exacerbate the risk of calcium-phosphorus stone formation by the distal nephron. In the proximal tubule, a luminal CaSR blunts the phosphaturic action of PTH and promotes acid secretion via stimulation of the Na+:H+ exchanger, NHE3 (Capasso et al., 2013). Thus, by monitoring both urine and plasma composition, together with the integration of inputs deriving from urinary phosphate content, concentration and acidification, the renal CaSR accomplishes divalent cation homeostasis while minimizing the risk of developing nephrolithiasis and nephrocalcinosis that could arise as a consequence of enhanced urinary calcium excretion by the TAL (Hebert et al., 1997). The corollary is that altered CaSR expression or function due to CaSR mutations leads to FHH, ADH or Bartter type V syndrome (see Section VI). In all circumstances, the aberrant calciuria is not the consequence of an impairment of renal function but rather the result of altered Ca2+ o sensing by the CaSR in the parathyroid glands and the kidney. In the context of CKD, hyperphosphatemia caused by decreased renal phosphate excretion, and acidosis, may both elicit CaSR underactivation, leading to secondary hyperparathyroidism. Therefore, similarly to the parathyroid CaSR, the kidney CaSR is a drug target and, indeed, pharmacological CaSR PAMs are employed to rectify abnormal Ca2+ o sensing by the kidney (Riccardi and Valenti, 2016). Furthermore, the use of NAMs for the treatment of nephrolithiasis and nephrocalcinosis could also be postulated (Riccardi and Valenti, 2016). Finally, it should be noted that CaSR PAMs increase urinary calcium excretion by means of their actions on both the parathyroid and kidney CaSR and indeed, cinacalcet promoted calciuria in patients with secondary hyperparathyroidism, but this occurs in the absence of an increase in urine output (Courbebaisse et al., 2012). Given the clinical use of PAMs, the impact of their long-term use on urine production, acidification and concentration, particularly in the context of kidney stone formation, remains to be fully understood (Riccardi and Valenti, 2016).
D. CaSR in the bone
The CaSR is expressed by several types of bone cells, including osteoblasts, osteocytes, osteoclasts and some chondrocytes (Santa Maria et al., 2016). Although some controversies exist, there is good evidence that Ca2+ o and the CaSR contribute to skeletal development and maintenance (Chang et al., 2008; Goltzman and Hendy, 2015; Hannan et al., 2018a) and that bone CaSRs may even contribute to overall Ca2+ o homeostasis (Al-Dujaili et al., 2016). Elucidation of the CaSR’s roles in skeletal tissue was historically complicated by models that examined global deletion of the Casr gene (Kos et al., 2003). Global Casr deletion has numerous effects, partly through large alterations in PTH secretion and changes in serum calcium and phosphate concentrations (Hannan et al., 2018a), thus it is difficult to elucidate tissue-specific CaSR effects. To further complicate matters, early Casr knockouts involved deletion of Casr exon 5, which results in mice encoding a non-functional CaSR lacking a portion of its extracellular domain (Kos et al., 2003). When Casr exon 5 deleted mice were crossed with mice that had a deletion of Gcm2 (which results in no parathyroid gland development) or the Pth gene, the skeletal abnormalities seen in the global Casr knockout mice were largely abolished (Kos et al., 2003; Tu et al., 2003). Further, studies of the Casr exon 5 deleted mice revealed an alternatively spliced Casr transcript in the growth plate and other organs such as skin that could compensate for the absence of full-length CaSR in cartilage and bone (Rodriguez et al., 2005). Nonetheless, alternative Casr knockout models and bone-specific Casr deletion has confirmed that the CaSR is critical to bone development and maintenance, as described below.
Osteoblast CaSRs
Perhaps the clearest evidence for a role of the CaSR in skeletal development and maintenance comes from studies in which exon 7 of the Casr gene was deleted during different stages of osteoblast differentiation. Casr exon 7 deletion removes most of the 7TM and C-terminal tail resulting in a non-functional receptor (Chang et al., 2008; Dvorak-Ewell et al., 2011). Casr exon 7 deletion was achieved by Cre-recombinase in osteoblasts under the control of the 2.3-kb Col(I) α1 subunit promoter [2.3Col(I)-Cre], which is expressed in early- and late-stage cells of the osteoblast lineage (Chang et al., 2008), an α1(I) collagen promoter [Col 3.6–Cre], which is expressed throughout cells of the osteoblastic lineage (Dvorak-Ewell et al., 2011), or the osterix promoter, which is expressed in early osteoblasts (Chang et al., 2008). In studies using the collagen-Cre promoters, heterozygous Casr knockout mice grew relatively normally (Chang et al., 2008; Dvorak-Ewell et al., 2011). In contrast, homozygous Casr knockout using any of the Cre promoters resulted in severe bone defects (Chang et al., 2008; Dvorak-Ewell et al., 2011). There was marked reduction in the size of the knockout mice and their skeleton, evident as early as 3 days after birth, and at day 20 the weight of the Casr knockout mice was only 30% that of controls (Chang et al., 2008; Dvorak-Ewell et al., 2011). Their skeletons were severely undermineralized, even in the skull, as well as in the vertebrae and long bones (Chang et al., 2008; Dvorak-Ewell et al., 2011). There was a marked reduction in bone volume in both the trabecular and cortical bones (Chang et al., 2008). Most of these mice died with multiple fractures by 3-4 weeks after birth (Chang et al., 2008; Dvorak-Ewell et al., 2011). Osteoblasts from Casr knockout mice were poorly differentiated, with both early and late differentiation markers markedly reduced, compared with controls (Chang et al., 2008; Dvorak-Ewell et al., 2011). mRNA levels of the local growth factor, IGF-1, were also substantially decreased, as were those of factors supporting cell survival, such as Bcl-2 and Bcl-2L1 (Chang et al., 2008). In contrast, mRNA for IL-10, an inducer of apoptosis in many cell types, was increased, along with evidence of augmented apoptotic osteoblast and osteocyte numbers in sections from bone (Chang et al., 2008; Dvorak-Ewell et al., 2011). mRNA encoding genes that inhibit mineralization, such as osteopontin (OPN), ankylosis protein (ANK) and nucleotide pyrophosphatase/phosphodiesterase 1 (NPP1), showed increased expression in the knockouts (Dvorak-Ewell et al., 2011). In addition to impaired osteoblastic differentiation and activity, deletion of Casr in early and late osteoblasts led to increased expression of mRNA for the bone resorption promoting protein, Receptor Activator of NFkB Ligand (RANK-L), together with a doubling of osteoclast numbers and activity, with bone loss in trabecular and cortical bone (Chang et al., 2008; Dvorak-Ewell et al., 2011). Whether these effects of osteoblast Casr knockout are entirely specific is not clear, since transplantation of vertebrae from 10-day old wild type or homozygous Casr knockout mice into athymice mice, resulted in no differences in the volume or composition of transplanted bones, assessed by microCT or histomorphometry after 4 weeks (Al-Dujaili et al., 2016). Furthermore, bones of transgenic mice that expressed a constitutively active mutant CaSR in late osteoblasts under the control of the osteocalcin promoter (Dvorak et al., 2007), also showed increased RANK-L expression and increased osteoclast activity, with resultant bone loss over the lifespan of the mice, but only in trabecular and not cortical bone (Dvorak et al., 2007). Other osteoblastic markers and function were largely unaffected, except for a slight decrease in bone-forming activity, indicated by a small drop in mineral apposition rate (Dvorak et al., 2007). It is difficult to explain these observations in a comprehensive model of the CaSR’s role in bone, despite attempts to propose age-related differences in the interactions between the CaSR and PTH/PTH1R in bone or the presence of mild hyperparathyroidism in the CaSR knockout models, but not in the mutant constitutively active CaSR mice (Dvorak et al., 2007; Goltzman and Hendy, 2015). Under conditions of expression of a constitutively active CaSR, however, normal feedback mechanisms would not function. The existing evidence indicates that Ca2+ o and the CaSR, together with PTH/PTHrP and PTH1R interact with one another in whole animals in ways that cannot easily be predicted (Goltzman and Hendy, 2015; Santa Maria et al., 2016; Yang and Wang, 2018).
Osteoclast CaSRs
Local regulatory pathways relevant to Ca2+ o and the CaSR are likely to involve osteoclasts, which, along with bone marrow monocytes and macrophages, express the CaSR (Diepenhorst et al., 2018; House et al., 1997; Kameda et al., 1998). Activation of the CaSR in these cells with high concentrations of Ca2+ o or strontium inhibited osteoclast maturation and secretion of acid phosphatase, critical for bone resorption, and increased apoptosis of mature osteoclasts, all of which would suppress bone resorption (Diepenhorst et al., 2018; Kameda et al., 1998; Kanatani et al., 1999; Mentaverri et al., 2006; Zaidi et al., 1991). Although high Ca2+ o concentrations were required to activate these osteoclast responses, this might be relevant in vivo, since an acid environment, as present in resorption pits, increases Ca2+ o potency at the CaSR (Quinn et al., 2004), and the Ca2+ o concentration in bone resorption pits can be as high as 40 mM (Silver et al., 1988). There is some recent evidence that cinacalcet can inhibit the actions of osteoclasts (Diepenhorst et al., 2018), raising the possibility of CaSR activation in osteoclasts as a potential anti-resorptive strategy in osteoporosis. However, another study found no effect of cinacalcet on osteoclast-mediated resorption (Shalhoub et al., 2003).
Osteoblast and osteoclast CaSRs as therapeutic targets
Given the negative effects of CaSR deletion on bone mass and bone cell survival (Chang et al., 2008; Dvorak-Ewell et al., 2011; Santa Maria et al., 2016), it follows that there would be interest in targeting the CaSR in osteoblasts/osteocytes for a bone anabolic effect (Diepenhorst et al., 2018; Goltzman and Hendy, 2015; Marie, 2010). Indeed, there is evidence that strontium, which displays higher potency than Ca2+ o in osteoblasts (Brennan et al., 2009), increased bone mineral density and reduced fractures in the clinic (Reginster et al., 2005). Other receptors including GPRC6A may also mediate the effects of strontium (Pi et al., 2005a; Rybchyn et al., 2009). Unfortunately, reported cardiovascular side effects of strontium ranelate (marketed as Protelos/Osseor) narrowed the potential patient population, so that this agent was withdrawn from the market. Nevertheless, pre-clinical studies showed strontium reduced bone resorbing signals and increased bone cell anabolism and survival under stress (Bonnelye et al., 2008; Brennan et al., 2009; Rybchyn et al., 2011). They also reported that strontium stimulated the important bone anabolic Wnt pathway downstream of the CaSR and Akt phosphorylation in osteoblasts (Rybchyn et al., 2011). CaSR-dependent activation of the Wnt pathway in bone cells was in turn dependent on the formation of a complex involving CaSR, Homer1 (a long isoform of this scaffold protein) and mechanistic target of rapamycin (mTOR) complex-2 (mTORC2), which phosphorylates Akt on Serine 475 (Rybchyn et al., 2019). These observations provide proof of principle that selective activation of the CaSR in osteoblasts might be a suitable strategy for osteoporosis therapies, either alone or in combination with other anabolic agents such as intermittent PTH. Intermittent PTH has anabolic effects on bone but also stimulates osteoclast activity. Given that CaSR activation in osteoclasts suppresses bone resorption, as discussed above, and has anabolic effects on bone, the use of CaSR PAMs in conjunction with intermittent PTH, may reduce the likelihood of hypercalcemia and enhance the bone anabolic effects of intermittent PTH. Indeed, administration of the PAM, NPS R-568, in combination with intermittent PTH in mice, reduced blood Ca2+ o concentrations, increased trabecular bone and increased cortical bone strength compared with intermittent PTH alone (Santa Maria et al., 2015). The effect of intermittent PTH on trabecular bone volume as a fraction of total bone volume was slightly, but significantly, blunted in mice in which the Casr gene was deleted in early and late osteoblasts (Al-Dujaili et al., 2016).
Chondrocyte CaSRs
High levels of CaSR protein are present in hypertrophic chondrocytes in the growth plate of long bones (Santa Maria et al., 2016). When mice with the loxP sites flanking exon 7 of the Casr were crossed with mice expressing the Cre transgene under the control of the type II collagen α1 subunit [Col(II)] promoter [Col(II)-Cre], which targets growth plate chondrocytes and other types of chondrocytes, they all died in utero at around E13 (Chang et al., 2008). Whether this was due to interference in heart valve development is unclear. When the Col(II)-Cre promoter was modified to a tamoxifen-inducible variant, and 4-hydroxytamoxifen was given at E18-19, the resultant growth plate chondrocyte targeted Casr knockout produced small, undermineralized skeletons, with expansion and reduced mineralization of the hypertrophic zone of the growth plate (Chang et al., 2008). Gene expression analysis confirmed reduced expression of terminal differentiation markers and reduced expression of IGF-1 and its receptor (Chang et al., 2008).
The CaSR is also expressed in articular cartilage chondrocytes, with increased expression reported in chondrocytes from osteoarthritic joints (Burton et al., 2005). Increased expression of the CaSR was also reported in cartilage endplate chondrocytes adjacent to degenerated intervertebral discs from human subjects along with high total calcium concentrations and low water content (Grant et al., 2016). Treatment of cartilage endplate chondrocytes in vitro with increasing Ca2+ o, resulted in lower accumulation of collagens I and II and aggrecan, while catabolic enzymes were increased, effects that were abrogated by knockdown of the CASR (Grant et al., 2016). The authors proposed a role for increased Ca2+ o and the CaSR in intervertebral disc degeneration (Grant et al., 2016). In a dental malocclusion model affecting the temporomandibular joint in rats, increased expression of CaSR in articular chondrocytes and in the endoplasmic reticulum of these cells was also observed (Zhang et al., 2019). These provided some evidence to support a role for endoplasmic reticulum expressed CaSR, as opposed to cell membrane CaSR in chondrocytes under stress (Zhang et al., 2018). Increased whole cell CaSR and increased endoplasmic reticulum CaSR were also observed in vitro using articular cartilage chondrocytes exposed to shear stress (Zhang et al., 2018). Shear stress resulted in increased expression of chondrocyte terminal differentiation markers such as alkaline phosphatase, osteocalcin and matrix metalloprotease-13, which contributes to cartilage degradation. Critically, local CaSR knockdown, or the use of the NAM, NPS 2143, reduced the shear stress-induced increases in terminal differentiation markers in chondrocytes in culture and reduced the severity of osteoarthritis in the temporomandibular joint of a rat model of dental malocclusion (Zhang et al., 2018). In contrast, injection of the PAM, cinacalcet, into the temporomandibular joint of these rats promoted thinning and loss of articular cartilage (Zhang et al., 2018). These studies in chondrocytes raise the possibility that the chondrocyte CaSR is a potential therapeutic target for prevention or management of joint degeneration.
E. CaSR in keratinocytes
The CaSR is highly expressed in keratinocytes, the main epidermal cell type. Moreover, an ionic calcium gradient exists in the epidermis, which increases from the basal proliferative layer to reach a maximum in the stratum granulosum, where the keratinocytes are well differentiated, decreasing again in the relatively water deficient lipid-containing cells of the stratum corneum (Celli et al., 2010; Menon et al., 1985). The epidermal calcium gradient and the CaSR are critically important for various epidermal functions including keratinocyte differentiation, water and xenobiotic barrier function and wound healing (Hannan et al., 2018a; Tu and Bikle, 2013). Interestingly, the epidermal calcium gradient is predominantly present in intracellular organelles of keratinocytes, such as the endoplasmic reticulum and Golgi, although an extracellular gradient makes some contribution to the gradient (Celli et al., 2010).
Keratinocytes cultured in low calcium media (<0.07 mM) proliferate well. Raising the Ca2+ o concentration above 0.1 mM promotes differentiation, indicated by the appearance of E-Cadherin/catenin complexes (adherens junctions) and desmosomes, upregulation of keratins 1 and 10, stratification of cells and then formation of cornified envelope precursors (Braga et al., 1995). Disruption of the permeability barrier of the skin by tape stripping, disrupts the epidermal calcium gradient resulting in disorganization of the normally differentiated cell layers (Menon et al., 1994). When the calcium gradient is re-established over the next day or so, the permeability barrier also recovers. Skin diseases such as psoriasis, characterized by abnormal barrier function, also exhibit a loss of the calcium gradient (Menon and Elias, 1991).
CaSR expression increases in upper layers of the epidermis with the increase in differentiation, with high expression in the stratum granulosum, but weak expression in the corneocytes (Komuves et al., 2002). There is some expression of the CaSR on the plasma membrane of keratinocytes, but its predominant localization in these cells is intracellular, in the cytoplasm around the nucleus (Komuves et al., 2002). This perinuclear localization is also seen, though not to the same extent, in rodent osteoblasts and chondrocytes (Chang et al., 1999). It is likely that Ca2+ o signals to the keratinocyte via the plasma membrane CaSR, in a manner similar to that of more classic calcium targets such as the parathyroid or kidney. The function of the intracellular CaSR is unclear at this time. Knockdown or inactivation of the CaSR in keratinocytes abrogates calcium-induced inhibition of proliferation and stimulation of differentiation of these cells (Tu et al., 2008). Not surprisingly, in mice in which there had been knockdown of the CaSR in the epidermis, skin barrier function was disrupted with impaired differentiation of keratinocytes, and these problems were exacerbated by a low calcium diet (Tu et al., 2012). Keratinocytes from these epidermal Casr -/- mice had blunted Ca2+ i mobilization in response to Ca2+ o, decreased Ca2+ i pools, defective cell-cell adhesion and reduced expression of differentiation markers (Tu et al., 2012).
In contrast, mice engineered to constitutively overexpress the CaSR in basal keratinocytes, displayed enhanced keratinocyte differentiation and barrier formation during development as well as accelerated hair growth at birth (Turksen and Troy, 2003). There was hypertrophy of the suprabasal keratinocyte layers with increased expression of early and late differentiation markers, together with upregulation of epidermal growth factor and non-canonical Wnt signaling pathways (Turksen and Troy, 2003).
In the epidermis, there are interactions between the CaSR and the vitamin D system in skin. The active vitamin D hormone, 1,25-dihydroxyvitamin D, increases the calcium response in keratinocytes (Ratnam et al., 1999). Deletion of the epidermal CaSR reduces expression of both the vitamin D receptor (VDR) and CYP27B1, the enzyme that produces 1,25-dihydroxyvitamin D from 25-hydroxyvitamin D (Tu et al., 2012). It is likely that these effects on the vitamin D system contribute to impaired differentiation of the epidermis in these mice and reduced function of the innate immune system (Schauber et al., 2007). Moreover, 1,25-dihydroxyvitamin D increases transcription of the CASR (Canaff and Hendy, 2002).
It has previously been reported that wound healing is impaired in mice with epidermal deletion of the VDR (Oda et al., 2017). Very low dietary calcium or deletion of the Casr gene exacerbates this impairment in wound healing in epidermal VDR-deficient mice (Oda et al., 2017). There is a robust increase in Ca2+ in the bed of wounds within minutes of injury (Jungman et al., 2012), a rapid increase in Ca2+ i in cells near the site of the wound and spreading to surrounding cells (Tsutsumi et al., 2013), all indicating that the CaSR may play an important role in wound healing. The CaSR is co-expressed with E-cadherin at the cell membranes of migratory keratinocytes. Blockade of either the CaSR or E-cadherin inhibited keratinocyte proliferation and migration after wound induction (Tu et al., 2019). Accordingly, the PAM, NPS R-568, accelerated wound healing in normal mice, potentially pointing to the epidermal CaSR as a therapeutic target to enhance repair of skin wound (Tu et al., 2019).
Mice with epidermal knock-out of the VDR are more susceptible to UV- or chemically-induced skin tumor formation (Ellison et al., 2008; Zinser et al., 2002). Neither the epidermal VDR knockout mice nor mice with epidermal Casr knockout develop skin tumors spontaneously, but mice null for both epidermal Vdr and Casr are reported to spontaneously develop squamous cell carcinomas (Bikle et al., 2015). In keratinocytes, stimulation of Wnt signaling results in β-catenin translocation to the nucleus and subsequent transcriptional activity, which may be important in skin tumorigenesis (Wei et al., 2007; Youssef et al., 2012). The VDR appears to suppress this β-catenin transcriptional activity in skin (Wei et al., 2007), in part by helping to keep β-catenin at the cell membrane as part of the E-cadherin/catenin complex (adherens junctions). As noted earlier, the CaSR is also important for the development of the E-cadherin/catenin complex, which helps to retain β-catenin at the cell membrane, promoting wound healing and differentiation of the skin cells (Oda et al., 2017) and inhibiting nuclear translocation and associated pro-tumorigenic activities of β-catenin (Wei et al., 2007). This is in direct contrast with osteoblasts, where activation of the CaSR predominantly promotes β-catenin stabilization and subsequent β-catenin translocation to the nucleus and increased transcriptional activity (Rybchyn et al., 2011). Some preliminary data indicates that both CaSR PAMs and NAMs enhance DNA repair after UV damage in cultured keratinocytes (Yang et al., 2016), though the mechanism, and why PAMs and NAMs have a similar effect is unclear. How this observation fits with observed effects of CaSR knockdown in mice, remains to be examined.
F. CaSR in the gastrointestinal (GI) tract
The CaSR is expressed along the entire GI tract, in parietal and G cells of stomach gastric glands (Busque et al., 2005; Engelstoft et al., 2013; Feng et al., 2010; Ray et al., 1997), epithelial and entero-endocrine cells of the small and large intestine (Alamshah et al., 2017; Cheng et al., 2014; Liou et al., 2011; Wang et al., 2011), and neurons of the submucosal and myenteric plexuses of the enteric nervous system (ENS) (Cheng, 2012; Geibel et al., 2006; Tang et al., 2018). In the GI tract the CaSR functions as a nutrient sensor, binding not only Ca2+, Mg2+, and other cations, but also L-amino acids, and di- and polypeptides (e.g. glutamyl dipeptides, poly-L-lysine). The CaSR is involved in regulation of gastric acid and hormone secretion, nutrient absorption, intestinal fluid homeostasis, energy metabolism, cellular differentiation and proliferation, motility and enteric nerve activity, maintenance of gut microbiota, immune homeostasis, and intestinal inflammation (Alamshah et al., 2017; Brennan et al., 2014; Ceglia et al., 2009; Cheng et al., 2014; Dufner et al., 2005; Feng et al., 2010; Geibel and Hebert, 2009; Sun et al., 2018; Tang et al., 2016b; Tang et al., 2018).
The CaSR responds to alterations in nutrient levels by regulating hormone secretion from entero-endocrine cells (Alamshah et al., 2017; Geibel and Hebert, 2009; Liou et al., 2011; Liu et al., 2018; Wang et al., 2011). In global Casr knockout mice, gastric G cell number was significantly reduced, suggesting the CaSR regulates G cell growth. Further, in wildtype but not knockout mice, NPS 2143 inhibited gastrin secretion following gavage of Ca2+ o, L-Phe or cinacalcet (Feng et al., 2010). In rat whole stomach preparations, ex vivo exposure to Ca2+ o increased acid production in parietal cells by enhancing H+-K+-ATPase activity. These effects were potentiated by L-but not D-amino acids, implicating the CaSR (Busque et al., 2005). The function of the recently identified acid secretory protein, vacuolar H+-ATPase, in parietal cells, is also dependent on CaSR activity (Kitay et al., 2018).
The amino acid-stimulated CaSR may influence appetite and satiety via stimulatory effects on satiety hormones, cholecystokinin (CCK), glucagon-like peptide 1 (GLP-1) and protein YY (PYY) (Alamshah et al., 2017) and/or inhibitory effects on the release of the appetite-stimulating hormone ghrelin (Engelstoft et al., 2013). In the mouse entero-endocrine cell line, STC-1, L-Phe increased PYY and GLP-1 secretion, an effect inhibited by NPS 2143, suggesting involvement of the CaSR (Alamshah et al., 2017). In a transgenic mouse model expressing a CCK promoter-driven enhanced green fluorescent protein (GFP), the CaSR was enriched in CCK-producing duodenal I cells, where L-Phe and cinacalcet induced Ca2+ i changes and stimulated CCK in the presence of Ca2+ o (Liou et al., 2011). L-Phe and Trp-stimulated CCK secretion was inhibited by calhex 231 (Wang et al., 2011). In intestinal L-cells the CaSR was involved in peptone-stimulated GLP-1 release (Pais et al., 2016), while in a ghrelinoma cell line the CaSR partially mediated the L-Phe, L-Ala, and peptone-induced secretion of octanoyl ghrelin (Vancleef et al., 2015). In swine duodenum, ex vivo L-Trp perfusion induced secretion of CCK and glucose-dependent insulinotropic peptide and upregulated CaSR expression. This effect was inhibited by NPS 2143 (Zhao et al., 2018). To conclude that the Trp-induced gut hormone secretion is mediated by the CaSR needs further proof. In mice and rats L-Phe reduced short-time food intake and plasma ghrelin release, affecting the appetite of the animals (Alamshah et al., 2017). In minks, CaSR-mediated secretion of CCK and PYY led to emesis (Wu et al., 2017). These findings might explain why cinacalcet and other CaSR PAMs cause gastrointestinal side effects (Block et al., 2017).
Specific knockout of intestinal epithelial cell Casr leads to epithelial cell hyperproliferation, and changes in intestinal crypt structure, driven by β-catenin signaling (Rey et al., 2012). Mice additionally have decreased intestinal transepithelial resistance and reduced levels of colonic tight junction proteins, suggesting that the epithelial CaSR maintains intestinal barrier function (Cheng et al., 2014). The inadequate epithelial barrier function was associated with lower amounts of beneficial lactobacilli, and more Deferribacteraceae bacteria, which are linked to colitis (Cheng et al., 2014). This intestinal dysbiosis has been associated with more severe pro-inflammatory responses in the intestinal epithelium-specific Casr null mice, compared with wild-type controls (Owen et al., 2016).
The amino acid-stimulated CaSR has recently been found to suppress intestinal inflammation in inflammatory bowel disease and other settings (reviews: (Owen et al., 2016; Sun et al., 2018)). Inflammatory cytokine expression, including IL-1R, was higher in the distal colons of the CaSR knockout mice, in addition to a marked increase in NFκB-dependent genes (Cheng et al., 2014). These mice developed more severe colitis with delayed recovery than the CaSR-expressing littermates when challenged with dextrane sulphate sodium (DSS) (Cheng et al., 2014). In a mouse model of colitis, poly-L-lysine (commonly used as a food preservative) and glutamyl dipeptides reduced DSS-induced inflammation, while intravenous administration of NPS 2143 inhibited this effect (Mine and Zhang, 2015; Zhang et al., 2015). Dietary supplementation of Trp, L-Phe, and tyrosine (Tyr) also reduced the expression of intestinal inflammatory markers in piglets, after short-term induction of inflammation by lipopolysaccharides (Liu et al., 2018). However, a recent study found that the CaSR PAMs, cinacalcet and GSK3004774 (an intestine-specific modulator), did not reduce the inflammatory effects of DSS, while NPS 2143 ameliorated the DSS-induced symptoms and reduced immune cell infiltration (Elajnaf et al., 2019).
The anti-inflammatory effect of the CaSR was also shown in vitro, in cell lines. In a colonic myofibroblast cell line, activation of the CaSR inhibited tumor necrosis factor alpha (TNFα) secretion (Kelly et al., 2011) and increased expression of bone morphogenetic protein-2, a promoter of colonic epithelial barrier maturation (Peiris et al., 2007). In colon cancer cell lines, amino acids and dipeptides inhibited pro-inflammatory cytokine secretion and the effect was reversed by NPS 2143 (Mine and Zhang, 2015; Zhang et al., 2015). Inflammatory cytokines, such as TNF-α, IL-1β and IL-6 increased the expression of the CaSR at the mRNA and protein level in some colon cancer cell lines (Fetahu et al., 2014), which could be a defense mechanism against inflammation in the intestines.
Bicarbonate (HCO3 -) secretion in the colon is fine-tuned by the CaSR (Tang et al., 2015). However, it seems that the neurogenic secretory responses in the intestinal epithelium are mediated mainly by the CaSR expressed in the ENS and not in the epithelium (Cheng, 2012; Geibel et al., 2006; Tang et al., 2018). As abnormalities of the ENS affect the severity of intestinal inflammation and contribute to the pathogenesis of IBD (Margolis et al., 2011), the CaSR could be a potential therapeutic target.
Increased dietary intake of calcium reduces the risk of several cancers. The inverse correlation between calcium intake and risk of colorectal cancer has been known for decades, although the mechanisms driving the protective effect of calcium were not clear. There is some evidence that the CaSR is one of the central mediators of the anti-tumorigenic effects of calcium, acting as a tumor suppressor (Kallay et al., 2003; Kallay et al., 1997; Yang et al., 2017). In colon cancer cells activation of the CaSR increased differentiation, reduced proliferation, epithelial-to-mesenchymal transition and expression of stem cell markers (Aggarwal et al., 2015; Aggarwal et al., 2017). The signaling pathways involved in these processes still need to be determined. Interestingly, in the upper intestinal tract, it seems that the CaSR functions as an oncogene, as it promoted gastric cancer cell proliferation (Xie et al., 2017).
G. CaSR in the pancreas
The CaSR is expressed in pancreatic acinar cells (Bruce et al., 1999), which promote digestion via nutrient-stimulated release of digestive enzymes and fluid. The CaSR is also expressed in the pancreatic islets, on glucagon-secreting α cells and insulin-secreting β cells (Babinsky et al., 2017). Thus, the CaSR may influence not only protein metabolism but also carbohydrate and fat metabolism.
Ca2+ o is critical for pancreatic islet function and acts via voltage-gated Ca2+ channels to trigger the exocytosis of insulin- and glucagon-containing secretory granules from β- and α-cells, respectively (Rorsman et al., 2012). Ca2+ o also activates the pancreatic islet CaSR, with ex vivo and in vitro studies demonstrating a role for the CaSR in mediating islet hormone secretion. Thus, stimulation of isolated human islets and an insulin secreting mouse cell line (MIN6) with the PAM, NPS R-568, potentiated Ca2+ o-mediated insulin secretion (Gray et al., 2006), whereas knockdown of the CaSR through RNA interference diminished glucose-induced insulin secretion in MIN6 cells that were cultured as pseudo-islets (Kitsou-Mylona et al., 2008). Studies involving MIN6 pseudo-islets have also revealed CaSR-stimulated insulin secretion to be mediated by PLC and the MAPK pathway (Gray et al., 2006). Furthermore, CaSR-mediated MAPK activation in MIN6 cells induces β-cell proliferation, thus highlighting a potential role for the CaSR in the regulation of β-cell mass (Kitsou-Mylona et al., 2008). The CaSR also upregulates the expression of E-cadherin in MIN6 cells, which is associated with increased adherence between neighboring β-cells (Hills et al., 2012). Thus, the CaSR may facilitate cell-to-cell communication within an individual pancreatic islet to coordinate insulin secretion from β-cells (Hodgkin et al., 2008). In addition, the level of islet CaSR expression correlates with insulin secretion from isolated wild-type mouse pancreatic islets (Oh et al., 2016). Studies in wild-type mice have shown that islet CaSR expression increases with age, which may compensate for the insulin resistance in aged mice by increasing insulin secretion (Oh et al., 2016). Transient stimulation of isolated human islets with Ca2+ o and NPS R-568 also promoted glucagon secretion, thereby indicating a role for the CaSR in α-cells (Gray et al., 2006). The intestinal CaSR, which is activated by dietary amino acids and peptides, may also influence pancreatic islet function by regulating the secretion of incretin hormones. In support of this, studies involving isolated mouse intestine have shown that the CaSR is expressed in glucagon-like peptide 1 (GLP-1)-secreting L-cells and also that oligopeptides enhance GLP-1 secretion by activation of the CaSR (Diakogiannaki et al., 2013).
The role of the CaSR in systemic glucose homeostasis has been investigated in studies involving human subjects. An association study reported a common coding-region CASR gene variant to be an independent determinant of plasma glucose concentrations in renal transplant recipients (Babinsky et al., 2015). However, another study involving patients with FHH1 (caused by germline loss-of-function CASR mutations), did not reveal any alterations in insulin secretion or glucose tolerance (Wolf et al., 2014). The effect of altered CaSR function on glucose tolerance has also been evaluated in an ADH1 mouse model, which is referred to as Nuclear flecks (Nuf) as the mutant mouse was initially identified to have nuclear cataracts (Babinsky et al., 2017). Nuf mice, which harbor a germline gain-of-function CaSR mutation (Leu723ICL2Gln) causing hypocalcemia, have impaired glucose tolerance and hypoinsulinemia in association with reductions in pancreatic islet mass and β-cell proliferation (Babinsky et al., 2017). Nuf mice also lack glucose-mediated suppression of glucagon secretion, which was associated with an increase in α-cell proliferation and an impairment of α-cell membrane depolarization (Babinsky et al., 2017). Administration of the NAM, ronacalceret, ameliorated the hypocalcemia and glucose intolerance of Nuf mice, and these findings highlight the potential utility of targeted CaSR compounds for modulating glucose metabolism (Babinsky et al., 2017).
H. CaSR in mammary glands
The CaSR is expressed in breast epithelial cells, where its main role is to fine-tune maternal Ca2+ metabolism by balancing Ca2+ o mobilization and usage: it ensures the supply of Ca2+ for milk while protecting against maternal hypocalcemia (Cheng et al., 1998; Kim and Wysolmerski, 2016; VanHouten et al., 2004; VanHouten et al., 2007). The expression of the CaSR is increased during lactation, when it regulates Ca2+ o transport into milk. In parallel, it inhibits synthesis of PTHrP by coupling with Gi to inhibit adenylyl cyclase activity and cAMP production (VanHouten et al., 2004). During milk production, the CaSR enables the lactating breast to participate in the regulation of systemic Ca2+ o and bone metabolism. VanHouten and Wysolmerski (Vanhouten and Wysolmerski, 2013) suggested a negative feedback between systemic Ca2+ o delivered to the lactating breast and PTHrP synthesis and secretion by mammary epithelial cells. When the mother’s serum calcium level is adequate, the CaSR in breast epithelial cells stimulates calcium secretion into milk, but reduces Ca2+ o usage when the mother’s calcium supply becomes limited (Vanhouten and Wysolmerski, 2013). In mammary epithelial cells, the CaSR regulates Ca2+ o transport by altering the activity of the plasma membrane Ca2+ ATPase 2, however, the detailed molecular mechanism is not yet known (VanHouten et al., 2007; VanHouten and Wysolmerski, 2007).
The CaSR is expressed also in the neoplastic mammary gland. In contrast to normal breast cells, in breast cancer cells the CaSR stimulates PTHrP secretion. This is possible because the CaSR switches coupling from Gi/o to Gs, leading to stimulation of cAMP and PTHrP synthesis (Mamillapalli et al., 2008). The higher PTHrP levels, secreted due to activation of the CaSR, inhibit the cell cycle inhibitor p27kip1 and the apoptosis inducing factor, stimulating cell proliferation and reducing apoptosis (Kim et al., 2016). Moreover, PTHrP is an activator of osteoclasts and often stimulates osteolytic bone destruction when secreted from cancer cells that metastasize to bone (Wysolmerski, 2012).
The CaSR is highly expressed by metastatic breast cancer cells and potentiates their osteolytic ability, promoting a more aggressive behavior. In vitro, the CaSR promoted breast cancer cell migration only in cells capable of forming bone metastases (e.g. MDA-MB-231, MCF-7), but not in BT474 cells that have no bone-metastatic potential, even though CaSR levels were similar in all cell types (Saidak et al., 2009). It has been shown recently that breast cancer cells overexpressing the wildtype CaSR injected intratibially into Balb/c-Nude mice, led to osteolytic lesions through an epiregulin-mediated mechanism (Boudot et al., 2017). The oncogenic potential of the CaSR in breast cancer cells was also suggested by the fact that activation of the CaSR by NPS R-568 or Ca2+ o increased secretion of pro-angiogenic and chemotactic cytokines and growth factors from the highly invasive MDA-MB-231 breast cancer cells (Hernandez-Bedolla et al., 2015). Another group, however, found that activating the CaSR with Ca2+ o induced sensitivity of MCF-7 and MDA-MB-435 cells to the chemotherapeutic drug paclitaxel and reduced malignant behavior. The paclitaxel-resistant cells expressed no CaSR (Liu et al., 2009). This group suggested a positive link between the tumor suppressive functions of BRCA1 and the CaSR (Promkan et al., 2011).
I. CaSR in airway smooth muscle and epithelium
Asthma is characterized by airway hyperresponsiveness, inflammation and remodeling of the conducting airways. A number of mechanisms, many driven by inflammation, have been hypothesized to contribute to airway hyperresponsiveness and/or remodeling. Among these, local increases of polycations are seen in the airways of asthmatic patients (Kurosawa et al., 1992) and, vice versa, increased inflammation increases the local concentration of polycations. Furthermore, the polycations, ECP and MBP, are markers for asthma severity and stability. Elevated arginase activity increases the consumption of L-arginine to enhance production of the polycations, spermine, spermidine and putrescine (North et al., 2013). Indeed, arginase inhibitors have been proposed to have therapeutic potential for allergic asthma (van den Berg et al., 2018). Recent evidence suggests that the CaSR is expressed in the airway epithelium, smooth muscle and inflammatory cells, and that polycations act at the CaSR and are directly implicated in the pathogenesis of asthma (Yarova et al., 2015). Yarova et al have also shown that inhaled CaSR NAMs, delivered topically, reverse airway hyperresponsiveness, inflammation and remodeling in in vivo models of allergic asthma, and other inflammatory lung diseases such as chronic obstructive pulmonary disease. Inhaled NAMs also show efficacy in non-allergic asthma, which is often associated with poor response to steroids and for which currently there is no treatment (Riccardi unpublished observations). Four CaSR NAMs have been studied in humans, NPSP795, ronacaleret, AXT914 and JTT-305 (Section IIB, Table 1), which could be repurposed, via the inhaled route, as novel asthma treatments. Crucially, delivery of CaSR NAMs directly to the lung does not significantly affect serum calcium levels up to 24 hours post-treatment, suggesting absence of any significant systemic overspill and possible effects on whole body mineral ion homeostasis in vivo. Thus, CaSR NAMs could provide a new therapeutic approach to treating inflammatory lung disease in humans.
J. CaSR in the vasculature
The CaSR is expressed in the intima, media and adventitia of the blood vessels, in endothelial, smooth muscle cells and in the perivascular neurons. While consumption of dietary calcium reduces blood pressure (Nakamura et al., 2019; Rietsema et al., 2019), direct actions of Ca2+ o on isolated blood vessels have yielded contrasting effects, with both relaxation and constriction reported (Bohr, 1963). Further, the molecular mechanisms underlying these actions are elusive. Studies carried out over the last two decades indicate that the CaSR could mediate at least some of the effects of Ca2+ o on vascular function, with opposing effects in the endothelium and smooth muscle cell layers of the blood vessels. Specifically, CaSR activation by the CaSR PAM, cinacalcet, in the vascular endothelium leads to hyperpolarization and attendant nitric oxide release and vasodilatation (Smajilovic et al., 2007). In contrast, studies of Casr gene ablation in the vascular smooth muscle cells show that activation of the CaSR in these cells leads to contraction, as evidenced by the fact that Casr knockout mice exhibit impaired vascular reactivity, hypotension and reduced contractile response to Ca2+ o (Schepelmann et al., 2016). Thus, the CaSR sets blood vessel tone by integrating pro-relaxing (endothelium-mediated) actions with pro-contractile (smooth-muscle mediated) effects (Schepelmann et al., 2016). Therefore, altered CaSR expression within either the endothelium or smooth muscle could account for the abnormal vascular reactivity seen in advanced CKD, or in type 2 diabetes. Indeed, systemic administration of the CaSR PAM, NPS R-568, initially evokes an increase in blood pressure in control and in uremic rats (a model for advanced CKD), which is followed by a reduction in blood pressure, but only in uremic animals (Odenwald et al., 2006), suggesting partial loss of the CaSR-dependent contractile component of the vasculature.
However, there is some controversy regarding the role of the CaSR in regulating blood pressure. In ex vivo studies in rat mesenteric arteries, relaxant responses to cinacalcet and calindol were not blocked by calhex 231 (Thakore and Ho, 2011), which at the time was believed to be a CaSR NAM (but has now been shown to have mixed PAM and NAM activity, see w IIB). Nonetheless, Ca2+ o influx into these vessels stimulated by the α1 adrenergic receptor agonist, methoxamine, was inhibited, not potentiated, by cinacalcet and calindol, as were contractions in response to an L-type calcium channel activator (Thakore and Ho, 2011). Given that arylalkylamine PAMs are structurally related to the nonselective calcium channel blocker, fendiline, and have low affinity for calcium channels, the relaxing effects of CaSR PAMs in arteries may in part arise from off-target calcium channel effects. Further support for a non-CaSR-mediated effect of PAMs in the vasculature comes from findings that while the S-enantiomers of arylalkylamine PAMs have little activity at the CaSR, the effects of NPS R-568 on vascular tone, BP and heart rate are not stereoselective, and only occur at concentrations in excess of those required to inhibit PTH secretion (Nakagawa et al., 2009).
End-stage CKD is associated with impaired mineral ion metabolism, which can lead to pathological vascular calcification of the medial layer of the blood vessel, left ventricular hypertrophy and increased cardiovascular mortality (Locatelli et al., 2002). CaSR expression is significantly reduced in the medial layer of calcifying blood vessels, and is completely absent in areas of extensive medial calcification, suggesting an involvement of the CaSR in the vascular calcification process (Alam et al., 2009). Human and bovine vascular smooth muscle cells exposed to Ca2+ and phosphate concentrations mimicking those seen during pathological CKD exhibit marked calcification in vitro, an effect that is exacerbated by CaSR downregulation and that is reversed by the CaSR PAM, NPS R-568 (Alam et al., 2009). In addition, NPS R-568 reduces blood pressure and ameliorates cardiac remodeling in animal models of advanced CKD in vivo (Ogata et al., 2003). Taken together, these results suggest that loss of CaSR expression by the medial layer of the blood vessels in advanced CKD leads to vascular calcification, and that CaSR PAMs might be vasculo-protective by directly restoring normal CaSR expression levels within the vasculature. However, CaSR PAMs reduce systemic levels of serum Pi and PTH through their actions on the parathyroid CaSR, and parathyroidectomy suppresses vascular calcification (Kawata et al., 2008), therefore PAM-mediated reduction of vascular calcification may be dependent on activation of parathyroid CaSRs. While in vitro and in vivo studies support a direct role for the vascular CaSR in protecting vascular function, human observational studies of clinical evaluation of the CaSR PAM, cinacalcet, assessed by the EValuation Of Cinacalcet Hydrochloride Therapy to Lower CardioVascular Events (EVOLVE) randomized controlled trial, failed to reach its endpoints (reduction of all-cause and cardiovascular mortality in advanced CKD patients) (Chertow et al., 2012). However, a recent Bayesian meta-analysis, combined with a systematic literature review concluded that, once subject ages and high drop-out rates throughout the trial are accounted for, cinacalcet treatment does reduce mortality rates in secondary hyperparathyroidism patients on hemodialysis (Lozano-Ortega et al., 2018). Therefore, further clinical studies are needed to fully evaluate the effects of CaSR PAMs on cardiovascular and all-cause mortality in advanced CKD patients.
Finally, it should be pointed out that the CaSR is also expressed in arterial smooth muscle cells of the pulmonary vasculature, where receptor activation leads to pulmonary vasoconstriction and proliferation. Here, CaSR NAMs prevent the development and progression of pulmonary hypertension in mouse and rat models in vivo (Tang et al., 2016a). Thus, targeting the CaSR in the pulmonary arteries with inhaled NAMs might provide a novel treatment for patients with idiopathic pulmonary hypertension.
K. CaSR in the brain
For a comprehensive review of all evidence for CaSR function in the brain, readers are directed to a recent review (Giudice et al., 2019).
While the role of the CaSR in human brain requires validation, the CaSR is expressed throughout the rat brain, with particular abundance in the hippocampus, striatum, cerebellum, pituitary and olfactory bulb (Ruat et al., 1995). However, CaSR expression can change with developmental age (Vizard et al., 2008), supporting a role for the CaSR in brain development. For instance, rat CaSR expression increases in fetal oligodendrocyte precursor cells and postnatal immature oligodendrocytes during myelination of nerve axons, but expression declines in mature oligodendrocytes (Chattopadhyay et al., 2008; Chattopadhyay et al., 1998; Ferry et al., 2000).
While hyperparathyroidism and consequent early lethality resulting from global Casr ablation precludes determination of the role of the CaSR in brain development, concomitant Casr and PTH ablation prevents hyperparathyroidism, and mice survive to adulthood (Kos et al., 2003). In brains of Casr-/-/Pth-/- mice, neuron and glial cell differentiation markers were reduced after birth, while differentiation of neural stem cells from Casr-/- mice was delayed (Liu et al., 2013). These mice also had reduced numbers of gonadotropin-releasing hormone positive neurons in the hypothalamus. These findings suggest a role for the CaSR in neuron and glial cell differentiation.
To elucidate region-specific CaSR brain functions, hippocampus-specific Casr ablation 3 weeks post birth has been undertaken (Kim et al., 2014). While mice did not display an obvious phenotype under normal conditions, they were protected from hippocampal neuronal damage in response to ischemia-induced injury, which mimics injury sustained during cardiac arrest or stroke. In line with these findings, hypoxia increases CaSR expression in rat hippocampal neurons in vivo and in vitro (Bai et al., 2015), but neuroprotection from ischemia is blocked when the related class C GPCR, GABABR1, was also ablated (Kim et al., 2014). This may be explained by findings of an increase in CaSR expression in hippocampal neurons in culture upon suppression of GABABR1 levels (Chang et al., 2007), which is also observed in cortical neurons of mice who have experienced controlled cortical impact as a model for traumatic brain injury (Kim et al., 2013). In support of a role for the CaSR in the hippocampus, in rat hippocampal neurons from wild type but not Casr-/- mice, CaSR activation opens nonselective cation channels (Ye et al., 1997b). Similarly, transfection of a dominant-negative Arg185Gln mutant CaSR into pyramidal neurons of hippocampal brain slice cultures resulted in significantly shorter and less complex dendritic branching (Vizard et al., 2008). Taken together, these studies suggest that CaSR NAMs could be neuroprotective.
In addition to neuroprotective effects upon brain injury, inhibition of brain CaSRs may afford neuroprotection in Alzheimer’s disease. The first evidence for a possible role of the CaSR in Alzheimer’s disease came from a study that demonstrated activation of nonselective cation channels in cultured hippocampal pyramidal neurons from wild type rats and mice, but not from Casr-/- mice (Ye et al., 1997a). While additional studies have since suggested β-amyloid proteins activate the CaSR (Conley et al., 2009; Dal Pra et al., 2014), these findings warrant further validation. Nonetheless, CaSR expression is increased in the hippocampus of an Alzheimer’s disease mouse model (Gardenal et al., 2017) and there is a positive association between CaSR SNPs and Alzheimer’s disease, although only patients who do not habor the Alzheimer’s risk allele encoding ApoE4 (Conley et al., 2009). Further, in human cortical astrocytes and neurons in culture, neurotoxic β-amyloid25-35 stimulates full-length β-amyloid42 secretion, an effect that is blocked by the CaSR NAM, NPS 2143 (Armato et al., 2013; Chiarini et al., 2017b). NPS 2143 also blocked β-amyloid25-35-mediated GSK-3β activation and subsequent phosphorylation of tau in cultured human astrocytes (Chiarini et al., 2017a).
The CaSR has also been implicated in the etiology of neuroblastoma, brain tumors originating from precursor nerve cells of the sympathetic nervous system. Approximately 98% of neuroblastomas are associated with spontaneous mutations in a variety of genes (Aygun, 2018). Analysis of mRNA from neuroblastoma tumors indicates that while the CaSR is expressed in benign differentiated tumors, epigenetic hypermethylation of the CASR P2 promoter region silences CASR transcription in some aggressive neuroblastomas (Casala et al., 2013; de Torres et al., 2009). Similarly, two non-coding CASR SNPs (rs7652579 and rs1501899) that reduce CaSR expression are present in homo- or heterozygous form in 58% of neuroblastoma tumors, but in only 47% of the general population (Masvidal et al., 2017). In a subset of ganglioneuromas, CASR expression was absent in 4/6 tumors haboring rs7652579 and rs1501899. However, neuroblastoma patients with rs7652579 and rs1501899 SNPs did not have poorer outcomes or survival (Masvidal et al., 2017). In contrast, neuroblastomas with a haplotype SNP in the CASR gene coding region were associated with poorer outcomes, including increased risk of death (Masvidal et al., 2013). Importantly, cinacalcet reduced neuroblastoma tumor growth in immunocompromised mice carrying neuroblastoma xenografts, by inducing ER stress, tumor differentiation and fibrosis, as well as upregulation of the tumor proteins, cancer-testis antigens (Rodriguez-Hernandez et al., 2016).
Finally, approximately 40% of patients haboring ADH1 gain-of-function CASR mutations present with seizures (Gorvin, 2019). While this could be associated with the consequent reduction in serum calcium concentrations, a gain-of-expression CaSR mutation, Arg898Gln, was identified in an idiopathic epilepsy patients who did not have low serum concentrations of calcium or PTH (Kapoor et al., 2008). These findings suggest a possible role for CaSR in neurotransmission, which is supported by numerous in vitro studies suggesting the CaSR regulates synaptic transmission and neuronal activity via activation of nonselective cation channels on presynaptic terminals (reviewed in Jones and Smith, 2016).
VI. CaSR-related genetic diseases and therapeutic interventions
A. Loss- and gain-of-function mutations in the CaSR and its signaling partners
Alterations in CaSR signaling, which lead to derangements of mineral homeostasis, can result from: loss-of-function germline mutations of the CASR gene on chromosome 3q21.1, which cause FHH1 and NSHPT; or gain-of-function germline CASR mutations, which lead to ADH1 and Bartter syndrome type V (Table 2, Figure 4) (Hannan et al., 2012; Hannan and Thakker, 2013). In addition, loss- and gain-of-function germline mutations of the GNA11 gene on chromosome 19p13.3, which encodes Gα 11, are associated with FHH2 and ADH2, respectively (Table 2) (Hannan et al., 2016; Nesbit et al., 2013a). Further, loss-of-function germline mutations of the AP2S1 gene on chromosome 19q13.3, which encodes AP2σ, cause FHH3 (Table 2) (Hannan et al., 2015a; Nesbit et al., 2013b).
Table 2. Calcitropic disorders caused by germline CASR, GNA11 and AP2S1 mutations.
Gene mutation and disease | Genotype | Serum calcium | Serum PTH | Urine calcium |
---|---|---|---|---|
CASR mutations | ||||
Loss-of-function | ||||
Familial hypocalciuric hypercalcemia type 1 (FHH1) | Heterozygous* | High | Normal or high | Low |
Neonatal severe hyperparathyroidism (NSHPT) | Heterozygous, compound heterozygous, or homozygous | High | High | Normal, low or high |
Primary hyperparathyroidism (PHPT)† | Heterozygous or homozygous | High | High | Normal, low or high |
Gain-of-function | ||||
Autosomal dominant hypocalcemia type 1 (ADH1) | Heterozygous* | Low | Normal or low | Normal, low or high |
Bartter syndrome type V | Heterozygous | Low | Low | High |
GNA11 mutations | ||||
Loss-of-function | ||||
Familial hypocalciuric hypercalcemia type 2 (FHH2) | Heterozygous | High | Normal or high | Normal or low |
Gain-of-function | ||||
Autosomal dominant hypocalcemia type 2 (ADH2) | Heterozygous | Low | Normal or low | Normal or low |
AP2S1 mutations | ||||
Loss-of-function | ||||
Familial hypocalciuric hypercalcemia type 3 (FHH3) | Heterozygous | High | Normal or high | Low |
Figure 4. CaSR snakeplot with residues linked to loss- and gain-of-function germline mutations.
Snake-plot of the CaSR showing the location of the ECD, 7TM, ICLs, ECLs and carboxy-terminus. Sites of loss- and gain-of-function germline mutations causing FHH1/NSHPT (red), ADH1/Bartter syndrome type V (green), or both FHH1/NSHPT and ADH1/Bartter syndrome type V (yellow), respectively. Snake-plot generated by GPCRdb (Munk et al., 2016) with data from the Human Gene Mutation Database (Stenson et al., 2012).
B. FHH and NSHPT
FHH is a genetically heterogeneous autosomal dominant disorder characterized by lifelong non-progressive elevations of serum calcium concentrations, mild hypermagnesemia, normal or mildly raised serum PTH concentrations and low urinary calcium excretion (Table 2) (Hannan and Thakker, 2013). FHH1 (OMIM #145980) accounts for ~65% of all FHH cases and is usually an asymptomatic disorder. It has been associated with >150 different CASR mutations (Hannan et al., 2018a). The majority (>85%) of these loss-of-function CASR mutations are heterozygous missense substitutions, which are predominantly located in the VFT of the CaSR ECD, and also in the 7TM (Hannan et al., 2012). These FHH1-associated missense mutations cause a loss-of-function by diminishing the signaling responses of CaSR-expressing cells (Leach et al., 2012), or by reducing CaSR anterograde trafficking and cell-surface expression (Huang and Breitwieser, 2007; White et al., 2009). In addition, FHH1-causing missense mutations may induce biased agonism by switching from a wild type CaSR that preferentially increases Ca2+ i mobilization to mutant receptors that demonstrate equal preference for Ca2+ i and MAPK pathways, or that preferentially act via MAPK (Leach and Gregory, 2017; Leach et al., 2013; Leach et al., 2012). Between 10 to 15% of FHH1 cases are caused by heterozygous deletion, insertion, nonsense and splice-site mutations that lead to nonsense-mediated decay of mRNA and CaSR haploinsufficiency, or truncate the CaSR protein (Hannan et al., 2012). The offspring of two parents with FHH1 can harbor biallelic loss-of-function CASR mutations that cause NSHPT (OMIM #239200), and which is associated with marked hyperparathyroidism that leads to hypercalcemia and bone demineralization causing fractures and respiratory distress (Hannan and Thakker, 2013). Occasionally, biallelic loss-of-function CASR mutations can lead to primary hyperparathyroidism, which presents in adulthood (Table 2) (Hannan et al., 2010). Further, some heterozygous mutations (e.g Arg185Gln) can cause NSHPT due to dominant negative effects on the wild-type CaSR (Bai et al., 1997).
FHH2 (OMIM#145981) is the least common form of FHH and has been reported in four probands to-date (Gorvin et al., 2016; Gorvin et al., 2018b; Nesbit et al., 2013a). FHH2 appears to have a mild clinical presentation with serum adjusted total calcium concentrations usually between 2.55 - 2.80 mM (normal range 2.10 - 2.55 mM). Urinary calcium excretion may be normal or low (Table 2) (Gorvin et al., 2016; Gorvin et al., 2018b; Nesbit et al., 2013a). The GNA11 mutations reported in FHH2 probands consist of three missense substitutions (Thr54Met, Leu135Gln, Phe220Ser) and an in-frame isoleucine deletion (Ile200del) (Gorvin et al., 2016; Gorvin et al., 2018b; Nesbit et al., 2013a). All of these mutations impair CaSR signaling responses, and are located within key domains of the Gα 11 protein (Gorvin et al., 2016; Gorvin et al., 2018b; Nesbit et al., 2013a). Thus, the Ile200del and Phe220Ser mutations are located within the Gα 11 GTPase domain and are predicted to diminish the interaction of Gα 11 with the CaSR or PLC, respectively (Gorvin et al., 2018b; Nesbit et al., 2013a). In contrast, the Leu135Gln mutation is situated within the PLC-interacting portion of the Gα 11 helical domain, and the Thr54Met mutation is located at the interface between the helical and GTPase domains, and may potentially affect GTP binding (Gorvin et al., 2016; Nesbit et al., 2013a).
FHH3 (OMIM#600740) has been reported in >60 FHH probands and has a more marked clinical phenotype than FHH1. Thus, FHH3 is associated with significant elevations of serum calcium and magnesium, and also a significantly reduced urinary calcium excretion compared to FHH1 (Table 2) (Hannan et al., 2015a; Vargas-Poussou et al., 2016). In addition, hypercalcemic symptoms, low bone mineral density and alterations in cognitive function have been described in some FHH3 patients (Hannan et al., 2015a; McMurtry et al., 1992). Nearly all FHH3 cases are caused by a missense mutation of the AP2σ Arg15 residue (Arg15Cys, Arg15His or Arg15Leu) (Fujisawa et al., 2013; Hannan et al., 2015a; Hendy et al., 2014; Hovden et al., 2017; Howles et al., 2016; Nesbit et al., 2013b; Vargas-Poussou et al., 2016). In addition, a genotype-phenotype correlation has been observed at the AP2σ Arg15 residue with the Arg15Leu mutation being associated with significant increases in serum calcium and an earlier age of presentation compared to patients harboring the Arg15Cys or Arg15His AP2σ mutations (Hannan et al., 2015a; Hovden et al., 2017). The AP2σ subunit forms part of the heterotetrameric AP2 complex, which is involved in clathrin-mediated endocytosis (Kelly et al., 2008) and the FHH3-causing AP2σ Arg15 mutations have been shown to reduce CaSR endocytosis and impair endosomal signaling from the internalized CaSR (Gorvin et al., 2018c). However, given the role of AP2 in clathrin-mediated endocytosis, it remains to be established whether phenotypic observations such as cognitive deficits in FHH3 are attributable to CaSR dysregulation, or potentially due to alterations in the endocytosis of other plasma membrane proteins.
C. ADH and Bartter’s syndrome V
ADH is comprised of two genetically distinct variants, designated ADH1 and 2, which are caused by germline gain-of-function mutations of the CaSR and Gα 11, respectively (Table 2) (Hannan et al., 2016). ADH1 (OMIM#601198) is characterized by mild-to-moderate hypocalcemia in association with mild hypomagnesemia, hyperphosphatemia, and serum PTH concentrations that are usually detectable but within the lower half of the reference range (Nesbit et al., 2013a). Patients with ADH1 have significantly increased urinary calcium excretion compared to hypoparathyroid patients (Yamamoto et al., 2000), and ~10% of ADH1 patients have an absolute hypercalciuria (Nesbit et al., 2013a). Some patients with ADH1 may have ectopic calcifications and/or elevations in bone mineral density (BMD) (Pearce et al., 1996), and patients with a severe gain-of-function CaSR mutation may also develop a Bartter syndrome (referred to as Bartter syndrome type V) (Table 2), which is characterized by renal salt wasting leading to volume depletion, hyper-reninemic hyperaldosteronism and hypokalemic alkalosis (Watanabe et al., 2002). Over 90 different ADH1-causing CaSR mutations have been reported (Hannan et al., 2016; Hannan and Thakker, 2013) and around 95% of these are heterozygous missense substitutions, whereas ~5% are in frame or frameshift insertion or deletion mutations (Hannan et al., 2012). ADH1 mutations cluster within the second loop of the VFT domain (residues 116-136), which contributes to the dimeric CaSR interface (Geng et al., 2016) (Figure 4). A second ADH1 mutational hotspot is located in a region that encompasses transmembrane helices 6 and 7, and the intervening third ECL of the CaSR (residues 819-837) (Hannan et al., 2016). This transmembrane region may participate in a network of interactions with other transmembrane helices (Dore et al., 2014), thereby causing the CaSR to adopt an inactive conformational state.
ADH2 (OMIM#615361) (Table 2) has been reported in seven probands (Li et al., 2014a; Mannstadt et al., 2013; Nesbit et al., 2013a; Piret et al., 2016; Tenhola et al., 2016). ADH2 patients generally have mild-to-moderate hypocalcemia, in keeping with the serum biochemical phenotype of ADH1 (Hannan et al., 2016). However, ADH2 is associated with a milder urinary phenotype, with significantly reduced urinary calcium excretion compared to ADH1 (Li et al., 2014a). Moreover, short stature caused by postnatal growth insufficiency has been reported in two ADH2 kindreds (Li et al., 2014a; Tenhola et al., 2016). ADH2-causing mutations all comprise missense substitutions (Arg60Cys, Arg60Leu, Arg181Gln, Ser211Trp, Val340Met and Phe341Leu), which enhance CaSR-mediated signaling responses, consistent with a gain-of-function (Li et al., 2014a; Nesbit et al., 2013a; Piret et al., 2016). ADH2-causing mutations cluster at the interface between the Gα 11 helical and GTPase domains (Piret et al., 2016), and may enhance the exchange of GDP and GTP, thereby leading to G protein activation. ADH2 mutations also affect the C-terminal portion of the Gα 11 protein, which facilitates G protein-GPCR coupling (Piret et al., 2016).
D. Animal models of genetic diseases
Mouse models for FHH, NSHPT and ADH have been generated using gene knockout and knock-in techniques, and also by using chemical mutagenesis (Piret and Thakker, 2011).
FHH/NSHPT mouse models. A mouse model lacking the CaSR was generated by replacing part of exon 5 with a neomycin resistance gene (Ho et al., 1995). Mice harboring this germline heterozygous CaSR deletion (Casr+/-) had mild hypercalcemia and hypocalciuria, similar to FHH1 patients, whereas, mice with a homozygous CaSR deletion (Casr -/-) had a phenotype resembling NSHPT, with parathyroid hyperplasia, severe hypercalcemia, bone abnormalities and retarded growth (Ho et al., 1995). The Casr -/- mice died within the first 30 days of life (Ho et al., 1995), which was attributed to severe hyperparathyroidism. In support of this, correction of the hyperparathyroidism by the additional germline ablation of the Pth or Gcm2 genes, rescued the early lethality and bone demineralization in Casr -/- mice (Kos et al., 2003; Tu et al., 2003). The importance of the parathyroid CaSR in the pathogenesis of NSHPT has been further highlighted by mice harboring a parathyroid-specific ablation of the CaSR, which developed severe hypercalcemia and hyperparathyroidism (Chang et al., 2008; Fan et al., 2018). In contrast, mice with a kidney-specific ablation of the CaSR do not have alterations in serum calcium or PTH, but are hypocalciuric, and these findings support an independent role of the kidney CaSR in the regulation of urinary calcium excretion (Toka et al., 2012).
A mouse model for FHH2 has been generated by chemical mutagenesis using the N-ethyl-N-nitrosourea (ENU) alkylating agent (Howles et al., 2017). The mutant mice harbor a germline loss-of-function Gna11 mutation, Asp195Gly (D195G) (Howles et al., 2017). Heterozygous (Gna11+/195G) mice have mild hypercalcemia and normal plasma PTH concentrations (Howles et al., 2017). Homozygous (Gna11195G/195G) mice have significantly increased plasma calcium and PTH concentrations compared to Gna11+/195G and wild type mice (Howles et al., 2017). However, Gna11195G/195G mice do not have growth retardation, bone demineralization or early lethality to suggest an NSHPT phenotype (Howles et al., 2017). Thus, these studies indicate that the loss-of-function D195G Gα 11 mutation is associated with a mild calcitropic phenotype. Furthermore, the Gna11+/195G and Gna11195G/195G mice have no alterations in urinary calcium excretion (Howles et al., 2017), which suggests that Gα 11 may not play a major role in the renal handling of calcium.
ADH mouse models. Three different ADH1 mouse models have been reported (Dong et al., 2015; Hough et al., 2004b). Nuf mice (described in Section VG) was generated by chemical mutagenesis using the isopropyl methane sulfonate (iPMS) alkylating agent (Hough et al., 2004a). Heterozygous and homozygous Nuf mice have hypocalcemia, hyperphosphatemia, reduced plasma PTH concentrations, and ectopic calcifications caused by a germline gain-of-function CaSR mutation, Leu723Gln (Hough et al., 2004a). Two knock-in mouse models, which harbor ADH1-causing germline Cys129Ser or Ala843Glu gain-of-function CaSR mutations have also been generated (Dong et al., 2015). Homozygous mutant knock-in mice exhibited embryonic or perinatal lethality, whereas heterozygous knock-in mice have hypocalcemia, hyperphosphatemia, reduced plasma PTH, hypercalciuria and renal calcifications, consistent with the phenotype of ADH1 patients (Dong et al., 2015).
Two mouse models for ADH2 have been described (Gorvin et al., 2017; Roszko et al., 2017). One mouse model, which is known as Dark skin 7 (Dsk7), was generated by ENU chemical mutagenesis (Gorvin et al., 2017), and harbors a germline gain-of-function Gα 11 mutation, Ile62Val, whilst the other ADH2 mouse model was generated by CRISPR-Cas9 gene editing and harbors a human ADH2-causing germline Gα 11 mutation, Arg60Cys (Roszko et al., 2017). Both of these ADH2 mouse models have hypocalcemia, hyperphosphatemia, reduced plasma PTH and normocalciuria in association with increased skin pigmentation (Gorvin et al., 2017; Roszko et al., 2017).
E. Therapeutic interventions – successes and failures
CaSR PAMs represent a targeted therapy for symptomatic forms of FHH (Hannan et al., 2018b), and potentiate the signaling responses of cells expressing FHH-associated CaSR, Gα 11 or AP2σ mutant proteins in vitro (Table 3) (Babinsky et al., 2016; Gorvin et al., 2018b; Howles et al., 2016; Leach et al., 2013; Rus et al., 2008). Furthermore, cinacalcet treatment is effective at decreasing serum calcium concentrations in FHH1 patients and has been reported to improve hypercalcemic symptoms occasionally associated with FHH1 such as anorexia, polydipsia and constipation (Table 3) (Alon and VandeVoorde, 2010; Rasmussen et al., 2011; Sethi et al., 2017). However, the response of NSHPT to cinacalcet is variable and appears to depend on the underlying CASR mutation. Indeed, cinacalcet rectifies the hypercalcemia and hyperparathyroidism in NSHPT patients harboring a heterozygous Arg185Gln CaSR mutation (Fisher et al., 2015; Gannon et al., 2014; Reh et al., 2011), but is less effective for NSHPT patients with biallelic truncating CASR mutations (Table 3) (Atay et al., 2014; Garcia Soblechero et al., 2013), which would be a consequence of the truncated mutant receptor being unable to bind cinacalcet or couple with intracellular signaling proteins (Hannan et al., 2018b). Cinacalcet has also rectified the hypercalcemia in a mouse model for FHH2 (Howles et al., 2017), and ameliorated the hypercalcemia in a symptomatic FHH2 patient (Table 3) (Gorvin et al., 2018b). Furthermore, cinacalcet is an effective therapy for symptomatic hypercalcemia caused by all three types of FHH3-causing Arg15 AP2σ mutations (Table 3) (Howles et al., 2016). However, hypocalcemic symptoms have occurred in a cinacalcet-treated child affected by FHH3 and the chromosome 22q11.2 deletion syndrome (Tenhola et al., 2015). Thus, long-term surveillance is required to detect hypocalcemia in cinacalcet-treated FHH patients (Howles et al., 2016).
Table 3. Summary of key studies assessing effectiveness of calcimimetics and calcilytics for FHH, NSHPT and ADH.
Disorder | In vitro studies | In vivo studies |
---|---|---|
Hypercalcemic disorders | ||
FHH1/NSHPT | NPS R-568 and cinacalcet enhance the signaling responses and cell-surface expression of loss-of-function FHH1/NSHPT-causing CaSR mutants (Leach et al., 2013; Rus et al., 2008) | Cinacalcet lowers serum calcium and PTH concentrations, and improves hypercalcemic symptoms in FHH1 patients (Alon and VandeVoorde, 2010; Rasmussen et al., 2011; Sethi et al., 2017) Cinacalcet lowers serum calcium and PTH concentrations in NSHPT patients harboring a heterozygous Arg185Gln CASR mutation (Fisher et al., 2015; Gannon et al., 2014; Reh et al., 2011), but is less effective for NSHPT caused by biallelic truncating CASR mutations (Atay et al., 2014; Garcia Soblechero et al., 2013) |
FHH2 | Cinacalcet enhances the signaling responses of cells expressing loss-of-function FHH2-causing Ga11 mutants (Babinsky et al., 2016) | Cinacalcet lowers serum calcium and PTH concentrations in a mouse model for FHH2 (Howles et al., 2017), and also normalizes serum calcium concentrations in an FHH2 patient (Gorvin et al., 2018b) |
FHH3 | Cinacalcet enhances the signaling responses of cells expressing loss-of-function FHH3-causing Arg15Cys, Arg15His, or Arg15Leu AP2σ mutants (Howles et al., 2016) | Cinacalcet lowers serum calcium and PTH concentrations, and improves hypercalcemic symptoms in FHH3 patients with Arg15Cys, Arg15His or Arg15Leu AP2S1 mutations (Howles et al., 2016) |
Hypocalcemic disorders | ||
ADH1 | NPS 2143 reduces the signaling responses of cells expressing gain-of-function ADH1-causing CaSR mutants, but has limited efficacy for constitutively active CaSR mutants (Leach et al., 2013; Letz et al., 2010) | Acute administration of NPS 2143 and JTT-305/MK-5442 increases serum calcium and PTH concentrations in mouse models for ADH1 (Dong et al., 2015; Hannan et al., 2015b) |
ATF936 and AXT914 rectify the gain-of-function caused by constitutively active CaSR mutants (Letz et al., 2014) | Administration of JTT-305/MK-5442 over 12 weeks reduces urinary calcium excretion and prevents nephrocalcinosis in mouse models for ADH1 (Dong et al., 2015) | |
Intravenous infusion of NPSP795 increases serum PTH concentrations and reduces urinary calcium excretion in ADH1 patients (Roberts et al., 2019) | ||
ADH2 | NPS 2143 reduces the signaling responses of cells expressing gain-of-function ADH2-causing Gα11 mutants (Babinsky et al., 2016; Gorvin et al., 2017; Roszko et al., 2017) | NPS 2143 increases serum calcium and PTH concentrations in mouse models for ADH2 (Gorvin et al., 2017; Roszko et al., 2017) |
CaSR NAMs have been evaluated as a potential targeted therapy for ADH. In vitro studies have demonstrated that NPS 2143 normalizes the signaling responses associated with gain-of-function CASR and GNA11 mutations, which cause ADH1 and ADH2, respectively (Table 3) (Babinsky et al., 2016; Hannan et al., 2015b; Leach et al., 2013; Letz et al., 2010). However, NPS 2143 is less effective for gain-of-function mutations causing Bartter syndrome type V (Leach et al., 2013; Letz et al., 2010). In contrast, the quinazolinone-derived NAMs rectify gain-of-function CASR mutations that cause Bartter syndrome V (Table 3) (Letz et al., 2014). CaSR NAMs have also been characterized in vivo, and single-dose studies have demonstrated that NPS 2143 significantly increases circulating concentrations of calcium and PTH in ADH1 and ADH2 mouse models (Table 3) (Gorvin et al., 2017; Hannan et al., 2015b; Roszko et al., 2017). In addition, repetitive dosing studies have shown that the NAM, JTT-305/MK-5442, prevents the occurrence of nephrocalcinosis in mouse models of ADH1 (Dong et al., 2015). Furthermore, the NAM, NPSP795, has been administered to five ADH1 patients in a phase IIa clinical trial, and increased plasma PTH concentrations and reduced urinary calcium excretion (Table 3) (Roberts et al., 2019). However, circulating calcium concentrations were not altered in these patients, and the optimal dosing regimen of NPSP795 for ADH remains to be established.
VII. Conclusions and perspective
The CaSR is a highly complex GPCR, evidenced by its widespread tissue expression and varied physiological roles, its capacity to respond to multiple stimuli that act via numerous binding sites, and the ability of different stimuli to bias CaSR signaling towards distinct subsets of G protein-dependent and independent signaling pathways. The existence of multiple allosterically-linked binding sites for endogenous CaSR ligands demonstrates how allostery is fundamental to CaSR activity. It is therefore unsurprising that the CaSR was the first GPCR for which an allosteric therapeutic, cinacalcet, was FDA approved. The clinical success of cinacalcet in treating various forms of hyperparathyroidism highlights the potential of targeting the CaSR with allosteric drugs. Given the many fundamental roles of the CaSR, the CaSR is a putative therapeutic target for numerous diseases beyond Ca2+ o homeostasis, including asthma, diabetes and cancer. Thus, drug discovery efforts at the CaSR will no doubt continue.
In addition to the aforementioned CaSR (patho)physiology, ongoing research is expanding the known roles of this receptor. Analysis of CASR SNPs supports associations between CaSR expression or activity and the risk of kidney stones (Vezzoli et al., 2011), vascular calcification (Babinsky et al., 2015), breast cancer (Li et al., 2014b; Wang et al., 2017), psoriasis, (Zuo et al., 2015) and serum glucose concentrations (Babinsky et al., 2015). Further, the sensitivity of the CaSR to amino acids and other stimuli raises the possibility that Ca2+ o is not solely responsible for CaSR-mediated Ca2+ o homeostasis. Indeed, high dietary protein intake modestly increases bone density at some sites and reduces hospital stay after fracture (Dawson-Hughes, 2003; Shams-White et al., 2017). In contrast, low protein diets induce secondary hyperparathyroidism (Dubois-Ferrière et al., 2011; Kerstetter et al., 2000), and acute increases in L-amino acids suppress PTH secretion and potentiate Ca2+ o-mediated Ca2+ i mobilization in human parathyroid cells (Conigrave et al., 2004). Thus, the CaSR could couple protein metabolism to changes in Ca2+ o homeostasis (Conigrave et al., 2008; Conigrave et al., 2002).
While novel analytical methods such the operational model of allosterism have facilitated quantification of CaSR drug actions, CaSR drug discovery still suffers from limited tools to directly probe drug binding (e.g. commercially available radioligands or fluorescently-labeled ligands), and from the lack of 7TM and full-length CaSR structures for structure-based drug discovery. Further, given the critical importance of Ca2+ o homeostasis to human health, novel drugs that target the CaSR outside the parathyroid glands and kidney must have limited on-target effects in these tissues, e.g. by delivery of the drug regio-specifically to the targeted tissue. Alternatively, biased signaling has the potential to revolutionize our ability to target GPCRs in a tissue-specific manner, by directing receptor signaling towards desirable pathways that mediate therapeutic effects at the expense of pathways linked to unwanted effects. Further, major advances in GPCR structural biology resulting in 7TM and full-length structures of the class C GPCRs, mGlu1 and mGlu5 (Dore et al., 2014; Koehl et al., 2019; Wu et al., 2014), provide confidence for forthcoming CaSR structural biology efforts. Thus, the future holds much promise for the design of novel drugs that target the CaSR.
Significance statement.
The CaSR is a complex GPCR that possesses multiple orthosteric and allosteric binding sites, is subject to biased signaling via several different G proteins, and has numerous (patho)physiological roles. Understanding the complexities of CaSR structure, function and biology will aid future drug discovery efforts seeking to target this receptor for a diversity of diseases. This review summarizes what is known to date regarding key structural, pharmacological and physiological features of the CaSR.
Acknowledgements
This article was written as a result of updating the IUPHAR-BPS database by the CaSR Nomenclature Subcommittee for the International Union of Pharmacology and colleagues. The authors acknowledge funding from the Lundbeck Foundation (TCM and HB-O), the Marie Sklodowska-Curie Actions of the European Union’s Horizon 2020 research and innovation programme under REA grant agreement NO 675228 (HB-O, FMH, DTW, EK, RVT and DR) and NO 797497 (TCM and HB-O), the Australian Research Council (KL), the Australian National Health and Medical Research Council (KL, ADC, RM), the Cass Foundation (ANK), Wellcome Trust Investigator Award (106995/Z/15/Z, to RVT); National Institute for Health Research Senior Investigator Award (NF-SI-0514-10091, to RVT).
Nonstandard abbreviations in alphabetical order
-
1,25(OH)2D
1,25-dihydroxyvitamin D
ADH
autosomal dominant hypocalcemia
ADIS
agonist-driven insertional signaling
AP2
adaptor-related protein complex-2
BTU
benzothiazole trisubstituted urea
Ca2+i
intracellular calcium
Ca2+o
extracellular calcium
CaSR
calcium-sensing receptor
CKD
chronic kidney disease
cryo-EM
cryogenic electron microscopy
cTAL
cortical thick ascending limb
ERK1/2
extracellular signal-regulated kinase 1/2
FHH
familial hypocalciuric hypercalcemia
GLP-1
glucagon-like peptide 1
GPCR
G protein-coupled receptor
GRK
GPCR kinase
IP
inositol phosphate
JNK
c-Jun N-terminal kinase
MAPK
mitogen-activated protein kinase
NAM
negative allosteric modulator
NSHPT
neonatal severe hyperparathyroidism
PAM
positive allosteric modulator
PKC
protein kinase C
PLC
phospholipase C
PTH
parathyroid hormone
TAL
thick ascending limb
TNCA
L-1,2,3,4-tetrahydronorharman-3-carboxylic acid.
Footnotes
Authorship contributions
All authors contributed to the writing of this manuscript. KL and HBO edited and coordinated the manuscript writing. KL, TMJ and HBO generated the manuscript figures.
Contributor Information
Leach Katie, Drug Discovery Biology, Monash Institute of Pharmaceutical Science, Monash University, Parkville, Australia.
M Hannan Fadil, Nuffield Department of Women’s & Reproductive Health, University of Oxford, UK; Academic Endocrine Unit, Radcliffe Department of Clinical Medicine, University of Oxford, UK.
C Møller Thor, Department of Drug Design and Pharmacology, Faculty of Health and Medical Sciences, University of Copenhagen, Denmark.
T Ward Donald, Faculty of Biology, Medicine and Health, The University of Manchester, UK.
Kallay Enikö, Department of Pathophysiology and Allergy Research, Medical University of Vienna, Austria.
S Mason Rebecca, Physiology and Bosch Institute, University of Sydney, Australia.
V Thakker Rajesh, Academic Endocrine Unit, Radcliffe Department of Clinical Medicine, University of Oxford, UK.
Riccardi Daniela, School of Biosciences, Cardiff University, UK.
D Conigrave Arthur, School of Life and Environmental Sciences, University of Sydney, Australia.
Bräuner-Osborne Hans, Department of Drug Design and Pharmacology, Faculty of Health and Medical Sciences, University of Copenhagen, Denmark.
References
- Aggarwal A, Prinz-Wohlgenannt M, Groschel C, Tennakoon S, Meshcheryakova A, Chang W, Brown EM, Mechtcheriakova D, Kallay E. The calcium-sensing receptor suppresses epithelial-to-mesenchymal transition and stem cell-like phenotype in the colon. Mol Cancer. 2015;14:61. doi: 10.1186/s12943-015-0330-4. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Aggarwal A, Schulz H, Manhardt T, Bilban M, Thakker RV, Kallay E. Expression profiling of colorectal cancer cells reveals inhibition of DNA replication licensing by extracellular calcium. Biochim Biophys Acta. 2017;1864:987–996. doi: 10.1016/j.bbamcr.2017.01.017. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Alexander SPH, Christopoulos A, Davenport AP, Kelly E, Marrion NV, Peters JA, Faccenda E, Harding SD, Pawson AJ, Sharman JL, Southan C, et al. The Concise Guide to PHARMACOLOGY 2017/18: G protein-coupled receptors. Br J Pharmacol. 2017;174 Suppl 1:S17–S129. doi: 10.1111/bph.13878. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Al-Dujaili SA, Koh AJ, Dang M, Mi X, Chang W, Ma PX, McCauley LK. Calcium sensing receptor function supports osteoblast survival and acts as a co-factor in PTH anabolic actions in bone. J Cell Biochem. 2016;117:1556–1567. doi: 10.1002/jcb.25447. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Alam MU, Kirton JP, Wilkinson FL, Towers E, Sinha S, Rouhi M, Vizard TN, Sage AP, Martin D, Ward DT, Alexander MY, et al. Calcification is associated with loss of functional calcium-sensing receptor in vascular smooth muscle cells. Cardiovasc Res. 2009;81:260–268. doi: 10.1093/cvr/cvn279. [DOI] [PubMed] [Google Scholar]
- Alamshah A, Spreckley E, Norton M, Kinsey-Jones JS, Amin A, Ramgulam A, Cao Y, Johnson R, Saleh K, Akalestou E, Malik Z, et al. L-phenylalanine modulates gut hormone release and glucose tolerance, and suppresses food intake through the calcium-sensing receptor in rodents. Int J Obes (Lond) 2017;41:1693–1701. doi: 10.1038/ijo.2017.164. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Alexander ST, Hunter T, Walter S, Dong J, Maclean D, Baruch A, Subramanian R, Tomlinson JE. Critical cysteine residues in both the calcium-sensing receptor and the allosteric activator AMG 416 underlie the mechanism of action. Mol Pharmacol. 2015;88:853–865. doi: 10.1124/mol.115.098392. [DOI] [PubMed] [Google Scholar]
- Alon US, VandeVoorde RG. Beneficial effect of cinacalcet in a child with familial hypocalciuric hypercalcemia. Pediatr Nephrol. 2010;25:1747–1750. doi: 10.1007/s00467-010-1547-5. [DOI] [PubMed] [Google Scholar]
- Amino Y, Nakazawa M, Kaneko M, Miyaki T, Miyamura N, Maruyama Y, Eto Y. Structure-CaSR-activity relation of kokumi γ-glutamyl peptides. Chem Pharm Bull (Tokyo) 2016;64:1181–1189. doi: 10.1248/cpb.c16-00293. [DOI] [PubMed] [Google Scholar]
- Arey BJ, Seethala R, Ma Z, Fura A, Morin J, Swartz J, Vyas V, Yang W, Dickson JK, Jr, Feyen JH. A novel calcium-sensing receptor antagonist transiently stimulates parathyroid hormone secretion in vivo. Endocrinology. 2005;146:2015–2022. doi: 10.1210/en.2004-1318. [DOI] [PubMed] [Google Scholar]
- Armato U, Chiarini A, Chakravarthy B, Chioffi F, Pacchiana R, Colarusso E, Whitfield JF, Dal Pra I. Calcium-sensing receptor antagonist (calcilytic) NPS 2143 specifically blocks the increased secretion of endogenous Aβ42 prompted by exogenous fibrillary or soluble Aβ25-35 in human cortical astrocytes and neurons-therapeutic relevance to Alzheimer’s disease. Biochim Biophys Acta. 2013;1832:1634–1652. doi: 10.1016/j.bbadis.2013.04.020. [DOI] [PubMed] [Google Scholar]
- Atay Z, Bereket A, Haliloglu B, Abali S, Ozdogan T, Altuncu E, Canaff L, Vilaca T, Wong BY, Cole DE, Hendy GN, et al. Novel homozygous inactivating mutation of the calcium-sensing receptor gene (CASR) in neonatal severe hyperparathyroidism-lack of effect of cinacalcet. Bone. 2014;64:102–107. doi: 10.1016/j.bone.2014.04.010. [DOI] [PubMed] [Google Scholar]
- Avlani V, Ma W, Mun H-C, Leach K, Delbridge L, Christopoulos A, Conigrave A. Calcium-sensing receptor-dependent activation of CREB phosphorylation in HEK-293 cells and human parathyroid cells. Am J Physiol Endocrinol Metab. 2013;304:E1097–E1104. doi: 10.1152/ajpendo.00054.2013. [DOI] [PubMed] [Google Scholar]
- Aygun N. Biological and genetic features of neuroblastoma and their clinical importance. Curr Pediatr Rev. 2018;14:73–90. doi: 10.2174/1573396314666180129101627. [DOI] [PubMed] [Google Scholar]
- Babinsky VN, Hannan FM, Gorvin CM, Howles SA, Nesbit MA, Rust N, Hanyaloglu AC, Hu J, Spiegel AM, Thakker RV. Allosteric modulation of the calcium-sensing receptor rectifies signaling abnormalities associated with G-protein α-11 mutations causing hypercalcemic and hypocalcemic disorders. J Biol Chem. 2016;291:10876–10885. doi: 10.1074/jbc.M115.696401. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Babinsky VN, Hannan FM, Ramracheya RD, Zhang Q, Nesbit MA, Hugill A, Bentley L, Hough TA, Joynson E, Stewart M, Aggarwal A, et al. Mutant mice with calcium-sensing receptor activation have hyperglycemia that is rectified by calcilytic therapy. Endocrinology. 2017;158:2486–2502. doi: 10.1210/en.2017-00111. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Babinsky VN, Hannan FM, Youhanna SC, Marechal C, Jadoul M, Devuyst O, Thakker RV. Association studies of calcium-sensing receptor (CaSR) polymorphisms with serum concentrations of glucose and phosphate, and vascular calcification in renal transplant recipients. PloS one. 2015;10:e0119459. doi: 10.1371/journal.pone.0119459. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Bai M, Pearce SH, Kifor O, Trivedi S, Stauffer UG, Thakker RV, Brown EM, Steinmann B. In vivo and in vitro characterization of neonatal hyperparathyroidism resulting from a de novo, heterozygous mutation in the Ca2+-sensing receptor gene: normal maternal calcium homeostasis as a cause of secondary hyperparathyroidism in familial benign hypocalciuric hypercalcemia. J Clin Invest. 1997;99:88–96. doi: 10.1172/JCI119137. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Bai M, Trivedi S, Brown EM. Dimerization of the extracellular calcium-sensing receptor (CaR) on the cell surface of CaR-transfected HEK293 cells. J Biol Chem. 1998a;273:23605–23610. doi: 10.1074/jbc.273.36.23605. [DOI] [PubMed] [Google Scholar]
- Bai M, Trivedi S, Lane CR, Yang Y, Quinn SJ, Brown EM. Protein kinase C phosphorylation of threonine at position 888 in Ca2+ o-sensing receptor (CaR) inhibits coupling to Ca2+ store release. J Biol Chem. 1998b;273:21267–21275. doi: 10.1074/jbc.273.33.21267. [DOI] [PubMed] [Google Scholar]
- Bai S, Mao M, Tian L, Yu Y, Zeng J, Ouyang K, Yu L, Li L, Wang D, Deng X, Wei C, Luo Y. Calcium sensing receptor mediated the excessive generation of β-amyloid peptide induced by hypoxia in vivo and in vitro. Biochem Biophys Res Commun. 2015;459:568–573. doi: 10.1016/j.bbrc.2015.02.141. [DOI] [PubMed] [Google Scholar]
- Ballesteros JA, Weinstein H. Integrated methods for the construction of three-dimensional models and computational probing of structure-function relations in G protein- coupled receptors. In: Stuart CS, editor. Methods in Neurosciences. Academic Press; San Diego: 1995. pp. 366–428. [Google Scholar]
- Bikle D, Bräuner-Osborne H, Brown EM, Chang W, Conigrave A, Hannan F, Leach K, Riccardi D, Shoback D, Ward DT, Yarova P. Calcium-sensing receptor (version 2019.4) in the IUPHAR/BPS Guide to Pharmacology Database. IUPHAR/BPS Guide to Pharmacology CITE. 2019;2019(4) doi: 10.2218/gtopdb/F12/2019.4. [DOI] [Google Scholar]
- Bikle DD, Oda Y, Tu CL, Jiang Y. Novel mechanisms for the vitamin D receptor (VDR) in the skin and in skin cancer. J Steroid Biochem Mol Biol. 2015;148:47–51. doi: 10.1016/j.jsbmb.2014.10.017. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Binmahfouz LS, Centeno PP, Conigrave AD, Ward DT. Identification of serine-875 as an inhibitory phosphorylation site in the calcium-sensing receptor. Mol Pharmacol. 2019;96:204–211. doi: 10.1124/mol.119.116178. [DOI] [PubMed] [Google Scholar]
- Bjarnadóttir TK, Fredriksson R, Schiöth HB. The gene repertoire and the common evolutionary history of glutamate, pheromone (V2R), taste(1) and other related G protein- coupled receptors. Gene. 2005;362:70–84. doi: 10.1016/j.gene.2005.07.029. [DOI] [PubMed] [Google Scholar]
- Block GA, Bushinsky DA, Cheng S, Cunningham J, Dehmel B, Drueke TB, Ketteler M, Kewalramani R, Martin KJ, Moe SM, Patel UD, Silver J, Sun Y, Wang H, Chertow GM. Effect of etelcalcetide vs cinacalcet on serum parathyroid hormone in patients receiving hemodialysis with secondary hyperparathyroidism: A randomized clinical trial. JAMA. 2017;317:156–164. doi: 10.1001/jama.2016.19468. [DOI] [PubMed] [Google Scholar]
- Bohr DF. Vascular smooth muscle: Dual effect of calcium. Science. 1963;139:597–599. doi: 10.1126/science.139.3555.597. [DOI] [PubMed] [Google Scholar]
- Bonnelye E, Chabadel A, Saltel F, Jurdic P. Dual effect of strontium ranelate: stimulation of osteoblast differentiation and inhibition of osteoclast formation and resorption in vitro. Bone. 2008;42:129–138. doi: 10.1016/j.bone.2007.08.043. [DOI] [PubMed] [Google Scholar]
- Boudot C, Hénaut L, Thiem U, Geraci S, Galante M, Saldanha P, Saidak Z, Six I, Clézardin P, Kamel S, Mentaverri R. Overexpression of a functional calcium-sensing receptor dramatically increases osteolytic potential of MDA-MB-231 cells in a mouse model of bone metastasis through epiregulin-mediated osteoprotegerin downregulation. Oncotarget. 2017;8:56460–56472. doi: 10.18632/oncotarget.16999. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Bradbury RA, Sunn KL, Crossley M, Bai M, Brown EM, Delbridge L, Conigrave AD. Expression of the parathyroid Ca2+-sensing receptor in cytotrophoblasts from human term placenta. J Endocrinol. 1998;156:425–430. doi: 10.1677/joe.0.1560425. [DOI] [PubMed] [Google Scholar]
- Braga VM, Hodivala KJ, Watt FM. Calcium-induced changes in distribution and solubility of cadherins, integrins and their associated cytoplasmic proteins in human keratinocytes. Cell Adhes Commun. 1995;3:201–215. doi: 10.3109/15419069509081287. [DOI] [PubMed] [Google Scholar]
- Bräuner-Osborne H, Jensen AA, Sheppard PO, O'Hara P, Krogsgaard-Larsen P. The agonist-binding domain of the calcium-sensing receptor is located at the amino-terminal domain. J Biol Chem. 1999;274:18382–18386. doi: 10.1074/jbc.274.26.18382. [DOI] [PubMed] [Google Scholar]
- Breitwieser G. The calcium sensing receptor life cycle: trafficking, cell surface expression, and degradation. Best Pract Res Clin Endocrinol Metab. 2013;27:303–313. doi: 10.1016/j.beem.2013.03.003. [DOI] [PubMed] [Google Scholar]
- Brennan SC, Davies TS, Schepelmann M, Riccardi D. Emerging roles of the extracellular calcium-sensing receptor in nutrient sensing: control of taste modulation and intestinal hormone secretion. Br J Nutr. 2014;111 Suppl 1:S16–22. doi: 10.1017/S0007114513002250. [DOI] [PubMed] [Google Scholar]
- Brennan TC, Rybchyn MS, Green W, Atwa S, Conigrave AD, Mason RS. Osteoblasts play key roles in the mechanisms of action of strontium ranelate. Brit J Pharmacol. 2009;157:1291–1300. doi: 10.1111/j.1476-5381.2009.00305.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Brini M, Ottolini D, Calì T, Carafoli E. Chapter 4. Calcium in Health and Disease. In: Sigel A, Sigel H, Sigel RKO, editors. Interrelations between essential metal ions and human diseases Metal ions in life sciences. Springer; 2013. pp. 81–137. [DOI] [PubMed] [Google Scholar]
- Broadhead GK, Mun HC, Avlani VA, Jourdon O, Church WB, Christopoulos A, Delbridge L, Conigrave AD. Allosteric modulation of the calcium-sensing receptor by γ-glutamyl peptides: inhibition of PTH secretion, suppression of intracellular cAMP levels, and a common mechanism of action with L-amino acids. J Biol Chem. 2011;286:8786–8797. doi: 10.1074/jbc.M110.149724. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Brown EM. Four-parameter model of the sigmoidal relationship between parathyroid hormone release and extracellular calcium concentration in normal and abnormal parathyroid tissue. J Clin Endocrinol Metab. 1983;56:572–581. doi: 10.1210/jcem-56-3-572. [DOI] [PubMed] [Google Scholar]
- Brown EM. Extracellular Ca2+ sensing, regulation of parathyroid cell function, and role of Ca2+ and other ions as extracellular (first) messengers. Physiol Rev. 1991;71:371–411. doi: 10.1152/physrev.1991.71.2.371. [DOI] [PubMed] [Google Scholar]
- Brown EM. Role of the calcium-sensing receptor in extracellular calcium homeostasis. Best Pract Res Clin Endocrinol Metab. 2013;27:333–343. doi: 10.1016/j.beem.2013.02.006. [DOI] [PubMed] [Google Scholar]
- Brown EM, Butters R, Katz C, Kifor O, Fuleihan GE. A comparison of the effects of concanavalin-A and tetradecanoylphorbol acetate on the modulation of parathyroid function by extracellular calcium and neomycin in dispersed bovine parathyroid cells. Endocrinology. 1992;130:3143–3151. doi: 10.1210/endo.130.6.1317777. [DOI] [PubMed] [Google Scholar]
- Brown EM, Enyedi P, Leboff M, Rotberg J, Preston J, Chen C. High extracellular Ca2+ and Mg2+ stimulate accumulation of inositol phosphates in bovine parathyroid cells. FEBS Lett. 1987a;218:113–118. doi: 10.1016/0014-5793(87)81029-3. [DOI] [PubMed] [Google Scholar]
- Brown EM, Fuleihan GE, Chen CJ, Kifor O. A comparison of the effects of divalent and trivalent cations on parathyroid hormone release, 3’,5’-cyclic-adenosine monophosphate accumulation, and the levels of inositol phosphates in bovine parathyroid cells. Endocrinology. 1990;127:1064–1071. doi: 10.1210/endo-127-3-1064. [DOI] [PubMed] [Google Scholar]
- Brown EM, Gamba G, Riccardi D, Lombardi M, Butters R, Kifor O, Sun A, Hediger MA, Lytton J, Hebert SC. Cloning and characterization of an extracellular Ca2+-sensing receptor from bovine parathyroid. Nature. 1993;366:575–580. doi: 10.1038/366575a0. [DOI] [PubMed] [Google Scholar]
- Brown EM, MacLeod RJ. Extracellular calcium sensing and extracellular calcium signaling. Physiol Rev. 2001;81:239–297. doi: 10.1152/physrev.2001.81.1.239. [DOI] [PubMed] [Google Scholar]
- Brown EM, Redgrave J, Thatcher J. Effect of the phorbol ester TPA on PTH secretion. Evidence for a role for protein kinase C in the control of PTH release. FEBS Lett. 1984;175:72–75. doi: 10.1016/0014-5793(84)80572-4. [DOI] [PubMed] [Google Scholar]
- Brown EM, Watson EJ, Thatcher JG, Koletsky R, Dawson-Hughes BF, Posillico JT, Shoback DM. Ouabain and low extracellular potassium inhibit PTH secretion from bovine parathyroid cells by a mechanism that does not involve increases in the cytosolic calcium concentration. Metabolism. 1987b;36:36–42. doi: 10.1016/0026-0495(87)90060-6. [DOI] [PubMed] [Google Scholar]
- Bruce JI, Yang X, Ferguson CJ, Elliott AC, Steward MC, Case RM, Riccardi D. Molecular and functional identification of a Ca2+ (polyvalent cation)-sensing receptor in rat pancreas. J Biol Chem. 1999;274:20561–20568. doi: 10.1074/jbc.274.29.20561. [DOI] [PubMed] [Google Scholar]
- Bu L, Michino M, Wolf RM, Brooks CL. Improved model building and assessment of the Calcium-sensing receptor transmembrane domain. Proteins. 2008;71:215–226. doi: 10.1002/prot.21685. [DOI] [PubMed] [Google Scholar]
- Burton DW, Foster M, Johnson KA, Hiramoto M, Deftos LJ, Terkeltaub R. Chondrocyte calcium-sensing receptor expression is up-regulated in early guinea pig knee osteoarthritis and modulates PTHrP, MMP-13, and TIMP-3 expression. Osteoarthr Cartilage. 2005;13:395–404. doi: 10.1016/j.joca.2005.01.002. [DOI] [PubMed] [Google Scholar]
- Busque SM, Kerstetter JE, Geibel JP, Insogna K. L-type amino acids stimulate gastric acid secretion by activation of the calcium-sensing receptor in parietal cells. Am J Physiol Gastrointest Liver Physiol. 2005;289:G664–669. doi: 10.1152/ajpgi.00096.2005. [DOI] [PubMed] [Google Scholar]
- Butters RR, Jr, Chattopadhyay N, Nielsen P, Smith CP, Mithal A, Kifor O, Bai M, Quinn S, Goldsmith P, Hurwitz S, Krapcho K, Busby J, Brown EM. Cloning and characterization of a calcium-sensing receptor from the hypercalcemic New Zealand white rabbit reveals unaltered responsiveness to extracellular calcium. J Bone Miner Res. 1997;12:568–579. doi: 10.1359/jbmr.1997.12.4.568. [DOI] [PubMed] [Google Scholar]
- Campion KL, McCormick WD, Warwicker J, Khayat ME, Atkinson-Dell R, Steward MC, Delbridge LW, Mun HC, Conigrave AD, Ward DT. Pathophysiologic changes in extracellular pH modulate parathyroid calcium-sensing receptor activity and secretion via a histidine-independent mechanism. J Am Soc Nephrol. 2015;26:2163–2171. doi: 10.1681/ASN.2014070653. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Canaff L, Hendy GN. Human calcium-sensing receptor gene. Vitamin D response elements in promoters P1 and P2 confer transcriptional responsiveness to 1,25- dihydroxyvitamin D. J Biol Chem. 2002;277:30337–30350. doi: 10.1074/jbc.M201804200. [DOI] [PubMed] [Google Scholar]
- Canaff L, Zhou X, Mosesova I, Cole DE, Hendy GN. Glial cells missing-2 (GCM2) transactivates the calcium-sensing receptor gene: effect of a dominant-negative GCM2 mutant associated with autosomal dominant hypoparathyroidism. Hum Mutat. 2009;30:85–92. doi: 10.1002/humu.20827. [DOI] [PubMed] [Google Scholar]
- Capasso G, Geibel PJ, Damiano S, Jaeger P, Richards WG, Geibel JP. The calcium sensing receptor modulates fluid reabsorption and acid secretion in the proximal tubule. Kidney Int. 2013;84:277–284. doi: 10.1038/ki.2013.137. [DOI] [PubMed] [Google Scholar]
- Casala C, Gil-Guinon E, Ordonez JL, Miguel-Queralt S, Rodriguez E, Galvan P, Lavarino C, Munell F, de Alava E, Mora J, de Torres C. The calcium-sensing receptor is silenced by genetic and epigenetic mechanisms in unfavorable neuroblastomas and its reactivation induces ERK1/2-dependent apoptosis. Carcinogenesis. 2013;34:268–276. doi: 10.1093/carcin/bgs338. [DOI] [PubMed] [Google Scholar]
- Casti A, Orlandini G, Reali N, Bacciottini F, Vanelli M, Bernasconi S. Pattern of blood polyamines in healthy subjects from infancy to the adult age. J Endocrinol Invest. 1982;5:263–266. doi: 10.1007/BF03348334. [DOI] [PubMed] [Google Scholar]
- Cavaco BM, Canaff L, Nolin-Lapalme A, Vieira M, Silva TN, Saramago A, Domingues R, Rutter MM, Hudon J, Gleason JL, Leite V, Hendy GN. Homozygous calcium-sensing receptor polymorphism R544Q presents as hypocalcemic hypoparathyroidism. J Clin Endocrinol Metab. 2018;103:2879–2888. doi: 10.1210/jc.2017-02407. [DOI] [PubMed] [Google Scholar]
- Ceglia L, Harris SS, Rasmussen HM, Dawson-Hughes B. Activation of the calcium sensing receptor stimulates gastrin and gastric acid secretion in healthy participants. Osteoporos Int. 2009;20:71–78. doi: 10.1007/s00198-008-0637-8. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Celli A, Sanchez S, Behne M, Hazlett T, Gratton E, Mauro T. The epidermal Ca(2+) gradient: Measurement using the phasor representation of fluorescent lifetime imaging. Biophysical journal. 2010;98:911–921. doi: 10.1016/j.bpj.2009.10.055. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Centeno PP, Herberger A, Mun H-C, Tu C, Nemeth EF, Chang W, Conigrave AD, Ward DT. Phosphate acts directly on the calcium-sensing receptor to stimulate parathyroid hormone secretion. Nat Commun. 2019;10 doi: 10.1038/s41467-019-12399-9. 4693. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Chang W, Pratt S, Chen TH, Bourguignon L, Shoback D. Amino acids in the cytoplasmic C terminus of the parathyroid Ca2+-sensing receptor mediate efficient cell-surface expression and phospholipase C activation. J Biol Chem. 2001;276:44129–44136. doi: 10.1074/jbc.M104834200. [DOI] [PubMed] [Google Scholar]
- Chang W, Tu C, Chen TH, Bikle D, Shoback D. The extracellular calcium-sensing receptor (CaSR) is a critical modulator of skeletal development. Sci Signal. 2008;1 doi: 10.1126/scisignal.1159945. ra1. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Chang W, Tu C, Chen TH, Komuves L, Oda Y, Pratt SA, Miller S, Shoback D. Expression and signal transduction of calcium-sensing receptors in cartilage and bone. Endocrinology. 1999;140:5883–5893. doi: 10.1210/endo.140.12.7190. [DOI] [PubMed] [Google Scholar]
- Chang W, Tu C, Cheng Z, Rodriguez L, Chen TH, Gassmann M, Bettler B, Margeta M, Jan LY, Shoback D. Complex formation with the Type B gamma-aminobutyric acid receptor affects the expression and signal transduction of the extracellular calcium-sensing receptor. Studies with HEK-293 cells and neurons. J Biol Chem. 2007;282:25030–25040. doi: 10.1074/jbc.M700924200. [DOI] [PubMed] [Google Scholar]
- Chang WH, Pratt S, Chen TH, Nemeth E, Huang ZM, Shoback D. Coupling of calcium receptors to inositol phosphate and cyclic AMP generation in mammalian cells and Xenopus laevis oocytes and immunodetection of receptor protein by region-specific antipeptide antisera. J Bone Miner Res. 1998;13:570–580. doi: 10.1359/jbmr.1998.13.4.570. [DOI] [PubMed] [Google Scholar]
- Chappell MD, Li R, Smith SC, Dressman BA, Tromiczak EG, Tripp AE, Blanco MJ, Vetman T, Quimby SJ, Matt J, Britton TC, et al. Discovery of (1S,2R,3S,4S,5R,6R)-2-Amino-3-[(3,4- difluorophenyl)sulfanylmethyl]-4-hydroxy-bicy clo[3.1.0]hexane-2,6-dicarboxylic Acid Hydrochloride (LY3020371.HCl): A potent, metabotropic glutamate 2/3 receptor antagonist with antidepressant-like activity. J Med Chem. 2016;59:10974–10993. doi: 10.1021/acs.jmedchem.6b01119. [DOI] [PubMed] [Google Scholar]
- Chattopadhyay N, Espinosa-Jeffrey A, Tfelt-Hansen J, Yano S, Bandyopadhyay S, Brown EM, de Vellis J. Calcium receptor expression and function in oligodendrocyte commitment and lineage progression: potential impact on reduced myelin basic protein in CaR-null mice. J Neurosci Res. 2008;86:2159–2167. doi: 10.1002/jnr.21662. [DOI] [PubMed] [Google Scholar]
- Chattopadhyay N, Yano S, Tfelt-Hansen J, Rooney P, Kanuparthi D, Bandyopadhyay S, Ren XH, Terwilliger E, Brown EM. Mitogenic action of calcium-sensing receptor on rat calvarial osteoblasts. Endocrinology. 2004;145:3451–3462. doi: 10.1210/en.2003-1127. [DOI] [PubMed] [Google Scholar]
- Chattopadhyay N, Ye CP, Yamaguchi T, Kifor O, Vassilev PM, Nishimura R, Brown EM. Extracellular calcium-sensing receptor in rat oligodendrocytes: expression and potential role in regulation of cellular proliferation and an outward K+ channel. Glia. 1998;24:449–458. [PubMed] [Google Scholar]
- Chen CJ, Anast CS, Brown EM. High osmolality: a potent parathyroid hormone secretogogue in dispersed parathyroid cells. Endocrinology. 1987;121:958–964. doi: 10.1210/endo-121-3-958. [DOI] [PubMed] [Google Scholar]
- Chen CJ, Barnett JV, Congo DA, Brown EM. Divalent cations suppress 3’,5’-adenosine monophosphate accumulation by stimulating a pertussis toxin-sensitive guanine nucleotide- binding protein in cultured bovine parathyroid cells. Endocrinology. 1989;124:233–239. doi: 10.1210/endo-124-1-233. [DOI] [PubMed] [Google Scholar]
- Chen R, Goodman W. Role of the calcium-sensing receptor in parathyroid gland physiology. Am J Physiol Renal Physiol. 2004;286:F1005–F1011. doi: 10.1152/ajprenal.00013.2004. [DOI] [PubMed] [Google Scholar]
- Cheng I, Klingensmith ME, Chattopadhyay N, Kifor O, Butters RR, Soybel DI, Brown EM. Identification and localization of the extracellular calcium-sensing receptor in human breast. J Clin Endocrinol Metab. 1998;83:703–707. doi: 10.1210/jcem.83.2.4558. [DOI] [PubMed] [Google Scholar]
- Cheng SX. Calcium-sensing receptor inhibits secretagogue-induced electrolyte secretion by intestine via the enteric nervous system. Am J Physiol Gastrointest Liver Physiol. 2012;303:G60–70. doi: 10.1152/ajpgi.00425.2011. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Cheng SX, Lightfoot YL, Yang T, Zadeh M, Tang L, Sahay B, Wang GP, Owen JL, Mohamadzadeh M. Epithelial CaSR deficiency alters intestinal integrity and promotes proinflammatory immune responses. FEBS Lett. 2014;588:4158–4166. doi: 10.1016/j.febslet.2014.05.007. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Cheng Z, Tu C, Rodriguez L, Chen TH, Dvorak MM, Margeta M, Gassmann M, Bettler B, Shoback D, Chang W. Type B γ-aminobutyric acid receptors modulate the function of the extracellular Ca2+-sensing receptor and cell differentiation in murine growth plate chondrocytes. Endocrinology. 2007;148:4984–4992. doi: 10.1210/en.2007-0653. [DOI] [PubMed] [Google Scholar]
- Chertow GM, Block GA, Correa-Rotter R, Drueke TB, Floege J, Goodman WG, Herzog CA, Kubo Y, London GM, Mahaffey KW, Mix TC, Moe SM, Trotman ML, Wheeler DC, Parfrey PS. Effect of cinacalcet on cardiovascular disease in patients undergoing dialysis. N Engl J Med. 2012;367:2482–2494. doi: 10.1056/NEJMoa1205624. [DOI] [PubMed] [Google Scholar]
- Chiarini A, Armato U, Gardenal E, Gui L, Dal Pra I. Amyloid β-exposed human astrocytes overproduce phospho-tau and overrelease it within exosomes, effects suppressed by calcilytic NPS 2143-Further implications for Alzheimer’s therapy. Front Neurosci. 2017a;11:217. doi: 10.3389/fnins.2017.00217. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Chiarini A, Armato U, Liu D, Dal Pra I. Calcium-sensing receptor antagonist NPS 2143 restores amyloid precursor protein physiological non-amyloidogenic processing in Aβ- exposed adult human astrocytes. Sci Rep. 2017b;7 doi: 10.1038/s41598-017-01215-3. 1277. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Chikatsu N, Fukumoto S, Takeuchi Y, Suzawa M, Obara T, Matsumoto T, Fujita T. Cloning and characterization of two promoters for the human calcium-sensing receptor (CaSR) and changes of CaSR expression in parathyroid adenomas. J Biol Chem. 2000;275:7553–7557. doi: 10.1074/jbc.275.11.7553. [DOI] [PubMed] [Google Scholar]
- Chou YH, Brown EM, Levi T, Crowe G, Atkinson AB, Arnqvist HJ, Toss G, Fuleihan GE, Seidman JG, Seidman CE. The gene responsible for familial hypocalciuric hypercalcemia maps to chromosome 3q in four unrelated families. Nat Genet. 1992;1:295–300. doi: 10.1038/ng0792-295. [DOI] [PubMed] [Google Scholar]
- Christopher JA, Aves SJ, Bennett KA, Dore AS, Errey JC, Jazayeri A, Marshall FH, Okrasa K, Serrano-Vega MJ, Tehan BG, Wiggin GR, Congreve M. Fragment and structure- based drug discovery for a class C GPCR: Discovery of the mGlu5 negative allosteric modulator HTL14242 (3-Chloro-5-[6-(5-fluoropyridin-2-yl)pyrimidin-4-yl]benzonitrile) J Med Chem. 2015;58:6653–6664. doi: 10.1021/acs.jmedchem.5b00892. [DOI] [PubMed] [Google Scholar]
- Christopher JA, Orgovan Z, Congreve M, Dore AS, Errey JC, Marshall FH, Mason JS, Okrasa K, Rucktooa P, Serrano-Vega MJ, Ferenczy GG, Keseru GM. Structure-based optimization strategies for G protein-coupled receptor (GPCR) allosteric modulators: A case study from analyses of new metabotropic glutamate receptor 5 (mGlu5) X-ray structures. J Med Chem. 2019;62:207–222. doi: 10.1021/acs.jmedchem.7b01722. [DOI] [PubMed] [Google Scholar]
- Colloton M, Shatzen E, Miller G, Stehman-Breen C, Wada M, Lacey D, Martin D. Cinacalcet HCl attenuates parathyroid hyperplasia in a rat model of secondary hyperparathyroidism. Kidney Int. 2005;67:467–476. doi: 10.1111/j.1523-1755.2005.67103.x. [DOI] [PubMed] [Google Scholar]
- Conigrave A. The calcium-sensing receptor and the parathyroid: Past, present, future. Front Physiol. 2016;7(563):1–13. doi: 10.3389/fphys.2016.00563. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Conigrave AD, Brown EM, Rizzoli R. Dietary protein and bone health: Roles of amino acid–sensing receptors in the control of calcium metabolism and bone homeostasis. Annu Rev Nutr. 2008;28:131–155. doi: 10.1146/annurev.nutr.28.061807.155328. [DOI] [PubMed] [Google Scholar]
- Conigrave AD, Franks AH, Brown EM, Quinn SJ. L-Amino acid sensing by the calcium- sensing receptor: a general mechanism for coupling protein and calcium metabolism? Eur J Clin Nutr. 2002;56:1072–1080. doi: 10.1038/sj.ejcn.1601463. [DOI] [PubMed] [Google Scholar]
- Conigrave AD, Hampson DR. Broad-spectrum amino acid sensing by class 3 G-protein coupled receptors. Trends Endocrinol Metab. 2006;17:398–407. doi: 10.1016/j.tem.2006.10.012. [DOI] [PubMed] [Google Scholar]
- Conigrave AD, Hampson DR. Broad-spectrum amino acid-sensing class C G-protein coupled receptors: molecular mechanisms, physiological significance and options for drug development. Pharmacol Therap. 2010;127:252–260. doi: 10.1016/j.pharmthera.2010.04.007. [DOI] [PubMed] [Google Scholar]
- Conigrave AD, Mun H-C, Delbridge L, Quinn SJ, Wilkinson M, Brown EM. L-amino acids regulate parathyroid hormone secretion. J Biol Chem. 2004;279:38151–38159. doi: 10.1074/jbc.M406373200. [DOI] [PubMed] [Google Scholar]
- Conigrave AD, Mun HC, Lok HC. Aromatic L-amino acids activate the calcium-sensing receptor. J Nutr. 2007;137:1524S–1527S. doi: 10.1093/jn/137.6.1524S. discussion 1548S. [DOI] [PubMed] [Google Scholar]
- Conigrave AD, Quinn SJ, Brown EM. L-amino acid sensing by the extracellular Ca2+- sensing receptor. Proc Natl Acad Sci. 2000;97:4814–4819. doi: 10.1073/pnas.97.9.4814. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Conigrave AD, Ward DT. Calcium-sensing receptor (CaSR): pharmacological properties and signaling pathways. Best Pract Res Clin Endocrinol Metab. 2013;27:315–331. doi: 10.1016/j.beem.2013.05.010. [DOI] [PubMed] [Google Scholar]
- Conley YP, Mukherjee A, Kammerer C, DeKosky ST, Kamboh MI, Finegold DN, Ferrell RE. Evidence supporting a role for the calcium-sensing receptor in Alzheimer disease. Am J Med Genet B Neuropsychiatr Genet. 2009;150B:703–709. doi: 10.1002/ajmg.b.30896. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Cook AE, Mistry SN, Gregory KJ, Furness SG, Sexton PM, Scammells PJ, Conigrave AD, Christopoulos A, Leach K. Biased allosteric modulation at the CaSR engendered by structurally diverse calcimimetics. Br J Pharmacol. 2015;172:185–200. doi: 10.1111/bph.12937. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Corbetta S, Lania A, Filopanti M, Vicentini L, Ballare E, Spada A. Mitogen-activated protein kinase cascade in human normal and tumoral parathyroid cells. J Clin Endocrinol Metab. 2002;87:2201–2205. doi: 10.1210/jcem.87.5.8492. [DOI] [PubMed] [Google Scholar]
- Courbebaisse M, Diet C, Timsit MO, Mamzer MF, Thervet E, Noel LH, Legendre C, Friedlander G, Martinez F, Prie D. Effects of cinacalcet in renal transplant patients with hyperparathyroidism. Am J Nephrol. 2012;35:341–348. doi: 10.1159/000337526. [DOI] [PubMed] [Google Scholar]
- Dal Pra I, Armato U, Chioffi F, Pacchiana R, Whitfield JF, Chakravarthy B, Gui L, Chiarini A. The Aβ peptides-activated calcium-sensing receptor stimulates the production and secretion of vascular endothelial growth factor-A by normoxic adult human cortical astrocytes. Neuromolecular Med. 2014;16:645–657. doi: 10.1007/s12017-014-8315-9. [DOI] [PubMed] [Google Scholar]
- Davey AE, Leach K, Valant C, Conigrave AD, Sexton PM, Christopoulos A. Positive and negative allosteric modulators promote biased signaling at the calcium-sensing receptor. Endocrinology. 2012;153:1232–1241. doi: 10.1210/en.2011-1426. [DOI] [PubMed] [Google Scholar]
- Davies SL, Gibbons CE, Vizard T, Ward DT. Calcium-sensing receptor induces rho kinase-mediated actin stress fiber assembly and altered cell morphology though not in response to aromatic amino acids. Am J Physiol Cell Physiol. 2006;290:1543–1551. doi: 10.1152/ajpcell.00482.2005. [DOI] [PubMed] [Google Scholar]
- Davies SL, Ozawa A, McCormick WD, Dvorak MM, Ward DT. Protein kinase C- mediated phosphorylation of the calcium-sensing receptor is stimulated by receptor activation and attenuated by calyculin-sensitive phosphatase activity. J Biol Chem. 2007;282:15048–15056. doi: 10.1074/jbc.M607469200. [DOI] [PubMed] [Google Scholar]
- Dawson-Hughes B. Calcium and protein in bone health. Proc Nutr Soc. 2003;62:505–509. doi: 10.1079/pns2003267. [DOI] [PubMed] [Google Scholar]
- de Torres C, Beleta H, Diaz R, Toran N, Rodriguez E, Lavarino C, Garcia I, Acosta S, Sunol M, Mora J. The calcium-sensing receptor and parathyroid hormone-related protein are expressed in differentiated, favorable neuroblastic tumors. Cancer. 2009;115:2792–2803. doi: 10.1002/cncr.24304. [DOI] [PubMed] [Google Scholar]
- Deprez P, Temal T, Jary H, Auberval M, Lively S, Guedin D, Vevert JP. New potent calcimietics: II. Discovery of benzothiazole trisubstituted ureas. Bioorg Med Chem. 2013;23:2455–2459. doi: 10.1016/j.bmcl.2013.01.077. [DOI] [PubMed] [Google Scholar]
- Diakogiannaki E, Pais R, Tolhurst G, Parker HE, Horscroft J, Rauscher B, Zietek T, Daniel H, Gribble FM, Reimann F. Oligopeptides stimulate glucagon-like peptide-1 secretion in mice through proton-coupled uptake and the calcium-sensing receptor. Diabetologia. 2013;56:2688–2696. doi: 10.1007/s00125-013-3037-3. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Diaz R, Hurwitz S, Chattopadhyay N, Pines M, Yang Y, Kifor O, Einat MS, Butters R, Hebert SC, Brown EM. Cloning, expression, and tissue localization of the calcium-sensing receptor in chicken (Gallus domesticus) Am J Physiol. 1997;273:R1008–1016. doi: 10.1152/ajpregu.1997.273.3.R1008. [DOI] [PubMed] [Google Scholar]
- Didiuk MT, Griffith DA, Benbow JW, Liu KK, Walker DP, Bi FC, Morris J, Guzman-Perez A, Gao H, Bechle BM, Kelley RM, Yang X, Dirico K, Ahmed S, Hungerford W, DiBrinno J, Zawistoski MP, Bagley SW, Li J, Zeng Y, Santucci S, Oliver R, Corbett M, Olson T, Chen C, Li M, Paralkar VM, Riccardi KA, Healy DR, Kalgutkar AS, Maurer TS, Nguyen HT, Frederick KS. Short-acting 5-(trifluoromethyl)pyrido[4,3-d]pyrimidin-4(3H)-one derivatives as orally-active calcium-sensing receptor antagonists. Bioorg Med Chem Lett. 2009;19:4555–4559. doi: 10.1016/j.bmcl.2009.07.004. [DOI] [PubMed] [Google Scholar]
- Diepenhorst NA, Leach K, Keller AN, Rueda P, Cook AE, Pierce TL, Nowell C, Pastoureau P, Sabatini M, Summers RJ, Charman WN, Sexton PM, Christopoulos A, Langmead CJ. Divergent effects of strontium and calcium-sensing receptor positive allosteric modulators (calcimimetics) on human osteoclast activity. Br J Pharmacol. 2018;175:4095–4108. doi: 10.1111/bph.14344. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Dobnig H, Turner RT. The effects of programmed administration of human parathyroid hormone fragment (1-34) on bone histomorphometry and serum chemistry in rats. Endocrinology. 1997;138:4607–4612. doi: 10.1210/endo.138.11.5505. [DOI] [PubMed] [Google Scholar]
- Dong B, Endo I, Ohnishi Y, Kondo T, Hasegawa T, Amizuka N, Kiyonari H, Shioi G, Abe M, Fukumoto S, Matsumoto T. Calcilytic Ameliorates Abnormalities of Mutant Calcium-Sensing Receptor (CaSR) Knock-in Mice Mimicking Autosomal Dominant Hypocalcemia (ADH) J Bone Miner Res. 2015;30:1980–1993. doi: 10.1002/jbmr.2551. [DOI] [PubMed] [Google Scholar]
- Dore AS, Okrasa K, Patel JC, Serrano-Vega M, Bennett K, Cooke RM, Errey JC, Jazayeri A, Khan S, Tehan B, Weir M, Wiggin GR, Marshall FH. Structure of class C GPCR metabotropic glutamate receptor 5 transmembrane domain. Nature. 2014;511:557–562. doi: 10.1038/nature13396. [DOI] [PubMed] [Google Scholar]
- Dubois-Ferrière V, Brennan T, Dayer R, Rizzoli R, Ammann P. Calcitropic hormones and IGF-I are influenced by dietary protein. Endocrinology. 2011;152:1839–1847. doi: 10.1210/en.2010-1079. [DOI] [PubMed] [Google Scholar]
- Dufner MM, Kirchhoff P, Remy C, Hafner P, Muller MK, Cheng SX, Tang LQ, Hebert SC, Geibel JP, Wagner CA. The calcium-sensing receptor acts as a modulator of gastric acid secretion in freshly isolated human gastric glands. Am J Physiol Gastrointest Liver Physiol. 2005;289:G1084–1090. doi: 10.1152/ajpgi.00571.2004. [DOI] [PubMed] [Google Scholar]
- Dvorak MM, Chen TH, Orwoll B, Garvey C, Chang W, Bikle DD, Shoback DM. Constitutive activity of the osteoblast Ca2+-sensing receptor promotes loss of cancellous bone. Endocrinology. 2007;148:3156–3163. doi: 10.1210/en.2007-0147. [DOI] [PubMed] [Google Scholar]
- Dvorak-Ewell MM, Chen TH, Liang N, Garvey C, Liu B, Tu C, Chang W, Bikle DD, Shoback DM. Osteoblast extracellular Ca2+-sensing receptor regulates bone development, mineralization, and turnover. J Bone Miner Res. 2011;26:2935–2947. doi: 10.1002/jbmr.520. [DOI] [PMC free article] [PubMed] [Google Scholar]
- El Hiani Y, Ahidouch A, Roudbaraki M, Guenin S, Brule G, Ouadid-Ahidouch H. Calcium-sensing receptor stimulation induces nonselective cation channel activation in breast cancer cells. J Membr Biol. 2006;211:127–137. doi: 10.1007/s00232-006-0017-2. [DOI] [PubMed] [Google Scholar]
- Elajnaf T, Iamartino L, Mesteri I, Muller C, Bassetto M, Mangardt T, Baumgartner-Parzer S, Kallay E, Schepelmann M. Nutritional and pharmacological targeting of the calcium- sensing receptor influences chemically induced colitis in mice. Nutrients. 2019;11 doi: 10.3390/nu11123072. pii: E3072. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Ellison TI, Smith MK, Gilliam AC, MacDonald PN. Inactivation of the vitamin D receptor enhances susceptibility of murine skin to UV-induced tumorigenesis. J Invest Dermatol. 2008;128:2508–2517. doi: 10.1038/jid.2008.131. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Engelstoft MS, Park WM, Sakata I, Kristensen LV, Husted AS, Osborne-Lawrence S, Piper PK, Walker AK, Pedersen MH, Nohr MK, Pan J, Sinz CJ, et al. Seven transmembrane G protein-coupled receptor repertoire of gastric ghrelin cells. Mol Metab. 2013;2:376–392. doi: 10.1016/j.molmet.2013.08.006. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Fajtova VT, Quinn SJ, Brown EM. Cytosolic calcium responses of single rMTC 44-2 cells to stimulation with external calcium and potassium. Am J Physiol. 1991;261:E151–E158. doi: 10.1152/ajpendo.1991.261.1.E151. [DOI] [PubMed] [Google Scholar]
- Fan GF, Ray K, Zhao XM, Goldsmith PK, Spiegel AM. Mutational analysis of the cysteines in the extracellular domain of the human Ca2+ receptor: effects on cell surface expression, dimerization and signal transduction. FEBS Lett. 1998;436:353–356. doi: 10.1016/s0014-5793(98)01165-x. [DOI] [PubMed] [Google Scholar]
- Fan Y, Liu W, Bi R, Densmore MJ, Sato T, Mannstadt M, Yuan Q, Zhou X, Olauson H, Larsson TE, Toka HR, et al. Interrelated role of Klotho and calcium-sensing receptor in parathyroid hormone synthesis and parathyroid hyperplasia. Proc Natl Acad Sci USA. 2018;115:E3749–E3758. doi: 10.1073/pnas.1717754115. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Feng J, Petersen CD, Coy DH, Jiang JK, Thomas CJ, Pollak MR, Wank SA. Calcium- sensing receptor is a physiologic multimodal chemosensor regulating gastric G-cell growth and gastrin secretion. Proc Natl Acad Sci USA. 2010;107:17791–17796. doi: 10.1073/pnas.1009078107. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Ferry S, Traiffort E, Stinnakre J, Ruat M. Developmental and adult expression of rat calcium-sensing receptor transcripts in neurons and oligodendrocytes. Eur J Neurosci. 2000;12:872–884. doi: 10.1046/j.1460-9568.2000.00980.x. [DOI] [PubMed] [Google Scholar]
- Fetahu IS, Hummel DM, Manhardt T, Aggarwal A, Baumgartner-Parzer S, Kallay E. Regulation of the calcium-sensing receptor expression by 1,25-dihydroxyvitamin D3, interleukin-6, and tumor necrosis factor alpha in colon cancer cells. J Steroid Biochem Mol Biol. 2014;144 Pt A:228–231. doi: 10.1016/j.jsbmb.2013.10.015. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Fisher JE, Scott K, Wei N, Zhao JZ, Cusick T, Tijerina M, Karanam B, Duong L, Glantschnig H. Pharmacodynamic responses to combined treatment regimens with the calcium sensing receptor antagonist JTT-305/MK-5442 and alendronate in osteopenic ovariectomized rats. Bone. 2012;50:1332–1342. doi: 10.1016/j.bone.2012.03.004. [DOI] [PubMed] [Google Scholar]
- Fisher MM, Cabrera SM, Imel EA. Successful treatment of neonatal severe hyperparathyroidism with cinacalcet in two patients. Endocrinol Diabetes Metab Case Rep. 2015;2015 doi: 10.1530/EDM-15-0040. 150040. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Fitzpatrick LA, Dabrowski CE, Cicconetti G, Gordon DN, Fuerst T, Engelke K, Genant HK. Ronacaleret, a calcium-sensing receptor antagonist, increases trabecular but not cortical bone in postmenopausal women. J Bone Miner Res. 2012;27:255–262. doi: 10.1002/jbmr.554. [DOI] [PubMed] [Google Scholar]
- Fitzpatrick LA, Dabrowski CE, Cicconetti G, Gordon DN, Papapoulos S, Bone HG, Bilezikian JP. The effects of ronacaleret, a calcium-sensing receptor antagonist, on bone mineral density and biochemical markers of bone turnover in postmenopausal women with low bone mineral density. J Clin Endocrinol Metab. 2011a;96:2441–2449. doi: 10.1210/jc.2010-2855. [DOI] [PubMed] [Google Scholar]
- Fitzpatrick LA, Smith PL, McBride TA, Fries MA, Hossain M, Dabrowski CE, Gordon DN. Ronacaleret, a calcium-sensing receptor antagonist, has no significant effect on radial fracture healing time: results of a randomized, double-blinded, placebo-controlled Phase II clinical trial. Bone. 2011b;49:845–852. doi: 10.1016/j.bone.2011.06.017. [DOI] [PubMed] [Google Scholar]
- Freichel M, Zink-Lorenz A, Holloschi A, Hafner M, Flockerzi V, Raue F. Expression of a calcium-sensing receptor in a human medullary thyroid carcinoma cell line and its contribution to calcitonin secretion. Endocrinology. 1996;137:3842–3848. doi: 10.1210/endo.137.9.8756555. [DOI] [PubMed] [Google Scholar]
- Friedman PA. Codependence of renal calcium and sodium transport. Annu Rev Physiol. 1998;60:179–197. doi: 10.1146/annurev.physiol.60.1.179. [DOI] [PubMed] [Google Scholar]
- Fritze O, Filipek S, Kuksa V, Palczewski K, Hofmann KP, Ernst OP. Role of the conserved NPxxY(x)5,6F motif in the rhodopsin ground state and during activation. Proc Natl Acad Sci USA. 2003;100:2290–2295. doi: 10.1073/pnas.0435715100. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Fudge NJ, Kovacs CS. Physiological studies in heterozygous calcium sensing receptor (CaSR) gene-ablated mice confirm that the CaSR regulates calcitonin release in vivo. BMC Physiol. 2004;4:5. doi: 10.1186/1472-6793-4-5. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Fujisawa Y, Yamaguchi R, Satake E, Ohtaka K, Nakanishi T, Ozono K, Ogata T. Identification of AP2S1 mutation and effects of low calcium formula in an infant with hypercalcemia and hypercalciuria. J Clin Endocrinol Metab. 2013;98:E2022–2027. doi: 10.1210/jc.2013-2571. [DOI] [PubMed] [Google Scholar]
- Fukagawa M, Shimazaki R, Akizawa T, Evocalcet study g Head-to-head comparison of the new calcimimetic agent evocalcet with cinacalcet in Japanese hemodialysis patients with secondary hyperparathyroidism. Kidney Int. 2018;94:818–825. doi: 10.1016/j.kint.2018.05.013. [DOI] [PubMed] [Google Scholar]
- Gama L, Wilt SG, Breitwieser GE. Heterodimerization of calcium sensing receptors with metabotropic glutamate receptors in neurons. J Biol Chem. 2001;276:39053–39059. doi: 10.1074/jbc.M105662200. [DOI] [PubMed] [Google Scholar]
- Gannon AW, Monk HM, Levine MA. Cinacalcet monotherapy in neonatal severe hyperparathyroidism: a case study and review. J Clin Endocrinol Metab. 2014;99:7–11. doi: 10.1210/jc.2013-2834. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Garcia Soblechero E, Ferrer Castillo MT, Jimenez Crespo B, Dominguez Quintero ML, Gonzalez Fuentes C. Neonatal hypercalcemia due to a homozygous mutation in the calcium-sensing receptor: failure of cinacalcet. Neonatology. 2013;104:104–108. doi: 10.1159/000350540. [DOI] [PubMed] [Google Scholar]
- Gardenal E, Chiarini A, Armato U, Dal Pra I, Verkhratsky A, Rodriguez JJ. Increased calcium-sensing receptor immunoreactivity in the hippocampus of a triple transgenic mouse model of Alzheimer’s disease. Front Neurosci. 2017;11:81. doi: 10.3389/fnins.2017.00081. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Garrett JE, Capuano IV, Hammerland LG, Hung BC, Brown EM, Hebert SC, Nemeth EF, Fuller F. Molecular cloning and functional expression of human parathyroid calcium receptor cDNAs. J Biol Chem. 1995a;270:12919–12925. doi: 10.1074/jbc.270.21.12919. [DOI] [PubMed] [Google Scholar]
- Garrett JE, Tamir H, Kifor O, Simin RT, Rogers KV, Mithal A, Gagel RF, Brown EM. Calcitonin-secreting cells of the thyroid gland express an extracellular calcium-sensing receptor gene. Endocrinology. 1995b;136:5202–5211. doi: 10.1210/endo.136.11.7588259. [DOI] [PubMed] [Google Scholar]
- Geibel J, Sritharan K, Geibel R, Geibel P, Persing JS, Seeger A, Roepke TK, Deichstetter M, Prinz C, Cheng SX, Martin D, Hebert SC. Calcium-sensing receptor abrogates secretagogue- induced increases in intestinal net fluid secretion by enhancing cyclic nucleotide destruction. Proc Natl Acad Sci USA. 2006;103:9390–9397. doi: 10.1073/pnas.0602996103. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Geibel JP, Hebert SC. The functions and roles of the extracellular Ca2+-sensing receptor along the gastrointestinal tract. Annu Rev Physiol. 2009;71:205–217. doi: 10.1146/annurev.physiol.010908.163128. [DOI] [PubMed] [Google Scholar]
- Geng Y, Bush M, Mosyak L, Wang F, Fan QR. Structural mechanism of ligand activation in human GABA(B) receptor. Nature. 2013;504:254–259. doi: 10.1038/nature12725. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Geng Y, Mosyak L, Kurinov I, Zuo H, Sturchler E, Cheng TC, Subramanyam P, Brown AP, Brennan SC, Mun HC, Bush M, et al. Structural mechanism of ligand activation in human calcium-sensing receptor. Elife. 2016;5:e13662. doi: 10.7554/eLife.13662. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Geng Y, Xiong D, Mosyak L, Malito DL, Kniazeff J, Chen Y, Burmakina S, Quick M, Bush M, Javitch JA, Pin JP, Fan QR. Structure and functional interaction of the extracellular domain of human GABA(B) receptor GBR2. Nat Neurosci. 2012;15:970–978. doi: 10.1038/nn.3133. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Gerspacher M, Altmann E, Beerli R, Buhl T, Endres R, Gamse R, Kameni-Tcheudji J, Kneissel M, Krawinkler KH, Missbach M, Schmidt A, Seuwen K, Weiler S, Widler L. Penta- substituted benzimidazoles as potent antagonists of the calcium-sensing receptor (CaSR- antagonists) Bioorg Med Chem Lett. 2010;20:5161–5164. doi: 10.1016/j.bmcl.2010.07.016. [DOI] [PubMed] [Google Scholar]
- Giudice ML, Mihalik B, Dinnyes A, Kobolak J. The Nervous System Relevance of the Calcium Sensing Receptor in Health and Disease. Molecules. 2019;24 doi: 10.3390/molecules24142546. pii:E2546. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Goltzman D, Hendy GN. The calcium-sensing receptor in bone--mechanistic and therapeutic insights. Nat Rev Endocrinol. 2015;11:298–307. doi: 10.1038/nrendo.2015.30. [DOI] [PubMed] [Google Scholar]
- Gong Y, Hou J. Claudin-14 underlies Ca++-sensing receptor-mediated Ca++ metabolism via NFAT-microRNA-based mechanisms. J Am Soc Nephrol. 2014;25:745–760. doi: 10.1681/ASN.2013050553. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Goolam MA, Ward JH, Avlani VA, Leach K, Christopoulos A, Conigrave AD. Roles of intraloops-2 and -3 and the proximal C-terminus in signalling pathway selection from the human calcium-sensing receptor. FEBS Lett. 2014;588:3340–3346. doi: 10.1016/j.febslet.2014.07.022. [DOI] [PubMed] [Google Scholar]
- Gorvin CM. Molecular and clinical insights from studies of calcium-sensing receptor mutations. J Mol Endocrinol. 2019;63:R1–R16. doi: 10.1530/JME-19-0104. [DOI] [PubMed] [Google Scholar]
- Gorvin CM, Babinsky VN, Malinauskas T, Nissen PH, Schou AJ, Hanyaloglu AC, Siebold C, Jones EY, Hannan FM, Thakker RV. A calcium-sensing receptor mutation causing hypocalcemia disrupts a transmembrane salt bridge to activate ß-arrestin-biased signaling. Sci Signal. 2018a;11 doi: 10.1126/scisignal.aan3714. pii:eaan3714. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Gorvin CM, Cranston T, Hannan FM, Rust N, Qureshi A, Nesbit MA, Thakker RV. G- protein subunit-α11 loss-of-function mutation, Thr54Met, causing familial hypocalciuric hypercalcemia type 2 (FHH2) J Bone Miner Res. 2016;31:1200–1206. doi: 10.1002/jbmr.2778. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Gorvin CM, Hannan FM, Cranston T, Valta H, Makitie O, Schalin-Jantti C, Thakker RV. Cinacalcet rectifies hypercalcemia in a patient with familial hypocalciuric hypercalcemia type 2 (FHH2) caused by a germline loss-of-function Gα11 mutation. J Bone Miner Res. 2018b;33:32–41. doi: 10.1002/jbmr.3241. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Gorvin CM, Hannan FM, Howles SA, Babinsky VN, Piret SE, Rogers A, Freidin AJ, Stewart M, Paudyal A, Hough TA, Nesbit MA, et al. Gα11 mutation in mice causes hypocalcemia rectifiable by calcilytic therapy. JCI Insight. 2017;2:e91103. doi: 10.1172/jci.insight.91103. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Gorvin CM, Rogers A, Hastoy B, Tarasov AI, Frost M, Sposini S, Inoue A, Whyte MP, Rorsman P, Hanyaloglu AC, Breitwieser GE, et al. AP2σ mutations impair calcium- sensing receptor trafficking and signaling, and show an endosomal pathway to spatially direct G-protein selectivity. Cell Rep. 2018c;22:1054–1066. doi: 10.1016/j.celrep.2017.12.089. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Gowen M, Stroup GB, Dodds RA, James IE, Votta BJ, Smith BR, Bhatnagar PK, Lago AM, Callahan JF, DelMar EG, Miller MA, et al. Antagonizing the parathyroid calcium receptor stimulates parathyroid hormone secretion and bone formation in osteopenic rats. J Clin Invest. 2000;105:1595–1604. doi: 10.1172/JCI9038. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Graca JA, Schepelmann M, Brennan SC, Reens J, Chang W, Yan P, Toka H, Riccardi D, Price SA. Comparative expression of the extracellular calcium-sensing receptor in the mouse, rat, and human kidney. Am J Physiol Renal Physiol. 2016;310:F518–533. doi: 10.1152/ajprenal.00208.2015. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Grant MP, Cavanaugh A, Breitwieser GE. 14-3-3 Proteins Buffer Intracellular Calcium Sensing Receptors to Constrain Signaling. PloS one. 2015;10:e0136702. doi: 10.1371/journal.pone.0136702. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Grant MP, Epure LM, Bokhari R, Roughley P, Antoniou J, Mwale F. Human cartilaginous endplate degeneration is induced by calcium and the extracellular calcium- sensing receptor in the intervertebral disc. Eur Cells Mater. 2016;32:137–151. doi: 10.22203/ecm.v032a09. [DOI] [PubMed] [Google Scholar]
- Grant MP, Stepanchick A, Cavanaugh A, Breitwieser GE. Agonist-driven maturation and plasma membrane insertion of calcium-sensing receptors dynamically control signal amplitude. Sci Signal. 2011;4 doi: 10.1126/scisignal.2002208. ra78. [DOI] [PubMed] [Google Scholar]
- Gray E, Muller D, Squires PE, Asare-Anane H, Huang GC, Amiel S, Persaud SJ, Jones PM. Activation of the extracellular calcium-sensing receptor initiates insulin secretion from human islets of Langerhans: involvement of protein kinases. J Endocrinol. 2006;190:703–710. doi: 10.1677/joe.1.06891. [DOI] [PubMed] [Google Scholar]
- Gregory KJ, Giraldo J, Daio J, Christopoulos A, Leach K. Evaluation of operational models of agonism and allosterism at receptors with multiple orthosteric binding sites. Mol Pharmacol. 2020;97:35–45. doi: 10.1124/mol.119.118091. [DOI] [PubMed] [Google Scholar]
- Gregory KJ, Kufareva I, Keller AN, Khajehali E, Mun HC, Goolam MA, Mason RS, Capuano B, Conigrave AD, Christopoulos A, Leach K. Dual action calcium-sensing receptor modulator unmasks novel mode-switching mechanism. Pharm Trans Sci. 2018;1:96–109. doi: 10.1021/acsptsci.8b00021. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Grevellec A, Graham A, Tucker AS. Shh signalling restricts the expression of Gcm2 and controls the position of the developing parathyroids. Dev Biol. 2011;353:194–205. doi: 10.1016/j.ydbio.2011.02.012. [DOI] [PubMed] [Google Scholar]
- Halse J, Greenspan S, Cosman F, Ellis G, Santora A, Leung A, Heyden N, Samanta S, Doleckyj S, Rosenberg E, Denker AE. A phase 2, randomized, placebo-controlled, dose- ranging study of the calcium-sensing receptor antagonist MK-5442 in the treatment of postmenopausal women with osteoporosis. J Clin Endocrinol Metab. 2014;99:E2207–2215. doi: 10.1210/jc.2013-4009. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Hamano N, Komaba H, Fukagawa M. Etelcalcetide for the treatment of secondary hyperparathyroidism. Expert Opin Pharmacother. 2017;18:529–534. doi: 10.1080/14656566.2017.1303482. [DOI] [PubMed] [Google Scholar]
- Handlogten ME, Huang CF, Shiraishi N, Awata H, Miller RT. The Ca2+-sensing receptor activates cytosolic phospholipase A2 via a Gqα-dependent ERK-independent pathway. J Biol Chem. 2001;276:13941–13948. doi: 10.1074/jbc.M007306200. [DOI] [PubMed] [Google Scholar]
- Handlogten ME, Shiraishi N, Awata H, Huang C, Miller RT. Extracellular Ca(2+)- sensing receptor is a promiscuous divalent cation sensor that responds to lead. Am J Physiol Renal Physiol. 2000;279:F1083–1091. doi: 10.1152/ajprenal.2000.279.6.F1083. [DOI] [PubMed] [Google Scholar]
- Hannan FM, Babinsky VN, Thakker RV. Disorders of the calcium-sensing receptor and partner proteins: insights into the molecular basis of calcium homeostasis. J Mol Endocrinol. 2016;57:R127–142. doi: 10.1530/JME-16-0124. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Hannan FM, Howles SA, Rogers A, Cranston T, Gorvin CM, Babinsky VN, Reed AA, Thakker CE, Bockenhauer D, Brown RS, Connell JM, et al. Adaptor protein-2 sigma subunit mutations causing familial hypocalciuric hypercalcaemia type 3 (FHH3) demonstrate genotype-phenotype correlations, codon bias and dominant-negative effects. Hum Mol Genet. 2015a;24:5079–5092. doi: 10.1093/hmg/ddv226. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Hannan FM, Kallay E, Chang W, Brandi ML, Thakker RV. The calcium-sensing receptor in physiology and in calcitropic and noncalcitropic diseases. Nat Rev Endocrinol. 2018a;15:33–51. doi: 10.1038/s41574-018-0115-0. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Hannan FM, Nesbit MA, Christie PT, Lissens W, Van der Schueren B, Bex M, Bouillon R, Thakker RV. A homozygous inactivating calcium-sensing receptor mutation, Pro339Thr, is associated with isolated primary hyperparathyroidism: correlation between location of mutations and severity of hypercalcaemia. Clin Endocrinol (Oxf) 2010;73:715–722. doi: 10.1111/j.1365-2265.2010.03870.x. [DOI] [PubMed] [Google Scholar]
- Hannan FM, Nesbit MA, Zhang C, Cranston T, Curley AJ, Harding B, Fratter C, Rust N, Christie PT, Turner JJ, Lemos MC, et al. Identification of 70 calcium-sensing receptor mutations in hyper- and hypo-calcaemic patients: evidence for clustering of extracellular domain mutations at calcium-binding sites. Hum Mol Genet. 2012;21:2768–2778. doi: 10.1093/hmg/dds105. [DOI] [PubMed] [Google Scholar]
- Hannan FM, Olesen MK, Thakker RV. Calcimimetic and calcilytic therapies for inherited disorders of the calcium-sensing receptor signalling pathway. Br J Pharmacol. 2018b;175:4083–4094. doi: 10.1111/bph.14086. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Hannan FM, Thakker RV. Calcium-sensing receptor (CaSR) mutations and disorders of calcium, electrolyte and water metabolism. Best Pract Res Clin Endocrinol Metab. 2013;27:359–371. doi: 10.1016/j.beem.2013.04.007. [DOI] [PubMed] [Google Scholar]
- Hannan FM, Walls GV, Babinsky VN, Nesbit MA, Kallay E, Hough TA, Fraser WD, Cox RD, Hu J, Spiegel AM, Thakker RV. The calcilytic agent NPS 2143 rectifies hypocalcemia in a mouse model with an activating calcium-sensing receptor (CaSR) mutation: Relevance to autosomal dominant hypocalcemia type 1 (ADH1) Endocrinology. 2015b;156:3114–3121. doi: 10.1210/en.2015-1269. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Hanyaloglu AC, von Zastrow M. Regulation of GPCRs by endocytic membrane trafficking and its potential implications. Annu Rev Pharmacol. 2008;48:537–568. doi: 10.1146/annurev.pharmtox.48.113006.094830. [DOI] [PubMed] [Google Scholar]
- Hauache OM, Hu J, Ray K, Spiegel AM. Functional interactions between the extracellular domain and the seven-transmembrane domain in Ca2+ receptor activation. Endocrine. 2000;13:63–70. doi: 10.1385/ENDO:13:1:63. [DOI] [PubMed] [Google Scholar]
- Hebert SC, Brown EM, Harris HW. Role of the Ca(2+)-sensing receptor in divalent mineral ion homeostasis. J Exp Biol. 1997;200:295–302. doi: 10.1242/jeb.200.2.295. [DOI] [PubMed] [Google Scholar]
- Hendy GN, Canaff L. Calcium-sensing receptor gene: Regulation of expression. Front Physiol. 2016;7:394. doi: 10.3389/fphys.2016.00394. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Hendy GN, Canaff L, Newfield RS, Tripto-Shkolnik L, Wong BY, Lee BS, Cole DE. Codon Arg15 mutations of the AP2S1 Gene: Common occurrence in familial hypocalciuric hypercalcemia cases negative for calcium-sensing receptor (CASR) mutations. J Clin Endocrinol Metab. 2014;99:E1311–1315. doi: 10.1210/jc.2014-1120. [DOI] [PubMed] [Google Scholar]
- Henley C, Yang Y, Davis J, Lu JY, Morony S, Fan W, Florio M, Sun B, Shatzen E, Pretorius JK, Richards WG, et al. Discovery of a calcimimetic with differential effects on parathyroid hormone and calcitonin secretion. J Pharmacol Exp Ther. 2011;337:681–691. doi: 10.1124/jpet.110.178681. [DOI] [PubMed] [Google Scholar]
- Hernandez-Bedolla MA, Carretero-Ortega J, Valadez-Sanchez M, Vazquez-Prado J, ReyesCruz G. Chemotactic and proangiogenic role of calcium sensing receptor is linked to secretion of multiple cytokines and growth factors in breast cancer MDA-MB-231 cells. Biochim Biophys Acta. 2015;1853:166–182. doi: 10.1016/j.bbamcr.2014.10.011. [DOI] [PubMed] [Google Scholar]
- Herrera-Vigenor F, Hernandez-Garcia R, Valadez-Sanchez M, Vazquez-Prado J, Reyes-Cruz G. AMSH regulates calcium-sensing receptor signaling through direct interactions. Biochem Biophys Res Commun. 2006;347:924–930. doi: 10.1016/j.bbrc.2006.06.169. [DOI] [PubMed] [Google Scholar]
- Hills CE, Younis MY, Bennett J, Siamantouras E, Liu KK, Squires PE. Calcium-sensing receptor activation increases cell-cell adhesion and beta-cell function. Cell Physiol Biochem. 2012;30:575–586. doi: 10.1159/000341439. [DOI] [PubMed] [Google Scholar]
- Ho C, Conner DA, Pollak MR, Ladd DJ, Kifor O, Warren HB, Brown EM, Seidman JG, Seidman CE. A mouse model of human familial hypocalciuric hypercalcemia and neonatal severe hyperparathyroidism. Nature genetics. 1995;11:389–394. doi: 10.1038/ng1295-389. [DOI] [PubMed] [Google Scholar]
- Hobson SA, Wright J, Lee F, McNeil SE, Bilderback T, Rodland KD. Activation of the MAP kinase cascade by exogenous calcium-sensing receptor. Mol Cell Endocrinol. 2003;200:189–198. doi: 10.1016/s0303-7207(01)00749-3. [DOI] [PubMed] [Google Scholar]
- Hodgkin MN, Hills CE, Squires PE. The calcium-sensing receptor and insulin secretion: a role outside systemic control 15 years on. J Endocrinol. 2008;199:1–4. doi: 10.1677/JOE-08-0261. [DOI] [PubMed] [Google Scholar]
- Hough TA, Bogani D, Cheeseman MT, Favor J, Nesbit MA, Thakker RV, Lyon MF. Activating calcium-sensing receptor mutation in the mouse is associated with cataracts and ectopic calcification. Proc Natl Acad Sci USA. 2004a;101:13566–13571. doi: 10.1073/pnas.0405516101. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Hough TA, Bogani D, Cheeseman MT, Favor J, Nesbit MA, Thakker RV, Lyon MF. Activating calcium-sensing receptor mutation in the mouse is associated with cataracts and ectopic calcification. Proc Natl Acad Sci USA. 2004b;101:13566–13571. doi: 10.1073/pnas.0405516101. [DOI] [PMC free article] [PubMed] [Google Scholar]
- House MG, Kohlmeier L, Chattopadhyay N, Kifor O, Yamaguchi T, Leboff MS, Glowacki J, Brown EM. Expression of an extracellular calcium-sensing receptor in human and mouse bone marrow cells. J Bone Miner Res. 1997;12:1959–1970. doi: 10.1359/jbmr.1997.12.12.1959. [DOI] [PubMed] [Google Scholar]
- Hovden S, Rejnmark L, Ladefoged SA, Nissen PH. AP2S1 and GNA11 mutations - not a common cause of familial hypocalciuric hypercalcemia. Eur J Endocrinol. 2017;176:177–185. doi: 10.1530/EJE-16-0842. [DOI] [PubMed] [Google Scholar]
- Howles SA, Hannan FM, Babinsky VN, Rogers A, Gorvin CM, Rust N, Nesbit MA, Thakker RV, Richardson T, McKenna MJ. Cinacalcet for symptomatic hypercalcemia caused by AP2S1 mutations. New Engl J Med. 2016;374:1396–1398. doi: 10.1056/NEJMc1511646. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Howles SA, Hannan FM, Gorvin CM, Piret SE, Paudyal A, Stewart M, Hough TA, Nesbit MA, Wells S, Brown SD, Cox RD, Thakker RV. Cinacalcet corrects hypercalcemia in mice with an inactivating Gα11 mutation. JCI Insight. 2017;2 doi: 10.1172/jci.insight.96540. pii:96540. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Hu J, Jiang J, Costanzi S, Thomas C, Yang W, Feyen JH, Jacobson KA, Spiegel AM. A missense mutation in the seven-transmembrane domain of the human Ca2+ receptor converts a negative allosteric modulator into a positive allosteric modulator. J Biol Chem. 2006;281:21558–21565. doi: 10.1074/jbc.M603682200. [DOI] [PubMed] [Google Scholar]
- Huang C, Miller RT. Novel Ca receptor signaling pathways for control of renal ion transport. Curr Opin Nephrol Hypertens. 2010;19:106–112. doi: 10.1097/MNH.0b013e328332e7b2. [DOI] [PubMed] [Google Scholar]
- Huang CF, Handlogten ME, Miller RT. Parallel activation of phosphatidylinositol 4- kinase and phospholipase C by the extracellular calcium-sensing receptor. J Biol Chem. 2002;277:20293–20300. doi: 10.1074/jbc.M200831200. [DOI] [PubMed] [Google Scholar]
- Huang CF, Hujer KM, Wu ZZ, Miller RT. The Ca2+-sensing receptor couples to Gα12/13 to activate phospholipase D in Madin-Darby canine kidney cells. Am J Physiol. 2004;286:C22–C30. doi: 10.1152/ajpcell.00229.2003. [DOI] [PubMed] [Google Scholar]
- Huang Y, Breitwieser GE. Rescue of calcium-sensing receptor mutants by allosteric modulators reveals a conformational checkpoint in receptor biogenesis. J Biol Chem. 2007;282:9517–9525. doi: 10.1074/jbc.M609045200. [DOI] [PubMed] [Google Scholar]
- Huang Y, Zhou Y, Castiblanco A, Yang W, Brown EM, Yang JJ. Multiple Ca(2+)- binding sites in the extracellular domain of the Ca(2+)-sensing receptor corresponding to cooperative Ca(2+) response. Biochemistry. 2009;48:388–398. doi: 10.1021/bi8014604. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Huang Y, Zhou Y, Yang W, Butters R, Lee HW, Li S, Castiblanco A, Brown EM, Yang JJ. Identification and dissection of Ca(2+)-binding sites in the extracellular domain of Ca(2+)-sensing receptor. J Biol Chem. 2007;282:19000–19010. doi: 10.1074/jbc.M701096200. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Hurtel-Lemaire AS, Mentaverri R, Caudrillier A, Cournarie F, Wattel A, Kamel S, Terwilliger EF, Brown EM, Brazier M. The calcium-sensing receptor is involved in strontium ranelate-induced osteoclast apoptosis. New insights into the associated signaling pathways. J Biol Chem. 2009;284:575–584. doi: 10.1074/jbc.M801668200. [DOI] [PubMed] [Google Scholar]
- Imura A, Tsuji Y, Murata M, Maeda R, Kubota K, Iwano A, Obuse C, Togashi K, Tominaga M, Kita N, Tomiyama K, et al. alpha-Klotho as a regulator of calcium homeostasis. Science. 2007;316:1615–1618. doi: 10.1126/science.1135901. [DOI] [PubMed] [Google Scholar]
- Jacobsen SE, Gether U, Bräuner-Osborne H. Investigating the molecular mechanism of positive and negative allosteric modulators in the calcium-sensing receptor dimer. Sci Rep. 2017;7 doi: 10.1038/srep46355. 46355. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Janicic N, Soliman E, Pausova Z, Seldin MF, Riviere M, Szpirer J, Szpirer C, Hendy GN. Mapping of the calcium-sensing receptor gene (CASR) to human chromosome 3q13.3-21 by fluorescence in situ hybridization, and localization to rat chromosome 11 and mouse chromosome 16. Mamm Genome. 1995;6:798–801. doi: 10.1007/BF00539007. [DOI] [PubMed] [Google Scholar]
- Jensen AA, Greenwood JR, Bräuner-Osborne H. The dance of the clams: twists and turns in the family C GPCR homodimer. Trends Pharmacol Sci. 2002;23:491–493. doi: 10.1016/s0165-6147(02)02107-7. [DOI] [PubMed] [Google Scholar]
- John MR, Harfst E, Loeffler J, Belleli R, Mason J, Bruin GJ, Seuwen K, Klickstein LB, Mindeholm L, Widler L, Kneissel M. AXT914 a novel, orally-active parathyroid hormone- releasing drug in two early studies of healthy volunteers and postmenopausal women. Bone. 2014;64:204–210. doi: 10.1016/j.bone.2014.04.015. [DOI] [PubMed] [Google Scholar]
- Jones BL, Smith SM. Calcium-sensing receptor: A key target for extracellular calcium signaling in neurons. Front Physiol. 2016;7:116. doi: 10.3389/fphys.2016.00116. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Josephs TM, Keller AN, Khajehali E, DeBono A, Langmead CJ, Conigrave AD, Capuano B, Kufareva I, Gregory KJ, Leach K. Negative allosteric modulators of the human calcium sensing receptor bind to topographically distinct sites within the 7 transmembrane domain. Br J Pharmacol. 2019 doi: 10.1111/bph.14961. In Press. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Jungman E, Pirot F, Maibach H. Ex vivo calcium percutaneous eggression in normal and tape-stripped human skin. Cutan Ocul Toxicol. 2012;31:1–6. doi: 10.3109/15569527.2011.590570. [DOI] [PubMed] [Google Scholar]
- Kallay E, Bonner E, Wrba F, Thakker RV, Peterlik M, Cross HS. Molecular and functional characterization of the extracellular calcium-sensing receptor in human colon cancer cells. Oncol Res. 2003;13:551–559. doi: 10.3727/000000003108748072. [DOI] [PubMed] [Google Scholar]
- Kallay E, Kifor O, Chattopadhyay N, Brown EM, Bischof MG, Peterlik M, Cross HS. Calcium-dependent c-myc proto-oncogene expression and proliferation of Caco-2 cells: a role for a luminal extracellular calcium-sensing receptor. Biochem Biophys Res Commun. 1997;232:80–83. doi: 10.1006/bbrc.1997.6225. [DOI] [PubMed] [Google Scholar]
- Kameda T, Mano H, Yamada Y, Takai H, Amizuka N, Kobori M, Izumi N, Kawashima H, Ozawa H, Ikeda K, Kameda A, Hakeda Y, Kumegawa M. Calcium-sensing receptor in mature osteoclasts, which are bone resorbing cells. Biochem Biophys Res Commun. 1998;245:419–422. doi: 10.1006/bbrc.1998.8448. [DOI] [PubMed] [Google Scholar]
- Kanatani M, Sugimoto T, Kanzawa M, Yano S, Chihara K. High extracellular calcium inhibits osteoclast-like cell formation by directly acting on the calcium-sensing receptor existing in osteoclast precursor cells. Biochem Biophys Res Commun. 1999;261:144–148. doi: 10.1006/bbrc.1999.0932. [DOI] [PubMed] [Google Scholar]
- Kantham L, Quinn S, Egbuna O, Baxi K, Butters R, Pang J, Pollak M, Goltzman D, Brown E. The calcium-sensing receptor (CaSR) defends against hypercalcemia independently of its regulation of parathyroid hormone secretion. Am J Physiol Endocrinol Metab. 2009;297:E915–E923. doi: 10.1152/ajpendo.00315.2009. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kapoor A, Satishchandra P, Ratnapriya R, Reddy R, Kadandale J, Shankar SK, Anand A. An idiopathic epilepsy syndrome linked to 3q13.3-q21 and missense mutations in the extracellular calcium sensing receptor gene. Ann Neurol. 2008;64:158–167. doi: 10.1002/ana.21428. [DOI] [PubMed] [Google Scholar]
- Katritch V, Cherezov V, Stevens RC. Structure-function of the G protein-coupled receptor superfamily. Annu Rev Pharmacol Toxicol. 2013;53:531–556. doi: 10.1146/annurev-pharmtox-032112-135923. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kawata T, Nagano N, Obi M, Miyata S, Koyama C, Kobayashi N, Wakita S, Wada M. Cinacalcet suppresses calcification of the aorta and heart in uremic rats. Kidney Int. 2008;74:1270–1277. doi: 10.1038/ki.2008.407. [DOI] [PubMed] [Google Scholar]
- Kawata T, Tokunaga S, Murai M, Masuda N, Haruyama W, Shoukei Y, Hisada Y, Yanagida T, Miyazaki H, Wada M, Akizawa T, Fukagawa M. A novel calcimimetic agent, evocalcet (MT-4580/KHK7580), suppresses the parathyroid cell function with little effect on the gastrointestinal tract or CYP isozymes in vivo and in vitro. PLoS One. 2018;13:e0195316. doi: 10.1371/journal.pone.0195316. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Keller AN, Kufareva I, Josephs TM, Diao J, Mai VT, Conigrave AD, Christopoulos A, Gregory KJ, Leach K. Identification of global and ligand-specific calcium sensing receptor activation mechanisms. Mol Pharmacol. 2018;93:619–630. doi: 10.1124/mol.118.112086. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kelly BT, McCoy AJ, Spate K, Miller SE, Evans PR, Honing S, Owen DJ. A structural explanation for the binding of endocytic dileucine motifs by the AP2 complex. Nature. 2008;456:976–979. doi: 10.1038/nature07422. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kelly JC, Lungchukiet P, Macleod RJ. Extracellular calcium-sensing receptor inhibition of intestinal epithelial TNF signaling requires CaSR-mediated Wnt5a/Ror2 interaction. Front Physiol. 2011;2:17. doi: 10.3389/fphys.2011.00017. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kenakin T, Christopoulos A. Signalling bias in new drug discovery: detection, quantification and therapeutic impact. Nat Rev Drug Discov. 2013;12:205–216. doi: 10.1038/nrd3954. [DOI] [PubMed] [Google Scholar]
- Kerstetter JE, Svastisalee CM, Caseria DM, Mitnick ME, Insogna KL. A threshold for low-protein-diet-induced elevations in parathyroid hormone. Am J Clin Nutr. 2000;72:168–173. doi: 10.1093/ajcn/72.1.168. [DOI] [PubMed] [Google Scholar]
- Kessler A, Faure H, Petrel C, Rognan D, Cesario M, Ruat M, Dauban P, Dodd RH. N1- Benzoyl-N2-[1-(1-naphthyl)ethyl]-trans-1,2-diaminocyclohexanes: Development of 4- chlorophenylcarboxamide (calhex 231) as a new calcium sensing receptor ligand demonstrating potent calcilytic activity. J Med Chem. 2006;49:5119–5128. doi: 10.1021/jm051233+. [DOI] [PubMed] [Google Scholar]
- Kessler A, Faure H, Roussanne MC, Ferry S, Ruat M, Dauban P, Dodd RH. N(1)- Arylsulfonyl-N(2)-(1-(1-naphthyl)ethyl)-1,2-diaminocyclohexanes: a new class of calcilytic agents acting at the calcium-sensing receptor. Chembiochem. 2004;5:1131–1136. doi: 10.1002/cbic.200400049. [DOI] [PubMed] [Google Scholar]
- Kifor O, Diaz R, Butters R, Brown EM. The Ca2+-sensing receptor (CaR) activates phospholipases C, A2, and D in bovine parathyroid and CaR-transfected, human embryonic kidney (HEK293) cells. J Bone Miner Res. 1997;12:715–725. doi: 10.1359/jbmr.1997.12.5.715. [DOI] [PubMed] [Google Scholar]
- Kifor O, Diaz R, Butters R, Kifor I, Brown EM. The calcium-sensing receptor is localized in caveolin-rich plasma membrane domains of bovine parathyroid cells. J Biol Chem. 1998;273:21708–21713. doi: 10.1074/jbc.273.34.21708. [DOI] [PubMed] [Google Scholar]
- Kifor O, Kifor I, Moore F, Butters R, Cantor T, Gao P, Brown E. Decreased expression of caveolin-1 and altered regulation of mitogen-activated protein kinase in cultured bovine parathyroid cells and human parathyroid adenomas. J Clin Endocrinol Metab. 2003;88:4455–4464. doi: 10.1210/jc.2002-021427. [DOI] [PubMed] [Google Scholar]
- Kifor O, MacLeod RJ, Diaz R, Bai M, Yamaguchi T, Yao T, I K, Brown EM. Regulation of MAP kinase by calcium-sensing receptor in bovine parathyroid and CaR-transfected HEK293 cells. Am J Physio. 2001;280:F291–F302. doi: 10.1152/ajprenal.2001.280.2.F291. [DOI] [PubMed] [Google Scholar]
- Kifor O, Moore FD, Wang P, Goldstein M, Vassilev P, Kifor I, Hebert SC, Brown EM. Reduced immunostaining for the extracellular Ca2+ sensing receptor in primary and uremic secondary hyperparathyroidism. J Clin Endocrinol Metab. 1996;81:1598–1606. doi: 10.1210/jcem.81.4.8636374. [DOI] [PubMed] [Google Scholar]
- Kim CH, Braud S, Isaac JT, Roche KW. Protein kinase C phosphorylation of the metabotropic glutamate receptor mGluR5 on Serine 839 regulates Ca2+ oscillations. J Biol Chem. 2005;280:25409–25415. doi: 10.1074/jbc.M502644200. [DOI] [PubMed] [Google Scholar]
- Kim JY, Ho H, Kim N, Liu J, Tu CL, Yenari MA, Chang W. Calcium-sensing receptor (CaSR) as a novel target for ischemic neuroprotection. Ann Clin Transl Neurol. 2014;1:851–866. doi: 10.1002/acn3.118. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kim JY, Kim N, Yenari MA, Chang W. Hypothermia and pharmacological regimens that prevent overexpression and overactivity of the extracellular calcium-sensing receptor protect neurons against traumatic brain injury. J Neurotrauma. 2013;30:1170–1176. doi: 10.1089/neu.2012.2691. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kim W, Takyar FM, Swan K, Jeong J, VanHouten J, Sullivan C, Dann P, Yu H, Fiaschi-Taesch N, Chang W, Wysolmerski J. Calcium-sensing receptor promotes breast cancer by stimulating intracrine actions of parathyroid hormone-related protein. Cancer Res. 2016;76:5348–5360. doi: 10.1158/0008-5472.CAN-15-2614. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kim W, Wysolmerski JJ. Calcium-sensing receptor in breast physiology and cancer. Front Physiol. 2016;7:440. doi: 10.3389/fphys.2016.00440. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kirberger M, Wang X, Deng H, Yang W, Chen G, Yang JJ. Statistical analysis of structural characteristics of protein Ca2+-binding sites. J Biol Inorg Chem. 2008;13:1169–1181. doi: 10.1007/s00775-008-0402-7. [DOI] [PubMed] [Google Scholar]
- Kitay AM, Schneebacher MT, Schmitt A, Heschl K, Kopic S, Alfadda T, Alsaihati A, Link A, Geibel JP. Modulations in extracellular calcium lead to H(+)-ATPase-dependent acid secretion: a clarification of PPI failure. Am J Physiol-Gastr L. 2018;315:G36–G42. doi: 10.1152/ajpgi.00132.2017. [DOI] [PubMed] [Google Scholar]
- Kitsou-Mylona I, Burns CJ, Squires PE, Persaud SJ, Jones PM. A role for the extracellular calcium-sensing receptor in cell-cell communication in pancreatic islets of langerhans. Cell Physiol Biochem. 2008;22:557–566. doi: 10.1159/000185540. [DOI] [PubMed] [Google Scholar]
- Koehl A, Hu H, Feng D, Sun B, Zhang Y, Robertson MJ, Chu M, Kobilka TS, Laermans T, Steyaert J, Tarrasch J, Dutta S, Fonseca R, Weis WI, Mathiesen JM, Skiniotis G, Kobilka BK. Structural insights into the activation of metabotropic glutamate receptors. Nature. 2019;566:79–84. doi: 10.1038/s41586-019-0881-4. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Koh J, Dar M, Untch BR, Dixit D, Shi Y, Yang Z, Adam MA, Dressman H, Wang X, Gesty-Palmer D, Marks JR, Spurney R, Druey KM, Olson JA., Jr Regulator of G protein signaling 5 is highly expressed in parathyroid tumors and inhibits signaling by the calcium- sensing receptor. Mol Endocrinol. 2011;25:867–876. doi: 10.1210/me.2010-0277. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Komuves L, Oda Y, Tu CL, Chang WH, Ho-Pao CL, Mauro T, Bikle DD. Epidermal expression of the full-length extracellular calcium-sensing receptor is required for normal keratinocyte differentiation. J Cell Physiol. 2002;192:45–54. doi: 10.1002/jcp.10107. [DOI] [PubMed] [Google Scholar]
- Kos CH, Karaplis AC, Peng JB, Hediger MA, Goltzman D, Mohammad KS, Guise TA, Pollak MR. The calcium-sensing receptor is required for normal calcium homeostasis independent of parathyroid hormone. J Clin Invest. 2003;111:1021–1028. doi: 10.1172/JCI17416. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Krissinel E, Henrick K. Inference of macromolecular assemblies from crystalline state. J Mol Biol. 2007;372:774–797. doi: 10.1016/j.jmb.2007.05.022. [DOI] [PubMed] [Google Scholar]
- Kruse AC, Ring AM, Manglik A, Hu J, Hu K, Eitel K, Hubner H, Pardon E, Valant C, Sexton PM, Christopoulos A, Felder CC, Gmeiner P, Steyaert J, Weis WI, Garcia KC, Wess J, Kobilka BK. Activation and allosteric modulation of a muscarinic acetylcholine receptor. Nature. 2013;504:101–106. doi: 10.1038/nature12735. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kumar S, Matheny CJ, Hoffman SJ, Marquis RW, Schultz M, Liang X, Vasko JA, Stroup GB, Vaden VR, Haley H, Fox J, DelMar EG, Nemeth EF, Lago AM, Callahan JF, Bhatnagar P, Huffman WF, Gowen M, Yi B, Danoff TM, Fitzpatrick LA. An orally active calcium-sensing receptor antagonist that transiently increases plasma concentrations of PTH and stimulates bone formation. Bone. 2010;46:534–542. doi: 10.1016/j.bone.2009.09.028. [DOI] [PubMed] [Google Scholar]
- Kunishima N, Shimada Y, Tsuji Y, Sato T, Yamamoto M, Kumasaka T, Nakanishi S, Jingami H, Morikawa K. Structural basis of glutamate recognition by a dimeric metabotropic glutamate receptor. Nature. 2000;407:971–977. doi: 10.1038/35039564. [DOI] [PubMed] [Google Scholar]
- Kurosawa M, Shimizu Y, Tsukagoshi H, Ueki M. Elevated levels of peripheral-blood, naturally occurring aliphatic polyamines in bronchial asthmatic patients with active symptoms. Allergy. 1992;47:638–643. doi: 10.1111/j.1398-9995.1992.tb02388.x. [DOI] [PubMed] [Google Scholar]
- Ladds G, Goddard A, Hill C, Thornton S, Davey J. Differential effects of RGS proteins on G alpha(q) and G alpha(11) activity. Cell Signal. 2007;19:103–113. doi: 10.1016/j.cellsig.2006.05.027. [DOI] [PubMed] [Google Scholar]
- Lagerström MC, Schiöth HB. Structural diversity of G protein-coupled receptors and significance for drug discovery. Nat Rev Drug Discov. 2008;7:339–357. doi: 10.1038/nrd2518. [DOI] [PubMed] [Google Scholar]
- Lazarus S, Pretorius CJ, Khafagi F, Campion KL, Brennan SC, Conigrave AD, Brown EM, Ward DT. A novel mutation of the primary protein kinase C phosphorylation site in the calcium-sensing receptor causes autosomal dominant hypocalcemia. Eur J Endocrinol. 2011;164:429–435. doi: 10.1530/EJE-10-0907. [DOI] [PubMed] [Google Scholar]
- Leach K, Gregory KJ. Molecular insights into allosteric modulation of Class C G protein- coupled receptors. Pharmacol Res. 2017;116:105–118. doi: 10.1016/j.phrs.2016.12.006. [DOI] [PubMed] [Google Scholar]
- Leach K, Gregory KJ, Kufareva I, Khajehali E, Cook AE, Abagyan R, Conigrave AD, Sexton PM, Christopoulos A. Towards a structural understanding of allosteric drugs at the human calcium-sensing receptor. Cell Res. 2016;26:574–592. doi: 10.1038/cr.2016.36. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Leach K, Sexton PM, Christopoulos A. Allosteric GPCR modulators: taking advantage of permissive receptor pharmacology. Trends Pharmacol Sci. 2007;28:382–389. doi: 10.1016/j.tips.2007.06.004. [DOI] [PubMed] [Google Scholar]
- Leach K, Sexton PM, Christopoulos A, Conigrave AD. Engendering biased signalling from the calcium-sensing receptor for the pharmacotherapy of diverse disorders. Br J Pharmacol. 2014;171:1142–1155. doi: 10.1111/bph.12420. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Leach K, Wen A, Cook AE, Sexton PM, Conigrave AD, Christopoulos A. Impact of clinically relevant mutations on the pharmacoregulation and signaling bias of the calcium- sensing receptor by positive and negative allosteric modulators. Endocrinology. 2013;154:1105–1116. doi: 10.1210/en.2012-1887. [DOI] [PubMed] [Google Scholar]
- Leach K, Wen A, Davey AE, Sexton PM, Conigrave AD, Christopoulos A. Identification of molecular phenotypes and biased signaling induced by naturally occurring mutations of the human calcium-sensing receptor. Endocrinology. 2012;153:4304–4316. doi: 10.1210/en.2012-1449. [DOI] [PubMed] [Google Scholar]
- LeBoff MS, Shoback D, Brown EM, Thatcher J, Leombruno R, Beaudoin D, Henry M, Wilson R, Pallotta J, Marynick S, Stock J, Leight G. Regulation of parathyroid hormone release and cytosolic calcium by extracellular calcium in dispersed and cultured bovine and pathological human parathyroid cells. J Clin Invest. 1985;75:49–57. doi: 10.1172/JCI111696. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Lee H, Mun H-C, Lewis NC, Crouch MF, Culverston EL, Mason RS, Conigrave AD. Allosteric activation of the extracellular Ca2+-sensing receptor by L-amino acids enhances ERK1/2 phosphorylation. Biochem J. 2007;404:141–149. doi: 10.1042/BJ20061826. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Lee JJ, Liu X, O'Neill D, Beggs MR, Weissgerber P, Flockerzi V, Chen XZ, Dimke H, Alexander RT. Activation of the calcium sensing receptor attenuates TRPV6- dependent intestinal calcium absorption. JCI Insight. 2019;4:e128013. doi: 10.1172/jci.insight.128013. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Letz S, Haag C, Schulze E, Frank-Raue K, Raue F, Hofner B, Mayr B, Schofl C. Amino alcohol- (NPS-2143) and quinazolinone-derived calcilytics (ATF936 and AXT914) differentially mitigate excessive signalling of calcium-sensing receptor mutants causing Bartter syndrome Type 5 and autosomal dominant hypocalcemia. PLoS One. 2014;9:e115178. doi: 10.1371/journal.pone.0115178. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Letz S, Rus R, Haag C, Dorr HG, Schnabel D, Mohlig M, Schulze E, Frank-Raue K, Raue F, Mayr B, Schofl C. Novel activating mutations of the calcium-sensing receptor: the calcilytic NPS-2143 mitigates excessive signal transduction of mutant receptors. J Clin Endocrinol Metab. 2010;95:E229–233. doi: 10.1210/jc.2010-0651. [DOI] [PubMed] [Google Scholar]
- Li D, Opas EE, Tuluc F, Metzger DL, Hou C, Hakonarson H, Levine MA. Autosomal dominant hypoparathyroidism caused by germline mutation in GNA11: phenotypic and molecular characterization. J Clin Endocrinol Metab. 2014a;99:E1774–1783. doi: 10.1210/jc.2014-1029. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Li X, Kong X, Jiang L, Ma T, Yan S, Yuan C, Yang Q. A genetic polymorphism (rs17251221) in the calcium-sensing receptor is associated with breast cancer susceptibility and prognosis. Cell Physiol Biochem. 2014b;33:165–172. doi: 10.1159/000356659. [DOI] [PubMed] [Google Scholar]
- Lietman SA, Tenenbaum-Rakover Y, Jap TS, Yi-Chi W, De-Ming Y, Ding C, Kussiny N, Levine MA. A novel loss-of-function mutation, Gln459Arg, of the calcium-sensing receptor gene associated with apparent autosomal recessive inheritance of familial hypocalciuric hypercalcemia. J Clin Endocrinol Metab. 2009;94:4372–4379. doi: 10.1210/jc.2008-2484. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Liou AP, Sei Y, Zhao X, Feng J, Lu X, Thomas C, Pechhold S, Raybould HE, Wank SA. The extracellular calcium-sensing receptor is required for cholecystokinin secretion in response to L-phenylalanine in acutely isolated intestinal I cells. Am J Physiol-Gastr L. 2011;300:G538–546. doi: 10.1152/ajpgi.00342.2010. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Liu G, Cao W, Jia G, Zhao H, Chen X, Wang J. Calcium-sensing receptor in nutrient sensing: an insight into the modulation of intestinal homoeostasis. Br J Nutr. 2018;120:881–890. doi: 10.1017/S0007114518002088. [DOI] [PubMed] [Google Scholar]
- Liu G, Hu X, Chakrabarty S. Calcium sensing receptor down-regulates malignant cell behavior and promotes chemosensitivity in human breast cancer cells. Cell Calcium. 2009;45:216–225. doi: 10.1016/j.ceca.2008.10.004. [DOI] [PubMed] [Google Scholar]
- Liu XL, Lu YS, Gao JY, Marshall C, Xiao M, Miao DS, Karaplis A, Goltzman D, Ding J. Calcium sensing receptor absence delays postnatal brain development via direct and indirect mechanisms. Mol Neurobiol. 2013;48:590–600. doi: 10.1007/s12035-013-8448-0. [DOI] [PubMed] [Google Scholar]
- Locatelli F, Cannata-Andia JB, Drueke TB, Horl WH, Fouque D, Heimburger O, Ritz E. Management of disturbances of calcium and phosphate metabolism in chronic renal insufficiency, with emphasis on the control of hyperphosphataemia. Nephrol Dial Transplant. 2002;17:723–731. doi: 10.1093/ndt/17.5.723. [DOI] [PubMed] [Google Scholar]
- Lorenz S, Frenzel R, Paschke R, Breitwieser GE, Miedlich SU. Functional desensitization of the extracellular calcium-sensing receptor is regulated via distinct mechanisms: Role of G protein-coupled receptor kinases, protein kinase C and β-arrestins. Endocrinology. 2007;148:2398–2404. doi: 10.1210/en.2006-1035. [DOI] [PubMed] [Google Scholar]
- Loupy A, Ramakrishnan S, Wootla B, Chambrey R, de la Faille R, Bourgeois S, Bruneval P, Mandet C, Christensen E, Faure H, Cheval L, Laghmani K, Collet C, Eladari D, Dodd R, Ruat M, Houillier P. PTH-independent regulation of blood calcium concentration by the calcium-sensing receptor. J Clin Invest. 2012;122:3355–3367. doi: 10.1172/JCI57407. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Lozano-Ortega G, Waser N, Bensink ME, Goring S, Bennett H, Block GA, Chertow GM, Trotman ML, Cooper K, Levy AR, Belozeroff V. Effects of calcimimetics on long-term outcomes in dialysis patients: literature review and Bayesian meta-analysis. J Comp Eff Res. 2018;7:693–707. doi: 10.2217/cer-2018-0015. [DOI] [PubMed] [Google Scholar]
- Lu JY, Yang Y, Gnacadja G, Christopoulos A, Reagan JD. Effect of the calcimimetic R- 568 [3-(2-chlorophenyl)-N-((1R)-1-(3-methoxyphenyl)ethyl)-1-propanamine] on correcting inactivating mutations in the human calcium-sensing receptor. J Pharmacol Exp Ther. 2009;331:775–786. doi: 10.1124/jpet.109.159228. [DOI] [PubMed] [Google Scholar]
- Ma JN, Owens M, Gustafsson M, Jensen J, Tabatabaei A, Schmelzer K, Olsson R, Burstein ES. Characterization of highly efficacious allosteric agonists of the human calcium- sensing receptor. J Pharmacol Exp Ther. 2011;337:275–284. doi: 10.1124/jpet.110.178194. [DOI] [PubMed] [Google Scholar]
- MacLeod RJ, Chattopadhyay N, Brown EM. PTHrP stimulated by the calcium-sensing receptor requires MAP kinase activation. Am J Physiol. 2003;284:E435–E442. doi: 10.1152/ajpendo.00143.2002. [DOI] [PubMed] [Google Scholar]
- MacLeod RJ, Yano S, Chattopadhyay N, Brown EM. Extracellular calcium-sensing receptor transactivates the epidermal growth factor receptor by a triple-membrane-spanning signaling mechanism. Biochem Biophys Res Comm. 2004;320:455–460. doi: 10.1016/j.bbrc.2004.05.198. [DOI] [PubMed] [Google Scholar]
- Makita N, Ando T, Sato J, Manaka K, Mitani K, Kikuchi Y, Niwa T, Ootaki M, Takeba Y, Matsumoto N, Kawakami A, Ogawa T, Nangaku M, Iiri T. Cinacalcet corrects biased allosteric modulation of CaSR by AHH autoantibody. JCI Insight. 2019;4 doi: 10.1172/jci.insight.126449. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Makita N, Iiri T. Biased agonism: a novel paradigm in G protein-coupled receptor signaling observed in acquired hypocalciuric hypercalcemia. Endocr J. 2014;61:303–309. doi: 10.1507/endocrj.ej13-0453. [DOI] [PubMed] [Google Scholar]
- Makita N, Sato J, Manaka K, Shoji Y, Oishi A, Hashimoto M, Fujita T, Iiri T. An acquired hypocalciuric hypercalcemia autoantibody induces allosteric transition among active human Ca-sensing receptor conformations. Proc Natl Acad Sci USA. 2007;104:5443–5448. doi: 10.1073/pnas.0701290104. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Mamillapalli R, VanHouten J, Zawalich W, Wysolmerski J. Switching of G-protein usage by the calcium-sensing receptor reverses its effect on parathyroid hormone-related protein secretion in normal versus malignant breast cells. J Biol Chem. 2008;283:24435–24447. doi: 10.1074/jbc.M801738200. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Mamillapalli R, Wysolmerski J. The calcium-sensing receptor couples to Gαs and regulates PTHrP and ACTH secretion in pituitary cells. J Endocrinol. 2010;204:287–297. doi: 10.1677/JOE-09-0183. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Mannstadt M, Harris M, Bravenboer B, Chitturi S, Dreijerink KM, Lambright DG, Lim ET, Daly MJ, Gabriel S, Juppner H. Germline mutations affecting Gα11 in hypoparathyroidism. N Engl J Med. 2013;368:2532–2534. doi: 10.1056/NEJMc1300278. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Margolis KG, Stevanovic K, Karamooz N, Li ZS, Ahuja A, D’Autreaux F, Saurman V, Chalazonitis A, Gershon MD. Enteric neuronal density contributes to the severity of intestinal inflammation. Gastroenterology. 2011;141:588–598. doi: 10.1053/j.gastro.2011.04.047. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Marie PJ. The calcium-sensing receptor in bone cells: a potential therapeutic target in osteoporosis. Bone. 2010;46:571–576. doi: 10.1016/j.bone.2009.07.082. [DOI] [PubMed] [Google Scholar]
- Masvidal L, Iniesta R, Casala C, Galvan P, Rodriguez E, Lavarino C, Mora J, de Torres C. Polymorphisms in the calcium-sensing receptor gene are associated with clinical outcome of neuroblastoma. PLoS One. 2013;8:e59762. doi: 10.1371/journal.pone.0059762. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Masvidal L, Iniesta R, Garcia M, Casala C, Lavarino C, Mora J, de Torres C. Genetic variants in the promoter region of the calcium-sensing receptor gene are associated with its down-regulation in neuroblastic tumors. Mol Carcinog. 2017;56:1281–1289. doi: 10.1002/mc.22589. [DOI] [PubMed] [Google Scholar]
- McCormick WD, Atkinson-Dell R, Campion KL, Mun HC, Conigrave AD, Ward DT. Increased receptor stimulation elicits differential calcium-sensing receptor (T888) dephosphorylation. J Biol Chem. 2010;285:14170–14177. doi: 10.1074/jbc.M109.071084. [DOI] [PMC free article] [PubMed] [Google Scholar]
- McGehee DS, Aldersberg M, Liu KP, Hsuing S, Heath MJ, Tamir H. Mechanism of extracellular Ca2+ receptor-stimulated hormone release from sheep thyroid parafollicular cells. J Physiol (Lond) 1997;502:31–44. doi: 10.1111/j.1469-7793.1997.031bl.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
- McLarnon S, Holden D, Ward D, Jones M, Elliott A, Riccardi D. Aminoglycoside antibiotics induce pH-sensitive activation of the calcium-sensing receptor. Biochem Biophys Res Commun. 2002;297:71–77. doi: 10.1016/s0006-291x(02)02133-2. [DOI] [PubMed] [Google Scholar]
- McLarnon SJ, Riccardi D. Physiological and pharmacological agonists of the extracellular Ca2+-sensing receptor. Eur J Pharmacol. 2002;447:271–278. doi: 10.1016/s0014-2999(02)01849-6. [DOI] [PubMed] [Google Scholar]
- McMurtry CT, Schranck FW, Walkenhorst DA, Murphy WA, Kocher DB, Teitelbaum SL, Rupich RC, Whyte MP. Significant developmental elevation in serum parathyroid hormone levels in a large kindred with familial benign (hypocalciuric) hypercalcemia. Am J Med. 1992;93:247–258. doi: 10.1016/0002-9343(92)90229-5. [DOI] [PubMed] [Google Scholar]
- Meng K, Xu J, Zhang C, Zhang R, Yang H, Liao C, Jiao J. Calcium sensing receptor modulates extracellular calcium entry and proliferation via TRPC3/6 channels in cultured human mesangial cells. PLoS One. 2014;9:e98777. doi: 10.1371/journal.pone.0098777. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Menon GK, Elias PM. Ultrastructural localization of calcium in psoriatic and normal human epidermis. Arch Dermatol. 1991;127:57–63. [PubMed] [Google Scholar]
- Menon GK, Elias PM, Feingold KR. Integrity of the permeability barrier is crucial for maintenance of the epidermal calcium gradient. Br J Dermatol. 1994;130:139–147. doi: 10.1111/j.1365-2133.1994.tb02892.x. [DOI] [PubMed] [Google Scholar]
- Menon GK, Grayson S, Elias PM. Ionic calcium reservoirs in mammalian epidermis: ultrastructural localization by ion-capture cytochemistry. J Invest Dermatol. 1985;84:508–512. doi: 10.1111/1523-1747.ep12273485. [DOI] [PubMed] [Google Scholar]
- Mentaverri R, Yano S, Chattopadhyay N, Petit L, Kifor O, Kamel S, Terwilliger EF, Brazier M, Brown EM. The calcium sensing receptor is directly involved in both osteoclast differentiation and apoptosis. FASEB J. 2006;20:2562–2564. doi: 10.1096/fj.06-6304fje. [DOI] [PubMed] [Google Scholar]
- Miedlich SU, Gama L, Seuwen K, Wolf RM, Breitwieser GE. Homology modeling of the transmembrane domain of the human calcium sensing receptor and localization of an allosteric binding site. J Biol Chem. 2004;279:7254–7263. doi: 10.1074/jbc.M307191200. [DOI] [PubMed] [Google Scholar]
- Miller RT. Control of renal calcium, phosphate, electrolyte, and water excretion by the calcium-sensing receptor. Best Pract Res Clin Endocrinol Metab. 2013;27:345–358. doi: 10.1016/j.beem.2013.04.009. [DOI] [PubMed] [Google Scholar]
- Miller-Fleming L, Olin-Sandoval V, Campbell K, Ralser M. Remaining mysteries of molecular biology: The role of polyamines in the cell. J Mol Biol. 2015;427:3389–3406. doi: 10.1016/j.jmb.2015.06.020. [DOI] [PubMed] [Google Scholar]
- Mine Y, Zhang H. Anti-inflammatory effects of poly-L-lysine in intestinal mucosal system mediated by calcium-sensing receptor activation. J Agric Food Chem. 2015;63:10437–10447. doi: 10.1021/acs.jafc.5b03812. [DOI] [PubMed] [Google Scholar]
- Mizobuchi M, Ritter CS, Krits I, Slatopolsky E, Sicard G, Brown AJ. Calcium-sensing receptor expression is regulated by glial cells missing-2 in human parathyroid cells. J Bone Miner Res. 2009;24:1173–1179. doi: 10.1359/JBMR.090211. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Mos I, Jacobsen SE, Foster SR, Brauner-Osborne H. Calcium-Sensing Receptor Internalization Is beta-Arrestin-Dependent and Modulated by Allosteric Ligands. Mol Pharmacol. 2019a;96:463–474. doi: 10.1124/mol.119.116772. [DOI] [PubMed] [Google Scholar]
- Mos I, Jacobsen SE, Foster SR, Brauner-Osborne H. Calcium-sensing receptor internalization is β-arrestin-dependent and modulated by allosteric ligands. Mol Pharmacol. 2019b;96:463–474. doi: 10.1124/mol.119.116772. [DOI] [PubMed] [Google Scholar]
- Muff R, Nemeth EF, Haller-Brem S, Fischer JA. Regulation of hormone secretion and cytosolic Ca2+ by extracellular Ca2+ in parathyroid cells and C-cells: role of voltage- sensitive Ca2+ channels. Arch Biochem Biophys. 1988;265:128–135. doi: 10.1016/0003-9861(88)90378-5. [DOI] [PubMed] [Google Scholar]
- Mun HC, Culverston EL, Franks AH, Collyer CA, Clifton-Bligh RJ, Conigrave AD. A double mutation in the extracellular Ca2+-sensing receptor’s venus flytrap domain that selectively disables L-amino acid sensing. J Biol Chem. 2005;280:29067–29072. doi: 10.1074/jbc.M500002200. [DOI] [PubMed] [Google Scholar]
- Mun HC, Franks AH, Culverston EL, Krapcho K, Nemeth EF, Conigrave AD. The Venus Fly Trap domain of the extracellular Ca2+ -sensing receptor is required for L-amino acid sensing. J Biol Chem. 2004;279:51739–51744. doi: 10.1074/jbc.M406164/200. [DOI] [PubMed] [Google Scholar]
- Mun HC, Leach KM, Conigrave AD. L-amino acids promote calcitonin release via a calcium-sensing receptor: Gq/11-mediated pathway in human C-cells. Endocrinology. 2019;160:1590–1599. doi: 10.1210/en.2018-00860. [DOI] [PubMed] [Google Scholar]
- Munk C, Isberg V, Mordalski S, Harpsoe K, Rataj K, Hauser AS, Kolb P, Bojarski AJ, Vriend G, Gloriam DE. GPCRdb: the G protein-coupled receptor database - an introduction. Br J Pharmacol. 2016;173:2195–2207. doi: 10.1111/bph.13509. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Murer H, Hernando N, Forster I, Biber J. Molecular aspects in the regulation of renal inorganic phosphate reabsorption: the type IIa sodium/inorganic phosphate co-transporter as the key player. Curr Opin Nephrol Hypertens. 2001;10:555–561. doi: 10.1097/00041552-200109000-00002. [DOI] [PubMed] [Google Scholar]
- Muresan Z, MacGregor R. The release of parathyroid hormone and the exocytosis of a proteoglycan are modulated by extracellular Ca2+ in a similar manner. Mol Biol Cell. 1994;5:725–737. doi: 10.1091/mbc.5.7.725. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Muto T, Tsuchiya D, Morikawa K, Jingami H. Structures of the extracellular regions of the group II/III metabotropic glutamate receptors. Proc Natl Acad Sci USA. 2007;104:3759–3764. doi: 10.1073/pnas.0611577104. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Nakagawa K, Parekh N, Koleganova N, Ritz E, Schaefer F, Schmitt CP. Acute cardiovascular effects of the calcimimetic R-568 and its enantiomer S-568 in rats. Pediatr Nephrol. 2009;24:1385–1389. doi: 10.1007/s00467-009-1153-6. [DOI] [PubMed] [Google Scholar]
- Nakamura H, Tsujiguchi H, Hara A, Kambayashi Y, Miyagi S, Thu Nguyen TT, Suzuki K, Tao Y, Sakamoto Y, Shimizu Y, Yamamoto N, et al. Dietary calcium intake and hypertension: importance of serum concentrations of 25-hydroxyvitamin D. Nutrients. 2019;11 doi: 10.3390/nu11040911. pii:E911. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Nearing J, Betka M, Quinn S, Hentschel H, Elger M, Baum M, Bai M, Chattopadyhay N, Brown EM, Hebert SC, Harris HW. Polyvalent cation receptor proteins (CaRs) are salinity sensors in fish. Proc Natl Acad Sci USA. 2002;99:9231–9236. doi: 10.1073/pnas.152294399. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Nemeth EF, Scarpa A. Cytosolic Ca2+ and the regulation of secretion in parathyroid cells. FEBS Lett. 1986;203:15–19. doi: 10.1016/0014-5793(86)81427-2. [DOI] [PubMed] [Google Scholar]
- Nemeth EF, Scarpa A. Rapid mobilization of cellular Ca2+ in bovine parathyroid cells evoked by extracellular divalent cations. Evidence for a cell surface calcium receptor. J Biol Chem. 1987;262:5188–5196. [PubMed] [Google Scholar]
- Nemeth EF, Van Wagenen BC, Balandrin MF. Discovery and development of calcimimetic and calcilytic compounds. Prog Med Chem. 2018;57:1–86. doi: 10.1016/bs.pmch.2017.12.001. [DOI] [PubMed] [Google Scholar]
- Nemeth EF, Wallace J, Scarpa A. Stimulus-secretion coupling in bovine parathyroid cells. Dissociation between secretion and net changes in cytosolic Ca2+ . J Biol Chem. 1986;261:2668–2674. [PubMed] [Google Scholar]
- Nesbit MA, Hannan FM, Howles SA, Babinsky VN, Head RA, Cranston T, Rust N, Hobbs MR, Heath H, Thakker RV. Mutations affecting G-protein subunit α11 in hypercalcemia and hypocalcemia. N Engl J Med. 2013a;368:2476–2486. doi: 10.1056/NEJMoa1300253. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Nesbit MA, Hannan FM, Howles SA, Reed AA, Cranston T, Thakker CE, Gregory L, Rimmer AJ, Rust N, Graham U, Morrison PJ, Hunter SJ, Whyte MP, McVean G, Buck D, Thakker RV. Mutations in AP2S1 cause familial hypocalciuric hypercalcemia type 3. Nat Genet. 2013b;45:93–97. doi: 10.1038/ng.2492. [DOI] [PMC free article] [PubMed] [Google Scholar]
- North ML, Grasemann H, Khanna N, Inman MD, Gauvreau GM, Scott JA. Increased ornithine-derived polyamines cause airway hyperresponsiveness in a mouse model of asthma. Am J Respir Cell Mol Biol. 2013;48:694–702. doi: 10.1165/rcmb.2012-0323OC. [DOI] [PubMed] [Google Scholar]
- O'Hara PJ, Sheppard PO, Thogersen H, Venezia D, Haldeman BA, McGrane V, Houamed KM, Thomsen C, Gilbert TL, Mulvihill ER. The ligand-binding domain in metabotropic glutamate receptors is related to bacterial periplasmic binding proteins. Neuron. 1993;11:41–52. doi: 10.1016/0896-6273(93)90269-w. [DOI] [PubMed] [Google Scholar]
- Oda Y, Hu L, Nguyen T, Fong C, Tu CL, Bikle DD. Combined deletion of the vitamin D receptor and calcium-sensing receptor delays wound re-epithelialization. Endocrinology. 2017;158:1929–1938. doi: 10.1210/en.2017-00061. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Oda Y, Tu CL, Pillai S, Bikle DD. The calcium sensing receptor and its alternatively spliced form in keratinocyte differentiation. J Biol Chem. 1998;273:23344–23352. doi: 10.1074/jbc.273.36.23344. [DOI] [PubMed] [Google Scholar]
- Odenwald T, Nakagawa K, Hadtstein C, Roesch F, Gohlke P, Ritz E, Schaefer F, Schmitt CP. Acute blood pressure effects and chronic hypotensive action of calcimimetics in uremic rats. J Am Soc Nephrol. 2006;17:655–662. doi: 10.1681/ASN.2005090914. [DOI] [PubMed] [Google Scholar]
- Ogata H, Ritz E, Odoni G, Amann K, Orth SR. Beneficial effects of calcimimetics on progression of renal failure and cardiovascular risk factors. J Am Soc Nephrol. 2003;14:959–967. doi: 10.1097/01.asn.0000056188.23717.e5. [DOI] [PubMed] [Google Scholar]
- Oh YS, Seo EH, Lee YS, Cho SC, Jung HS, Park SC, Jun HS. Increase of calcium sensing receptor expression is related to compensatory insulin secretion during aging in mice. PLoS One. 2016;11:e0159689. doi: 10.1371/journal.pone.0159689. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Ohsu T, Amino Y, Nagasaki H, Yamanaka T, Takeshita S, Hatanaka T, Maruyama Y, Miyamura N, Eto Y. Involvement of the calcium-sensing receptor in human taste perception. J Biol Chem. 2010;285:1016–1022. doi: 10.1074/jbc.M109.029165. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Owen JL, Cheng SX, Ge Y, Sahay B, Mohamadzadeh M. The role of the calcium-sensing receptor in gastrointestinal inflammation. Semin Cell Dev Biol. 2016;49:44–51. doi: 10.1016/j.semcdb.2015.10.040. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Pais R, Gribble FM, Reimann F. Signalling pathways involved in the detection of peptones by murine small intestinal enteroendocrine L-cells. Peptides. 2016;77:9–15. doi: 10.1016/j.peptides.2015.07.019. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Pallais JC, Kemp EH, Bergwitz C, Kantham L, Slovik DM, Weetman AP, Brown EM. Autoimmune hypocalciuric hypercalcemia unresponsive to glucocorticoid therapy in a patient with blocking autoantibodies against the calcium-sensing receptor. J Clin Endocrinol Metab. 2011;96:672–680. doi: 10.1210/jc.2010-1739. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Pearce SH, Trump D, Wooding C, Besser GM, Chew SL, Grant DB, Heath DA, Hughes IA, Paterson CR, Whyte MP, Thakker RV. Calcium-sensing receptor mutations in familial benign hypercalcemia and neonatal hyperparathyroidism. J Clin Invest. 1995;96:2683–2692. doi: 10.1172/JCI118335. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Pearce SH, Williamson C, Kifor O, Bai M, Coulthard MG, Davies M, Lewis-Barned N, McCredie D, Powell H, Kendall-Taylor P, Brown EM, Thakker RV. A familial syndrome of hypocalcemia with hypercalciuria due to mutations in the calcium-sensing receptor. N Engl J Med. 1996;335:1115–1122. doi: 10.1056/NEJM199610103351505. [DOI] [PubMed] [Google Scholar]
- Peiris D, Pacheco I, Spencer C, MacLeod RJ. The extracellular calcium-sensing receptor reciprocally regulates the secretion of BMP-2 and the BMP antagonist Noggin in colonic myofibroblasts. Am J Physiol Gastrointest Liver Physiol. 2007;292:G753–766. doi: 10.1152/ajpgi.00225.2006. [DOI] [PubMed] [Google Scholar]
- Perez-Benito L, Doornbos MLJ, Cordomi A, Peeters L, Lavreysen H, Pardo L, Tresadern G. Molecular switches of allosteric modulation of the metabotropic glutamate 2 receptor. Structure. 2017;25:1153–1162 e1154. doi: 10.1016/j.str.2017.05.021. [DOI] [PubMed] [Google Scholar]
- Petrel C, Kessler A, Dauban P, Dodd RH, Rognan D, Ruat M. Positive and negative allosteric modulators of the Ca2+-sensing receptor interact within overlapping but not identical binding sites in the transmembrane domain. J Biol Chem. 2004;279:18990–18997. doi: 10.1074/jbc.M400724200. [DOI] [PubMed] [Google Scholar]
- Petrel C, Kessler A, Maslah F, Dauban P, Dodd RH, Rognan D, Ruat M. Modeling and mutagenesis of the binding site of Calhex 231, a novel negative allosteric modulator of the extracellular Ca(2+)-sensing receptor. J Biol Chem. 2003;278:49487–49494. doi: 10.1074/jbc.M308010200. [DOI] [PubMed] [Google Scholar]
- Pi M, Faber P, Ekema G, Jackson PD, Ting A, Wang N, Fontilla-Poole M, Mays RW, Brunden KR, Harrington JJ, Quarles LD. Identification of a novel extracellular cation-sensing G-protein-coupled receptor. J Biol Chem. 2005a;280:40201–40209. doi: 10.1074/jbc.M505186200. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Pi M, Oakley RH, Gesty-Palmer D, Cruickshank RD, Spurney RF, Luttrell LM, Quarles LD. β-arrestin- and G protein receptor kinase-mediated calcium-sensing receptor desensitization. Mol Endocrinol. 2005b;19:1078–1087. doi: 10.1210/me.2004-0450. [DOI] [PubMed] [Google Scholar]
- Pi M, Spurney RF, Tu QS, Hinson T, Quarles LD. Calcium-sensing receptor activation of Rho involves filamin and Rho-guanine nucleotide exchange factor. Endocrinology. 2002;143:3830–3838. doi: 10.1210/en.2002-220240. [DOI] [PubMed] [Google Scholar]
- Pidasheva S, Grant M, Canaff L, Ercan O, Kumar U, Hendy GN. Calcium-sensing receptor dimerizes in the endoplasmic reticulum: biochemical and biophysical characterization of CASR mutants retained intracellularly. Hum Mol Genet. 2006;15:2200–2209. doi: 10.1093/hmg/ddl145. [DOI] [PubMed] [Google Scholar]
- Pin JP, Bettler B. Organization and functions of mGlu and GABAB receptor complexes. Nature. 2016;540:60–68. doi: 10.1038/nature20566. [DOI] [PubMed] [Google Scholar]
- Piret SE, Gorvin CM, Pagnamenta AT, Howles SA, Cranston T, Rust N, Nesbit MA, Glaser B, Taylor JC, Buchs AE, Hannan FM, Thakker RV. Identification of a G-Protein subunit-α11 gain-of-function mutation, Val340Met, in a family with autosomal dominant hypocalcemia type 2 (ADH2) J Bone Miner Res. 2016;31:1207–1214. doi: 10.1002/jbmr.2797. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Piret SE, Thakker RV. Mouse models for inherited endocrine and metabolic disorders. J Endocrinol. 2011;211:211–230. doi: 10.1530/JOE-11-0193. [DOI] [PubMed] [Google Scholar]
- Pollak MR, Brown EM, Chou YH, Hebert SC, Marx SJ, Steinmann B, Levi T, Seidman CE, Seidman JG. Mutations in the human Ca2+-sensing receptor gene cause familial hypocalciuric hypercalcemia and neonatal severe hyperparathyroidism. Cell. 1993;75:1297–1303. doi: 10.1016/0092-8674(93)90617-y. [DOI] [PubMed] [Google Scholar]
- Pollak MR, Brown EM, Estep HL, McLaine PN, Kifor O, Park J, Hebert SC, Seidman CE, Seidman JG. Autosomal dominant hypocalcaemia caused by a Ca2+-sensing receptor gene mutation. Nat Genet. 1994;8:303–307. doi: 10.1038/ng1194-303. [DOI] [PubMed] [Google Scholar]
- Procino G, Mastrofrancesco L, Tamma G, Lasorsa DR, Ranieri M, Stringini G, Emma F, Svelto M, Valenti G. Calcium-sensing receptor and aquaporin 2 interplay in hypercalciuria- associated renal concentrating defect in humans. An in vivo and in vitro study. PLoS One. 2012;7:e33145. doi: 10.1371/journal.pone.0033145. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Promkan M, Liu G, Patmasiriwat P, Chakrabarty S. BRCA1 suppresses the expression of survivin and promotes sensitivity to paclitaxel through the calcium sensing receptor (CaSR) in human breast cancer cells. Cell Calcium. 2011;49:79–88. doi: 10.1016/j.ceca.2011.01.003. [DOI] [PubMed] [Google Scholar]
- Quinn SJ, Bai M, Brown EM. pH Sensing by the calcium-sensing receptor. J Biol Chem. 2004;279:37241–37249. doi: 10.1074/jbc.M404520200. [DOI] [PubMed] [Google Scholar]
- Quinn SJ, Kifor O, Trivedi S, Diaz R, Vassilev P, Brown EM. Sodium and ionic strength sensing by the calcium receptor. J Biol Chem. 1998;273:19579–19586. doi: 10.1074/jbc.273.31.19579. [DOI] [PubMed] [Google Scholar]
- Quinn SJ, Ye CP, Diaz R, Kifor O, Bai M, Vassilev P, Brown E. The Ca2+-sensing receptor: a target for polyamines. Am J Physiol. 1997;273:C1315–1323. doi: 10.1152/ajpcell.1997.273.4.C1315. [DOI] [PubMed] [Google Scholar]
- Racke FK, Hammerland LG, Dubyak GR, Nemeth EF. Functional expression of the parathyroid cell calcium receptor in Xenopus oocytes. FEBS Lett. 1993;333:132–136. doi: 10.1016/0014-5793(93)80390-g. [DOI] [PubMed] [Google Scholar]
- Ramirez JA, Goodman WG, Gornbein J, Menezes C, Moulton L, Segre GV, Salusky IB. Direct in vivo comparison of calcium-regulated parathyroid hormone secretion in normal volunteers and patients with secondary hyperparathyroidism. J Clin Endocrinol Metab. 1993;76:1489–1494. doi: 10.1210/jcem.76.6.8501155. [DOI] [PubMed] [Google Scholar]
- Rasmussen AQ, Jorgensen NR, Schwarz P. Clinical and biochemical outcomes of cinacalcet treatment of familial hypocalciuric hypercalcemia: a case series. J Med Case Rep. 2011;5:564. doi: 10.1186/1752-1947-5-564. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Ratnam AV, Bikle DD, Cho JK. 1,25 dihydroxyvitamin D3 enhances the calcium response of keratinocytes. J Cell Physiol. 1999;178:188–196. doi: 10.1002/(SICI)1097-4652(199902)178:2<188::AID-JCP8>3.0.CO;2-4. [DOI] [PubMed] [Google Scholar]
- Ray JM, Squires PE, Curtis SB, Meloche MR, Buchan AM. Expression of the calcium- sensing receptor on human antral gastrin cells in culture. J Clin Invest. 1997;99:2328–2333. doi: 10.1172/JCI119413. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Ray K. Calcium-sensing receptor: Trafficking, endocytosis, recycling, and importance of interacting proteins. Prog Mol Biol Transl Sci. 2015;132:127–150. doi: 10.1016/bs.pmbts.2015.02.006. [DOI] [PubMed] [Google Scholar]
- Ray K, Clapp P, Goldsmith PK, Spiegel AM. Identification of the sites of N-linked glycosylation on the human calcium receptor and assessment of their role in cell surface expression and signal transduction. J Biol Chem. 1998;273:34558–34567. doi: 10.1074/jbc.273.51.34558. [DOI] [PubMed] [Google Scholar]
- Ray K, Hauschild BC, Steinbach PJ, Goldsmith PK, Hauache O, Spiegel AM. Identification of the cysteine residues in the amino-terminal extracellular domain of the human Ca(2+) receptor critical for dimerization. Implications for function of monomeric Ca(2+) receptor. J Biol Chem. 1999;274:27642–27650. doi: 10.1074/jbc.274.39.27642. [DOI] [PubMed] [Google Scholar]
- Ray K, Northup J. Evidence for distinct cation and calcimimetic compound (NPS 568) recognition domains in the transmembrane regions of the human Ca2+ receptor. J Biol Chem. 2002;277:18908–18913. doi: 10.1074/jbc.M202113200. [DOI] [PubMed] [Google Scholar]
- Reginster JY, Seeman E, De Vernejoul MC, Adami S, Compston J, Phenekos C, Devogelaer JP, Curiel MD, Sawicki A, Goemaere S, Sorensen OH, Felsenberg D, Meunier PJ. Strontium ranelate reduces the risk of nonvertebral fractures in postmenopausal women with osteoporosis: Treatment of Peripheral Osteoporosis (TROPOS) study. J Clin Endocrinol Metab. 2005;90:2816–2822. doi: 10.1210/jc.2004-1774. [DOI] [PubMed] [Google Scholar]
- Reh CM, Hendy GN, Cole DE, Jeandron DD. Neonatal hyperparathyroidism with a heterozygous calcium-sensing receptor (CaSR) R185Q mutation: clinical benefit from cinacalcet. J Clin Endocrinol Metab. 2011;96:E707–712. doi: 10.1210/jc.2010-1306. [DOI] [PubMed] [Google Scholar]
- Renkema KY, Velic A, Dijkman HB, Verkaart S, van der Kemp AW, Nowik M, Timmermans K, Doucet A, Wagner CA, Bindels RJ, Hoenderop JG. The calcium-sensing receptor promotes urinary acidification to prevent nephrolithiasis. J Am Soc Nephrol. 2009;20:1705–1713. doi: 10.1681/ASN.2008111195. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Rey O, Chang W, Bikle D, Rozengurt N, Young SH, Rozengurt E. Negative cross-talk between calcium-sensing receptor and β-catenin signaling systems in colonic epithelium. J Biol Chem. 2012;287:1158–1167. doi: 10.1074/jbc.M111.274589. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Rey O, Young SH, Papazyan R, Shapiro MS, Rozengurt E. Requirement of the TRPC1 cation channel in the generation of transient Ca2+ oscillations by the calcium-sensing receptor. J Biol Chem. 2006;281:38730–38737. doi: 10.1074/jbc.M605956200. [DOI] [PubMed] [Google Scholar]
- Rey O, Young SH, Yuan J, Slice L, Rozengurt E. Amino acid-stimulated Ca2+ oscillations produced by the Ca2+-sensing receptor are mediated by a phospholipase C/inositol 1,4,5-trisphosphate-independent pathway that requires G12, Rho, filamin-A, and the actin cytoskeleton. J Biol Chem. 2005;280:22875–22882. doi: 10.1074/jbc.M503455200. [DOI] [PubMed] [Google Scholar]
- Reyes-Ibarra AP, Garcia-Regalado A, Ramirez-Rangel I, Esparza-Silva AL, Valadez-Sanchez M, Vazquez-Prado J, Reyes-Cruz G. Calcium-sensing receptor endocytosis links extracellular calcium signaling to parathyroid hormone-related peptide secretion via a Rab11a-dependent and AMSH-sensitive mechanism. Mol Endocrinol. 2007;21:1394–1407. doi: 10.1210/me.2006-0523. [DOI] [PubMed] [Google Scholar]
- Riccardi D, Hall AE, Chattopadhyay N, Xu JZ, Brown EM, Hebert SC. Localization of the extracellular Ca2+/polyvalent cation-sensing protein in rat kidney. Am J Physiol. 1998;274:F611–622. doi: 10.1152/ajprenal.1998.274.3.F611. [DOI] [PubMed] [Google Scholar]
- Riccardi D, Kemp PJ. The calcium-sensing receptor beyond extracellular calcium homeostasis: conception, development, adult physiology, and disease. Annu Rev Physiol. 2012;74:271–297. doi: 10.1146/annurev-physiol-020911-153318. [DOI] [PubMed] [Google Scholar]
- Riccardi D, Park J, Lee WS, Gamba G, Brown EM, Hebert SC. Cloning and functional expression of a rat kidney extracellular calcium/polyvalent cation-sensing receptor. Proc Natl Acad Sci USA. 1995;92:131–135. doi: 10.1073/pnas.92.1.131. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Riccardi D, Valenti G. Localization and function of the renal calcium-sensing receptor. Nat Rev Nephrol. 2016;12:414–425. doi: 10.1038/nrneph.2016.59. [DOI] [PubMed] [Google Scholar]
- Rietsema S, Eelderink C, Joustra ML, van Vliet IMY, van Londen M, Corpeleijn E, Singh-Povel CM, Geurts JMW, Kootstra-Ros JE, Westerhuis R, Navis G, Bakker SJL. Effect of high compared with low dairy intake on blood pressure in overweight middle-aged adults: results of a randomized crossover intervention study. Am J Clin Nutr. 2019;110:340–348. doi: 10.1093/ajcn/nqz116. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Ritter CS, Haughey BH, Armbrecht HJ, Brown AJ. Distribution and regulation of the 25- hydroxyvitamin D3 1α-hydroxylase in human parathyroid glands. J Steroid Biochem Mol Biol. 2012;130:73–80. doi: 10.1016/j.jsbmb.2012.01.010. [DOI] [PubMed] [Google Scholar]
- Roberts MS, Gafni RI, Brillante B, Guthrie LC, Streit J, Gash D, Gelb J, Krusinska E, Brennan SC, Schepelmann M, Riccardi D, Bin Khayat ME, Ward DT, Nemeth EF, Rosskamp R, Collins MT. Treatment of Autosomal Dominant Hypocalcemia Type 1 with the Calcilytic NPSP795 (SHP635) J Bone Miner Res. 2019;34:1609–1618. doi: 10.1002/jbmr.3747. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Rodriguez L, Tu C, Cheng Z, Chen TH, Bikle D, Shoback D, Chang W. Expression and functional assessment of an alternatively spliced extracellular Ca2+-sensing receptor in growth plate chondrocytes. Endocrinology. 2005;146:5294–5303. doi: 10.1210/en.2005-0256. [DOI] [PubMed] [Google Scholar]
- Rodriguez-Hernandez CJ, Mateo-Lozano S, Garcia M, Casala C, Brianso F, Castrejon N, Rodriguez E, Sunol M, Carcaboso AM, Lavarino C, Mora J, de Torres C. Cinacalcet inhibits neuroblastoma tumor growth and upregulates cancer-testis antigens. Oncotarget. 2016;7:16112–16129. doi: 10.18632/oncotarget.7448. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Romano C, Yang WL, O'Malley KL. Metabotropic glutamate receptor 5 is a disulfide- linked dimer. J Biol Chem. 1996;271:28612–28616. doi: 10.1074/jbc.271.45.28612. [DOI] [PubMed] [Google Scholar]
- Rorsman P, Braun M, Zhang Q. Regulation of calcium in pancreatic α- and β-cells in health and disease. Cell Calcium. 2012;51:300–308. doi: 10.1016/j.ceca.2011.11.006. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Roszko KL, Bi R, Gorvin CM, Bräuner-Osborne H, Xiong XF, Inoue A, Thakker RV, Stromgaard K, Gardella T, Mannstadt M. Knockin mouse with mutant Gα11 mimics human inherited hypocalcemia and is rescued by pharmacologic inhibitors. JCI Insight. 2017;2:e91079. doi: 10.1172/jci.insight.91079. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Ruat M, Molliver ME, Snowman AM, Snyder SH. Calcium sensing receptor: molecular cloning in rat and localization to nerve terminals. Proc Natl Acad Sci USA. 1995;92:3161–3165. doi: 10.1073/pnas.92.8.3161. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Ruat M, Snowman AM, Hester LD, Snyder SH. Cloned and expressed rat Ca2+-sensing receptor. J Biol Chem. 1996;271:5972–5975. doi: 10.1074/jbc.271.11.5972. [DOI] [PubMed] [Google Scholar]
- Rus R, Haag C, Bumke-Vogt C, Bahr V, Mayr B, Mohlig M, Schulze E, Frank-Raue K, Raue F, Schofl C. Novel inactivating mutations of the calcium-sensing receptor: the calcimimetic NPS R-568 improves signal transduction of mutant receptors. J Clin Endocrinol Metab. 2008;93:4797–4803. doi: 10.1210/jc.2008-1076. [DOI] [PubMed] [Google Scholar]
- Rybchyn MS, Green WL, Conigrave AD, Mason RS. Involvement of both GPRC6A and the calcium-sensing receptor in strontium ranelate-induced osteoclastogenic signal expression and replication in primary human osteoblasts. Bone. 2009;44:S317–S317. [Google Scholar]
- Rybchyn MS, Islam KS, Brennan-Speranza TC, Cheng Z, Brennan SC, Chang W, Mason RS, Conigrave AD. Homer1 mediates CaSR-dependent activation of mTOR complex 2 and initiates a novel pathway for AKT-dependent β-catenin stabilization in osteoblasts. J Biol Chem. 2019;294:16337–16350. doi: 10.1074/jbc.RA118.006587. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Rybchyn MS, Slater M, Conigrave AD, Mason RS. An Akt-dependent increase in canonical Wnt signaling and a decrease in sclerostin protein levels are involved in strontium ranelate-induced osteogenic effects in human osteoblasts. J Biol Chem. 2011;286:23771–23779. doi: 10.1074/jbc.M111.251116. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Saidak Z, Boudot C, Abdoune R, Petit L, Brazier M, Mentaverri R, Kamel S. Extracellular calcium promotes the migration of breast cancer cells through the activation of the calcium sensing receptor. Exp Cell Res. 2009;315:2072–2080. doi: 10.1016/j.yexcr.2009.03.003. [DOI] [PubMed] [Google Scholar]
- Sands JM, Naruse M, Baum M, Jo I, Hebert SC, Brown EM, Harris HW. Apical extracellular calcium/polyvalent cation-sensing receptor regulates vasopressin-elicited water permeability in rat kidney inner medullary collecting duct. J Clin Invest. 1997;99:1399–1405. doi: 10.1172/JCI119299. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Santa Maria C, Cheng Z, Li A, Wang J, Shoback D, Tu CL, Chang W. Interplay between CaSR and PTH1R signaling in skeletal development and osteoanabolism. Semin Cell Dev Biol. 2016;49:11–23. doi: 10.1016/j.semcdb.2015.12.004. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Santa Maria C, Li A, Cheng Z, Song F, Shoback D, Tu C. ASBMR Annual Meeting. Washington State Convention Center: 2015. Skeletal anabolism by concurrently targeting the PTH1R and CaSR. [Google Scholar]
- Schauber J, Dorschner RA, Coda AB, Buchau AS, Liu PT, Kiken D, Helfrich YR, Kang S, Elalieh HZ, Steinmeyer A, Zugel U, Bikle DD, Modlin RL, Gallo RL. Injury enhances TLR2 function and antimicrobial peptide expression through a vitamin D-dependent mechanism. J Clin Invest. 2007;117:803–811. doi: 10.1172/JCI30142. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Schepelmann M, Yarova PL, Lopez-Fernandez I, Davies TS, Brennan SC, Edwards PJ, Aggarwal A, Graca J, Rietdorf K, Matchkov V, Fenton RA, Chang W, Krssak M, Stewart A, Broadley KJ, Ward DT, Price SA, Edwards DH, Kemp PJ, Riccardi D. The vascular Ca2+- sensing receptor regulates blood vessel tone and blood pressure. Am J Physiol Cell Physiol. 2016;310:C193–204. doi: 10.1152/ajpcell.00248.2015. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Sethi BK, Nagesh VS, Kelwade J, Parekh H, Dukle V. Utility of cinacalcet in familial hypocalciuric hypercalcemia. Indian J Endocrinol Metab. 2017;21:362–363. doi: 10.4103/2230-8210.202034. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Shalhoub V, Grisanti M, Padagas J, Scully S, Rattan A, Qi M, Varnum B, Vezina C, Lacey D, Martin D. In vitro studies with the calcimimetic, cinacalcet HCl, on normal human adult osteoblastic and osteoclastic cells. Crit Rev Eukaryot Gene Expr. 2003;13:89–106. doi: 10.1615/critreveukaryotgeneexpr.v13.i24.30. [DOI] [PubMed] [Google Scholar]
- Shams-White MM, Chung M, Du M, Fu Z, Insogna KL, Karlsen MC, LeBoff MS, Shapses SA, Sackey J, Wallace TC, Weaver CM. Dietary protein and bone health: a systematic review and meta-analysis from the National Osteoporosis Foundation. Am J Clin Nutr. 2017;105:1528–1543. doi: 10.3945/ajcn.116.145110. [DOI] [PubMed] [Google Scholar]
- Shcherbakova I, Huang G, Geoffroy OJ, Nair SK, Swierczek K, Balandrin MF, Fox J, Heaton WL, Conklin RL. Design, new synthesis, and calcilytic activity of substituted 3H- pyrimidin-4-ones. Bioorg Med Chem Lett. 2005;15:2537–2540. doi: 10.1016/j.bmcl.2005.03.054. [DOI] [PubMed] [Google Scholar]
- Shinagawa Y, Inoue T, Katsushima T, Kiguchi T, Ikenogami T, Ogawa N, Fukuda K, Hirata K, Harada K, Takagi M, Nakagawa T, Kimura S, Matsuo Y, Maekawa M, Hayashi M, Soejima Y, Takahashi M, Shindo M, Hashimoto H. Discovery of a potent and short-acting oral calcilytic with a pulsatile secretion of parathyroid hormone. ACS medicinal chemistry letters. 2011;2:238–242. doi: 10.1021/ml100268k. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Shoback DM, Membreno LA, McGhee JG. High calcium and other divalent cations increase inositol trisphosphate in bovine parathyroid cells. Endocrinology. 1988;123:382–389. doi: 10.1210/endo-123-1-382. [DOI] [PubMed] [Google Scholar]
- Silver IA, Murrills RJ, Etherington DJ. Microelectrode studies on the acid microenvironment beneath adherent macrophages and osteoclasts. Exp Cell Res. 1988;175:266–276. doi: 10.1016/0014-4827(88)90191-7. [DOI] [PubMed] [Google Scholar]
- Smajilovic S, Sheykhzade M, Holmegard HN, Haunso S, Tfelt-Hansen J. Calcimimetic, AMG 073, induces relaxation on isolated rat aorta. Vascul Pharmacol. 2007;47:222–228. doi: 10.1016/j.vph.2007.06.010. [DOI] [PubMed] [Google Scholar]
- Smajilovic S, Wellendorph P, Bräuner-Osborne H. Promiscuous seven transmembrane receptors sensing L-α-amino acids. Curr Pharm Des. 2014;20:2693–2702. doi: 10.2174/13816128113199990576. [DOI] [PubMed] [Google Scholar]
- Soda K, Kano Y, Sakuragi M, Takao K, Lefor A, Konishi F. Long-term oral polyamine intake increases blood polyamine concentrations. J Nutr Sci Vitaminol (Tokyo) 2009;55:361–366. doi: 10.3177/jnsv.55.361. [DOI] [PubMed] [Google Scholar]
- Stenson PD, Ball EV, Mort M, Phillips AD, Shaw K, Cooper DN. The Human Gene Mutation Database (HGMD) and its exploitation in the fields of personalized genomics and molecular evolution. Curr Protoc Bioinformatics. 2012;Chapter 1:13. doi: 10.1002/0471250953.bi0113s39. Unit1. [DOI] [PubMed] [Google Scholar]
- Stepanchick A, McKenna J, McGovern O, Huang Y, Breitwieser GE. Calcium sensing receptor mutations implicated in pancreatitis and idiopathic epilepsy syndrome disrupt an arginine-rich retention motif. Cell Physiol Biochem. 2010;26:363–374. doi: 10.1159/000320560. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Sun X, Tang L, Winesett S, Chang W, Cheng SX. Calcimimetic R568 inhibits tetrodotoxin-sensitive colonic electrolyte secretion and reduces c-fos expression in myenteric neurons. Life Sci. 2018;194:49–58. doi: 10.1016/j.lfs.2017.12.019. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Suzuki K, Lavaroni S, Mori A, Okajima F, Kimura S, Katoh R, Kawaoi A, Kohn L. Thyroid transcription factor 1 is calcium modulated and coordinately regulates genes involved in calcium homeostasis in C cells. Mol Cell Biol. 1998;18:7410–7422. doi: 10.1128/mcb.18.12.7410. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Tang H, Yamamura A, Yamamura H, Song S, Fraidenburg DR, Chen J, Gu Y, Pohl NM, Zhou T, Jimenez-Perez L, Ayon RJ, et al. Pathogenic role of calcium-sensing receptors in the development and progression of pulmonary hypertension. Am J Physiol Lung Cell Mol Physiol. 2016a;310:L846–859. doi: 10.1152/ajplung.00050.2016. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Tang L, Cheng CY, Sun X, Pedicone AJ, Mahamadzadeh M, Cheng SX. The extracellular calcium-sensing receptor in the intestine: Evidence for regulation of colonic absorption, secretion, motility, and immunity. Front Physiol. 2016b;7:245. doi: 10.3389/fphys.2016.00245. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Tang L, Jiang L, McIntyre ME, Petrova E, Cheng SX. Calcimimetic acts on enteric neuronal CaSR to reverse cholera toxin-induced intestinal electrolyte secretion. Sci Rep. 2018;8 doi: 10.1038/s41598-018-26171-4. 7851. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Tang L, Peng M, Liu L, Chang W, Binder HJ, Cheng SX. Calcium-sensing receptor stimulates Cl(-)- and SCFA-dependent but inhibits cAMP-dependent HCO3(-) secretion in colon. Am J Physiol Gastrointest Liver Physiol. 2015;308:G874–883. doi: 10.1152/ajpgi.00341.2014. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Tatsumi R, Komaba H, Kanai G, Miyakogawa T, Sawada K, Kakuta T, Fukagawa M. Cinacalcet induces apoptosis in parathyroid cells in patients with secondary hyperparathyroidism: histological and cytological analyses. Nephron Clin Pract. 2013;124:224–231. doi: 10.1159/000357951. [DOI] [PubMed] [Google Scholar]
- Temal T, Jary H, Auberval M, Lively S, Guedin D, Vevert JP, Deprez P. New potent calcimimetics: I. Discovery of a series of novel trisubstituted ureas. Bioorg Med Chem Lett. 2013;23:2451–2454. doi: 10.1016/j.bmcl.2013.01.078. [DOI] [PubMed] [Google Scholar]
- Tenhola S, Hendy GN, Valta H, Canaff L, Lee BS, Wong BY, Valimaki MJ, Cole DE, Makitie O. Cinacalcet treatment in an adolescent with concurrent 22q11.2 deletion syndrome and FHH3 caused by AP2S1 mutation. J Clin Endocrinol Metab. 2015;100:2515–2518. doi: 10.1210/jc.2015-1518. [DOI] [PubMed] [Google Scholar]
- Tenhola S, Voutilainen R, Reyes M, Toiviainen-Salo S, Juppner H, Makitie O. Impaired growth and intracranial calcifications in autosomal dominant hypocalcemia caused by a GNA11 mutation. Eur J Endocrinol. 2016;175:211–218. doi: 10.1530/EJE-16-0109. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Tfelt-Hansen J, MacLeod RJ, Chattopadhyay N, Yano S, Quinn S, Ren X, Terwilliger EF, Schwarz P, Brown EM. Calcium-sensing receptor stimulates PTHrP release by pathways dependent on PKC, p38 MAPK, JNK, and ERK1/2 in H-500 cells. Am J Physiol. 2003;285:E329–E337. doi: 10.1152/ajpendo.00489.2002. [DOI] [PubMed] [Google Scholar]
- Thakore P, Ho WS. Vascular actions of calcimimetics: role of Ca2(+) -sensing receptors versus Ca2(+) influx through L-type Ca2(+) channels. Br J Pharmacol. 2011;162:749–762. doi: 10.1111/j.1476-5381.2010.01079.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Thomsen AR, Hvidtfeldt M, Bräuner-Osborne H. Biased agonism of the calcium- sensing receptor. Cell Calcium. 2012a;51:107–116. doi: 10.1016/j.ceca.2011.11.009. [DOI] [PubMed] [Google Scholar]
- Thomsen AR, Worm J, Jacobsen SE, Stahlhut M, Latta M, Bräuner-Osborne H. Strontium is a biased agonist of the calcium-sensing receptor in rat medullary thyroid carcinoma 6-23 cells. J Pharmacol Exp Ther. 2012b;343:638–649. doi: 10.1124/jpet.112.197210. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Toka HR, Al-Romaih K, Koshy JM, DiBartolo S, Kos CH, Quinn SJ, Curhan GC, Mount DB, Brown EM, Pollak MR. Deficiency of the calcium-sensing receptor in the kidney causes parathyroid hormone-independent hypocalciuria. J Am Soc Nephrol. 2012;23:1879–1890. doi: 10.1681/ASN.2012030323. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Trzaskowski B, Latek D, Yuan S, Ghoshdastider U, Debinski A, Filipek S. Action of molecular switches in GPCRs--theoretical and experimental studies. Curr Med Chem. 2012;19:1090–1109. doi: 10.2174/092986712799320556. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Tsuchiya D, Kunishima N, Kamiya N, Jingami H, Morikawa K. Structural views of the ligand-binding cores of a metabotropic glutamate receptor complexed with an antagonist and both glutamate and Gd3+ Proc Natl Acad Sci USA. 2002;99:2660–2665. doi: 10.1073/pnas.052708599. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Tsutsumi M, Goto M, Denda M. Dynamics of intracellular calcium in cultured human keratinocytes after localized cell damage. Exp Dermatol. 2013;22:367–369. doi: 10.1111/exd.12136. [DOI] [PubMed] [Google Scholar]
- Tu CL, Bikle DD. Role of the calcium-sensing receptor in calcium regulation of epidermal differentiation and function. Best Pract Res Clin Endocrinol Metab. 2013;27:415–427. doi: 10.1016/j.beem.2013.03.002. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Tu CL, Celli A, Mauro T, Chang W. Calcium-sensing receptor regulates epidermal intracellular Ca2+ signaling and re-epithelialization after wounding. J Invest Dermatol. 2019;139:919–929. doi: 10.1016/j.jid.2018.09.033. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Tu CL, Chang WH, Xie ZJ, Bikle DD. Inactivation of the calcium sensing receptor inhibits E-cadherin-mediated cell-cell adhesion and calcium-induced differentiation in human epidermal keratinocytes. J Biol Chem. 2008;283:3519–3528. doi: 10.1074/jbc.M708318200. [DOI] [PubMed] [Google Scholar]
- Tu CL, Crumrine DA, Man MQ, Chang W, Elalieh H, You M, Elias PM, Bikle DD. Ablation of the calcium-sensing receptor in keratinocytes impairs epidermal differentiation and barrier function. The Journal of investigative dermatology. 2012;132:2350–2359. doi: 10.1038/jid.2012.159. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Tu Q, Pi M, Karsenty G, Simpson L, Liu S, Quarles LD. Rescue of the skeletal phenotype in CasR-deficient mice by transfer onto the Gcm2 null background. J Clin Invest. 2003;111:1029–1037. doi: 10.1172/JCI17054. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Turksen K, Troy TC. Overexpression of the calcium sensing receptor accelerates epidermal differentiation and permeability barrier formation in vivo. Mech Dev. 2003;120:733–744. doi: 10.1016/s0925-4773(03)00045-5. [DOI] [PubMed] [Google Scholar]
- van den Berg MP, Meurs H, Gosens R. Targeting arginase and nitric oxide metabolism in chronic airway diseases and their co-morbidities. Curr Opin Pharmacol. 2018;40:126–133. doi: 10.1016/j.coph.2018.04.010. [DOI] [PubMed] [Google Scholar]
- Vancleef L, Van Den Broeck T, Thijs T, Steensels S, Briand L, Tack J, Depoortere I. Chemosensory signalling pathways involved in sensing of amino acids by the ghrelin cell. Sci Rep. 2015;5 doi: 10.1038/srep15725. 15725. [DOI] [PMC free article] [PubMed] [Google Scholar]
- VanHouten J, Dann P, McGeoch G, Brown EM, Krapcho K, Neville M, Wysolmerski JJ. The calcium-sensing receptor regulates mammary gland parathyroid hormone-related protein production and calcium transport. J Clin Invest. 2004;113:598–608. doi: 10.1172/JCI18776. [DOI] [PMC free article] [PubMed] [Google Scholar]
- VanHouten JN, Neville MC, Wysolmerski JJ. The calcium-sensing receptor regulates plasma membrane calcium adenosine triphosphatase isoform 2 activity in mammary epithelial cells: a mechanism for calcium-regulated calcium transport into milk. Endocrinology. 2007;148:5943–5954. doi: 10.1210/en.2007-0850. [DOI] [PMC free article] [PubMed] [Google Scholar]
- VanHouten JN, Wysolmerski JJ. Transcellular calcium transport in mammary epithelial cells. J Mammary Gland Biol Neoplasia. 2007;12:223–235. doi: 10.1007/s10911-007-9057-1. [DOI] [PubMed] [Google Scholar]
- Vanhouten JN, Wysolmerski JJ. The calcium-sensing receptor in the breast. Best Pract Res Clin Endocrinol Metab. 2013;27:403–414. doi: 10.1016/j.beem.2013.02.011. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Vargas-Poussou R, Mansour-Hendili L, Baron S, Bertocchio JP, Travers C, Simian C, Treard C, Baudouin V, Beltran S, Broux F, Camard O, Cloarec S, Cormier C, Debussche X, Dubosclard E, Eid C, Haymann JP, Romuald Kiando S, Kuhn JM, Lefort G, Linglart A, Lucas-Pouliquen B, Macher MA, Maruani G, Ouzounian S, Polak M, Requeda E, Robier D, Silve C, Souberbielle JC, Tack I, Vezzosi D, Jeunemaitre X, Houillier P. Familial hypocalciuric hypercalcemia types 1 and 3 and primary hyperparathyroidism: similarities and differences. J Clin Endocrinol Metab. 2016;101:2185–2195. doi: 10.1210/jc.2015-3442. [DOI] [PubMed] [Google Scholar]
- Vezzoli G, Scillitani A, Corbetta S, Terranegra A, Dogliotti E, Guarnieri V, Arcidiacono T, Paloschi V, Rainone F, Eller-Vainicher C, Borghi L, Nouvenne A, Guerra A, Meschi T, Allegri F, Cusi D, Spada A, Cole DE, Hendy GN, Spotti D, Soldati L. Polymorphisms at the regulatory regions of the CASR gene influence stone risk in primary hyperparathyroidism. Eur J Endocrinol. 2011;164:421–427. doi: 10.1530/EJE-10-0915. [DOI] [PubMed] [Google Scholar]
- Vizard TN, O'Keeffe GW, Gutierrez H, Kos CH, Riccardi D, Davies AM. Regulation of axonal and dendritic growth by the extracellular calcium-sensing receptor. Nat Neurosci. 2011;11:285–291. doi: 10.1038/nn2044. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Walter S, Baruch A, Dong J, Tomlinson JE, Alexander ST, Janes J, Hunter T, Yin Q, Maclean D, Bell G, Mendel DB, Johnson RM, Karim F. Pharmacology of AMG 416 (Velcalcetide), a novel peptide agonist of the calcium-sensing receptor, for the treatment of secondary hyperparathyroidism in hemodialysis patients. J Pharmacol Exp Ther. 2013;346:229–240. doi: 10.1124/jpet.113.204834. [DOI] [PubMed] [Google Scholar]
- Wang L, Widatalla SE, Whalen DS, Ochieng J, Sakwe AM. Association of calcium sensing receptor polymorphisms at rs1801725 with circulating calcium in breast cancer patients. BMC Cancer. 2017;17:511. doi: 10.1186/s12885-017-3502-3. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Wang M, Yao Y, Kuang D, Hampson DR. Activation of family C G-protein-coupled receptors by the tripeptide glutathione. J Biol Chem. 2006;281:8864–8870. doi: 10.1074/jbc.M512865200. [DOI] [PubMed] [Google Scholar]
- Wang X, Kirberger M, Qiu F, Chen G, Yang JJ. Towards predicting Ca2+-binding sites with different coordination numbers in proteins with atomic resolution. Proteins. 2009;75:787–798. doi: 10.1002/prot.22285. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Wang X, Zhao K, Kirberger M, Wong H, Chen G, Yang JJ. Analysis and prediction of calcium-binding pockets from apo-protein structures exhibiting calcium-induced localized conformational changes. Protein Sci. 2010;19:1180–1190. doi: 10.1002/pro.394. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Wang Y, Chandra R, Samsa LA, Gooch B, Fee BE, Cook JM, Vigna SR, Grant AO, Liddle RA. Amino acids stimulate cholecystokinin release through the Ca2+-sensing receptor. Am J Physiol Gastrointest Liver Physiol. 2011;300:G528–537. doi: 10.1152/ajpgi.00387.2010. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Ward DT, Brown EM, Harris HW. Disulfide bonds in the extracellular calcium- polyvalent cation-sensing receptor correlate with dimer formation and its response to divalent cations in vitro. J Biol Chem. 1998;273:14476–14483. doi: 10.1074/jbc.273.23.14476. [DOI] [PubMed] [Google Scholar]
- Watanabe S, Fukumoto S, Chang H, Takeuchi Y, Hasegawa Y, Okazaki R, Chikatsu N, Fujita T. Association between activating mutations of calcium-sensing receptor and Bartter’s syndrome. Lancet. 2002;360:692–694. doi: 10.1016/S0140-6736(02)09842-2. [DOI] [PubMed] [Google Scholar]
- Wei G, Ku S, Ma GK, Saito S, Tang AA, Zhang J, Mao JH, Appella E, Balmain A, Huang EJ. HIPK2 represses beta-catenin-mediated transcription, epidermal stem cell expansion, and skin tumorigenesis. Proc Natl Acad Sci USA. 2007;104:13040–13045. doi: 10.1073/pnas.0703213104. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Wellendorph P, Bräuner-Osborne H. Molecular basis for amino acid sensing by family C G protein-coupled receptors. Br J Pharmacol. 2009;156:869–884. doi: 10.1111/j.1476-5381.2008.00078.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Wettschureck N, Lee E, Libutti SK, Offermanns S, Robey PG, Spiegel AM. Parathyroid- specific double knockout of Gq and G11 alpha-subunits leads to a phenotype resembling germline knockout of the extracellular Ca2+ -sensing receptor. Mol Endocrinol. 2007;21:274–280. doi: 10.1210/me.2006-0110. [DOI] [PubMed] [Google Scholar]
- White E, McKenna J, Cavanaugh A, Breitwieser GE. Pharmacochaperone-mediated rescue of calcium-sensing receptor loss-of-function mutants. Mol Endocrinol. 2009;23:1115–1123. doi: 10.1210/me.2009-0041. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Wolf P, Krssak M, Winhofer Y, Anderwald CH, Zwettler E, Just Kukurova I, Gessl A, Trattnig S, Luger A, Baumgartner-Parzer S, Krebs M. Cardiometabolic phenotyping of patients with familial hypocalcuric hypercalcemia. J Clin Endocrinol Metab. 2014;99:E1721–1726. doi: 10.1210/jc.2014-1541. [DOI] [PubMed] [Google Scholar]
- Wu H, Wang C, Gregory KJ, Han GW, Cho HP, Xia Y, Niswender CM, Katritch V, Meiler J, Cherezov V, Conn PJ, Stevens RC. Structure of a class C GPCR metabotropic glutamate receptor 1 bound to an allosteric modulator. Science. 2014;344:58–64. doi: 10.1126/science.1249489. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Wu W, Zhou HR, Bursian SJ, Link JE, Pestka JJ. Calcium-sensing receptor and transient receptor ankyrin-1 mediate emesis induction by deoxynivalenol (Vomitoxin) Toxicol Sci. 2017;155:32–42. doi: 10.1093/toxsci/kfw191. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Wysolmerski JJ. Parathyroid hormone-related protein: an update. J Clin Endocrinol Metab. 2012;97:2947–2956. doi: 10.1210/jc.2012-2142. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Xie R, Xu J, Xiao Y, Wu J, Wan H, Tang B, Liu J, Fan Y, Wang S, Wu Y, Dong TX, Zhu MX, Carethers JM, Dong H, Yang S. Calcium promotes human gastric cancer via a novel coupling of calcium-sensing receptor and TRPV4 channel. Cancer Res. 2017;77:6499–6512. doi: 10.1158/0008-5472.CAN-17-0360. [DOI] [PubMed] [Google Scholar]
- Yamada S, Tokumoto M, Taniguchi M, Toyonaga J, Suehiro T, Eriguchi R, Fujimi S, Ooboshi H, Kitazono T, Tsuruya K. Two years of cinacalcet hydrochloride treatment decreased parathyroid gland volume and serum parathyroid hormone level in hemodialysis patients with advanced secondary hyperparathyroidism. Ther Apher Dial. 2015;19:367–377. doi: 10.1111/1744-9987.12292. [DOI] [PubMed] [Google Scholar]
- Yamamoto M, Akatsu T, Nagase T, Ogata E. Comparison of hypocalcemic hypercalciuria between patients with idiopathic hypoparathyroidism and those with gain-of- function mutations in the calcium-sensing receptor: is it possible to differentiate the two disorders? J Clin Endocrinol Metab. 2000;85:4583–4591. doi: 10.1210/jcem.85.12.7035. [DOI] [PubMed] [Google Scholar]
- Yang C, Rybchyn MS, Conigrave AD, Mason RS. Role of Calcium sensing receptor (CaSR) modifiers in photoprotection. Wound Repair Regen. 2016;24:A16. [Google Scholar]
- Yang W, Liu L, Masugi Y, Qian ZR, Nishihara R, Keum N, Wu K, Smith-Warner S, Ma Y, Nowak JA, Momen-Heravi F, Zhang L, Bowden M, Morikawa T, Silva AD, Wang M, Chan AT, Fuchs CS, Meyerhardt JA, Ng K, Giovannucci E, Ogino S, Zhang X. Calcium intake and risk of colorectal cancer according to expression status of calcium-sensing receptor (CASR) Gut. 2017;67:1475–1483. doi: 10.1136/gutjnl-2017-314163. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Yang Y, Wang B. PTH1R-CaSR Cross Talk: New Treatment Options for Breast Cancer Osteolytic Bone Metastases. Int J Endocrinol. 2018;2018 doi: 10.1155/2018/7120979. 7120979. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Yarova PL, Stewart AL, Sathish V, Britt RD, Jr, Thompson MA, AP PL, Freeman M, Aravamudan B, Kita H, Brennan SC, Schepelmann M, Davies T, Yung S, Cholisoh Z, Kidd EJ, Ford WR, Broadley KJ, Rietdorf K, Chang W, Bin Khayat ME, Ward DT, Corrigan CJ, JP TW, Kemp PJ, Pabelick CM, Prakash YS, Riccardi D. Calcium-sensing receptor antagonists abrogate airway hyperresponsiveness and inflammation in allergic asthma. Sci Transl Med. 2015;7 doi: 10.1126/scitranslmed.aaa0282. 284ra260. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Ye C, Ho-Pao CL, Kanazirska M, Quinn S, Rogers K, Seidman CE, Seidman JG, Brown EM, Vassilev PM. Amyloid-beta proteins activate Ca(2+)-permeable channels through calcium-sensing receptors. J Neurosci Res. 1997a;47:547–554. doi: 10.1002/(sici)1097-4547(19970301)47:5<547::aid-jnr10>3.0.co;2-v. [DOI] [PubMed] [Google Scholar]
- Ye C, Ho-Pao CL, Kanazirska M, Quinn S, Seidman CE, Seidman JG, Brown EM, Vassilev PM. Deficient cation channel regulation in neurons from mice with targeted disruption of the extracellular Ca2+-sensing receptor gene. Brain Res Bull. 1997b;44:75–84. doi: 10.1016/s0361-9230(97)00088-9. [DOI] [PubMed] [Google Scholar]
- Ye C, Rogers K, Bai M, Quinn SJ, Brown EM, Vassilev PM. Agonists of the Ca(2+)- sensing receptor (CaR) activate nonselective cation channels in HEK293 cells stably transfected with the human CaR. Biochem Biophys Res Commun. 1996;226:572–579. doi: 10.1006/bbrc.1996.1396. [DOI] [PubMed] [Google Scholar]
- Young SH, Rey O, Sinnett-Smith J, Rozengurt E. Intracellular Ca2+ oscillations generated via the Ca2+-sensing receptor are mediated by negative feedback by PKCalpha at Thr888. Am J Physiol Cell Physiol. 2014;306:C298–306. doi: 10.1152/ajpcell.00194.2013. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Youssef KK, Lapouge G, Bouvree K, Rorive S, Brohee S, Appelstein O, Larsimont JC, Sukumaran V, Van de Sande B, Pucci D, Dekoninck S, Berthe JV, Aerts S, Salmon I, del Marmol V, Blanpain C. Adult interfollicular tumour-initiating cells are reprogrammed into an embryonic hair follicle progenitor-like fate during basal cell carcinoma initiation. Nat Cell Biol. 2012;14:1282–1294. doi: 10.1038/ncb2628. [DOI] [PubMed] [Google Scholar]
- Zaidi M, Kerby J, Huang CL, Alam T, Rathod H, Chambers TJ, Moonga BS. Divalent cations mimic the inhibitory effect of extracellular ionised calcium on bone resorption by isolated rat osteoclasts: further evidence for a “calcium receptor”. J Cell Physiol. 1991;149:422–427. doi: 10.1002/jcp.1041490310. [DOI] [PubMed] [Google Scholar]
- Zhang C, Huang Y, Jiang Y, Mulpuri N, Wei L, Hamelberg D, Brown EM, Yang JJ. Identification of an L-phenylalanine binding site enhancing the cooperative responses of the calcium-sensing receptor to calcium. J Biol Chem. 2014a;289:5296–5309. doi: 10.1074/jbc.M113.537357. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Zhang C, Zhang T, Zou J, Miller CL, Gorkhali R, Yang JY, Schilmiller A, Wang S, Huang K, Brown EM, Moremen KW, Hu J, Yang JJ. Structural basis for regulation of human calcium-sensing receptor by magnesium ions and an unexpected tryptophan derivative co-agonist. Sci Adv. 2016;2:e1600241. doi: 10.1126/sciadv.1600241. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Zhang C, Zhuo Y, Moniz HA, Wang S, Moremen KW, Prestegard JH, Brown EM, Yang JJ. Direct determination of multiple ligand interactions with the extracellular domain of the calcium-sensing receptor. J Biol Chem. 2014b;289:33529–33542. doi: 10.1074/jbc.M114.604652. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Zhang H, Kovacs-Nolan J, Kodera T, Eto Y, Mine Y. γ-Glutamyl cysteine and γ- glutamyl valine inhibit TNF-alpha signaling in intestinal epithelial cells and reduce inflammation in a mouse model of colitis via allosteric activation of the calcium-sensing receptor. Biochim Biophys Acta. 2015;1852:792–804. doi: 10.1016/j.bbadis.2014.12.023. [DOI] [PubMed] [Google Scholar]
- Zhang M, Yang H, Wan X, Lu L, Zhang J, Zhang H, Ye T, Liu Q, Xie M, Liu X, Yu S, Guo S, Chang W, Wang M. Prevention of Injury-Induced Osteoarthritis in Rodent Temporomandibular Joint by Targeting Chondrocyte CaSR. J Bone Miner Res. 2018 doi: 10.1002/jbmr.3643. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Zhang M, Yang H, Wan X, Lu L, Zhang J, Zhang H, Ye T, Liu Q, Xie M, Liu X, Yu S, Guo S, Chang W, Wang M. Prevention of injury-induced osteoarthritis in rodent temporomandibular joint by targeting chondrocyte CaSR. J Bone Miner Res. 2019;34:726–738. doi: 10.1002/jbmr.3643. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Zhang Z, Jiang Y, Quinn SJ, Krapcho K, Nemeth EF, Bai M. L-Phenylalanine and NPS R-467 synergistically potentiate the function of the extracellular calcium-sensing receptor through distinct sites. J Biol Chem. 2002a;277:33736–33741. doi: 10.1074/jbc.M200978200. [DOI] [PubMed] [Google Scholar]
- Zhang Z, Qiu W, Quinn SJ, Conigrave AD, Brown EM, Bai M. Three adjacent serines in the extracellular domains of the CaR are required for L-amino acid-mediated potentiation of receptor function. J Biol Chem. 2002b;277:33727–33735. doi: 10.1074/jbc.M200976200. [DOI] [PubMed] [Google Scholar]
- Zhang Z, Sun S, Quinn SJ, Brown EM, Bai M. The extracellular calcium-sensing receptor dimerizes through multiple types of intermolecular interactions. J Biol Chem. 2001;276:5316–5322. doi: 10.1074/jbc.M005958200. [DOI] [PubMed] [Google Scholar]
- Zhao K, Wang X, Wong HC, Wohlhueter R, Kirberger MP, Chen G, Yang JJ. Predicting Ca2+ -binding sites using refined carbon clusters. Proteins. 2012;80:2666–2679. doi: 10.1002/prot.24149. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Zhao X, Xian Y, Wang C, Ding L, Meng X, Zhu W, Hang S. Calcium-sensing receptor- mediated L-tryptophan-induced secretion of cholecystokinin and glucose-dependent insulinotropic peptide in swine duodenum. J Vet Sci. 2018;19:179–187. doi: 10.4142/jvs.2018.19.2.179. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Zhuang X, Northup JK, Ray K. Large putative PEST-like sequence motif at the carboxyl tail of human calcium receptor directs lysosomal degradation and regulates cell surface receptor level. J Biol Chem. 2012;287:4165–4176. doi: 10.1074/jbc.M111.271528. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Zinser GM, Sundberg JP, Welsh J. Vitamin D(3) receptor ablation sensitizes skin to chemically induced tumorigenesis. Carcinogenesis. 2002;23:2103–2109. doi: 10.1093/carcin/23.12.2103. [DOI] [PubMed] [Google Scholar]
- Zuo X, Sun L, Yin X, Gao J, Sheng Y, Xu J, Zhang J, He C, Qiu Y, Wen G, Tian H, Zheng X, Liu S, Wang W, Li W, Cheng Y, Liu L, Chang Y, Wang Z, Li Z, Li L, Wu J, Fang L, Shen C, Zhou F, Liang B, Chen G, Li H, Cui Y, Xu A, Yang X, Hao F, Xu L, Fan X, Li Y, Wu R, Wang X, Liu X, Zheng M, Song S, Ji B, Fang H, Yu J, Sun Y, Hui Y, Zhang F, Yang R, Yang S, Zhang X. Whole-exome SNP array identifies 15 new susceptibility loci for psoriasis. Nat Commun. 2015;6 doi: 10.1038/ncomms7793. 6793. [DOI] [PMC free article] [PubMed] [Google Scholar]