[go: up one dir, main page]
More Web Proxy on the site http://driver.im/
Next Article in Journal
Ultrasonic Microfluidic Method Used for siHSP47 Loaded in Human Embryonic Kidney Cell-Derived Exosomes for Inhibiting TGF-β1 Induced Fibroblast Differentiation and Migration
Previous Article in Journal
Chemoprotective Mechanism of Sodium Thiosulfate Against Cisplatin-Induced Nephrotoxicity Is via Renal Hydrogen Sulfide, Arginine/cAMP and NO/cGMP Signaling Pathways
Previous Article in Special Issue
OCTN1 (SLC22A4) as a Target of Heavy Metals: Its Possible Role in Microplastic Threats
You seem to have javascript disabled. Please note that many of the page functionalities won't work as expected without javascript enabled.
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

The Molecular Biology of Placental Transport of Calcium to the Human Foetus

Department of Clinical Biochemistry, University Hospital Southampton NHS Foundation Trust, Southampton General Hospital, Southampton SO16 6YD, UK
Int. J. Mol. Sci. 2025, 26(1), 383; https://doi.org/10.3390/ijms26010383
Submission received: 30 October 2024 / Revised: 23 December 2024 / Accepted: 30 December 2024 / Published: 4 January 2025
(This article belongs to the Special Issue Transport of Nutrients and Ions Relevant to Human Pathophysiology)
Figure 1
<p>Development and implantation of the blastocyst. Compiled from Burton GJ and Jauniaux E 2021 [<a href="#B6-ijms-26-00383" class="html-bibr">6</a>], Mitchell B and Sharma R 2005 [<a href="#B39-ijms-26-00383" class="html-bibr">39</a>], Harrison RG 1963 [<a href="#B40-ijms-26-00383" class="html-bibr">40</a>], and Sadler TW 2023 [<a href="#B41-ijms-26-00383" class="html-bibr">41</a>]; ~, approximately.</p> ">
Figure 2
<p>Section of a human placenta showing three cotyledons separated by septa (S). Within each of these is a mass of foetal villi branching from a primary villous stem anchored to the decidua basalis, which forms the maternal placenta. The villi are covered by trophoblasts: an inner layer of cytotrophoblasts and an outer layer of syncytiotrophoblasts. These erode the walls of small uterine spiral arteries, and the blood empties into the intervillous spaces. The highly branched terminal villi (red arrows) float freely in a lake of maternal blood. Nutrients and minerals pass from mother to foetus, but there is no continuity between foetal and maternal circulations. AM, amnion; U.A., umbilical arteries; U.V., umbilical vein; UT.A., uterine artery; V., uterine vein. Source: Harrison RG: A Textbook of Human Embryology 2nd Ed. Blackwell Scientific Publications Ltd. Oxford 1963 [<a href="#B40-ijms-26-00383" class="html-bibr">40</a>]. Figure 35 reproduced with permission from Wiley-Blackwell, John Wiley and Sons.</p> ">
Figure 3
<p>Proposed scheme in which ezrin bridges the scaffolding protein NHERF1 and the actin skeleton, so stabilising proteins in the microvillar membrane and connecting them with cytosolic signalling complexes. (i) In a dormant form, the C-terminal tail of ezrin binds to the N-terminal and closes the molecule. (ii) Phosphorylation of residues located between the N- and C-terminals blocks the association and opens the ezrin molecule. (iii) The freed C-terminal binds to actin; the N-terminal binds to the scaffolding protein NHERF1 associated with the plasma membrane. (iv) NHERF1 connects membrane-associated proteins with transiently assembled cytosolic signalling complexes. NaPi-2a, PTH1R, and NHE3 are shown as examples. NaPi-2a, sodium–phosphate cotransporter 2a; NHE3, sodium–hydrogen exchanger 3; NHERF1, Na<sup>+</sup>/H<sup>+</sup> exchanger regulatory factor 1; PTH1R, parathyroid hormone 1 receptor.</p> ">
Figure 4
<p>Schematic of Na<sup>+</sup>/H<sup>+</sup> exchanger regulatory factor 1 (NHERF1) based on the models of Zhang et al. (2019) [<a href="#B62-ijms-26-00383" class="html-bibr">62</a>] and Bhattycharia et al. (2019) [<a href="#B63-ijms-26-00383" class="html-bibr">63</a>]. GYGF, core PDZ-binding motif; PPI, protein-phosphatase-1-binding site; S290, key phosphorylation site.</p> ">
Figure 5
<p>Phospholipase C releases inositol trisphosphate from phosphatidylinositol 4,5 bisphosphate (PIP2). Arachidonic acid is released from PIP2 by phospholipase A2.</p> ">
Figure 6
<p>Structure of TRPV6, as deduced from structural studies (PDB codes 5IWK, 5IWP, and 62EF) [<a href="#B132-ijms-26-00383" class="html-bibr">132</a>,<a href="#B133-ijms-26-00383" class="html-bibr">133</a>,<a href="#B134-ijms-26-00383" class="html-bibr">134</a>]. (<b>A</b>) Side and (<b>B</b>) top views of the four TRPV6 monomers, displayed as ribbons with α-helices and β-sheets as cylinders and arrows, respectively. Three monomers are coloured uniformly (purple, blue, and dark pink); one monomer is differentially coloured to highlight crucial regions: N-terminal helix (yellow), ankyrin repeats (cyan), intracellular N-terminal region (dark green), membrane region (green), and intra-cellular C-terminal region (orange). The Ca<sup>2+</sup> ion in the pore is displayed as a yellow ball. (<b>C</b>) Schematic overview of secondary structural elements and other structural features. Colour coding as in panels (<b>A</b>,<b>B</b>). Note that the C-terminal helices 1–3 (CtH1–3; light orange) are not observed in the 5IWK structure displayed in panels (<b>A</b>,<b>B</b>,<b>D</b>). (<b>D</b>) Close-up of the TRPV6 pore-forming region. The ion selectivity filter forming the side chains of D542 and surrounding residues are shown in ball-and-stick representation, colour-coded on a by-atom-type basis. The carboxylic groups lining the pore are clearly visible. The ion selectivity filter is the narrowest part of the channel. This figure was published in <span class="html-italic">Advances in Clinical Chemistry</span>, vol. 113, Walker V, Vuister GW, Biochemistry and pathophysiology of the transient potential receptor vanilloid 6 (TRPV6) calcium channel, pp 43–100, 2023 [<a href="#B115-ijms-26-00383" class="html-bibr">115</a>]. Copyright Elsevier. Reproduced with permission.</p> ">
Figure 7
<p>The domains of STIM1 and their functions [<a href="#B61-ijms-26-00383" class="html-bibr">61</a>,<a href="#B185-ijms-26-00383" class="html-bibr">185</a>,<a href="#B187-ijms-26-00383" class="html-bibr">187</a>]: SP, signal peptide; EF1 and EF2, EF-hand domains; SAM, sterile alpha motif; TM, transmembrane domain; CC1, CC2, and CC3, coiled-coil domains; CAD, CRAC activation domain; SOAR, STIM-Orai activating domain; ID, inactivation domain; STIM1L, a peptide insert; P/S, proline/serine-rich peptide; EB, EB1-binding domain; PBD, polybasic domain; U, peptide sequences.</p> ">
Figure 8
<p>Binding of activated STIM1 to Orai1 to open the Orai1 channel for Ca<sup>2+</sup> influx. SAM, sterile alpha motif; TM, transmembrane domain; CC1, CC2, and CC3, coiled-coil domains; PBD, polybasic domain; PIP2, phosphatidylinositol 4,5 bisphosphate.</p> ">
Figure 9
<p>Store-operated calcium entry (SOCE). (<b>A</b>) (1) Agonist stimulation of a membrane receptor (2) activates phospholipase C (PLC), releasing inositol trisphosphate (IP3), and (3) IP3 activates the IP3 receptor in the ER membrane, leading to release of Ca<sup>2+</sup> into the cytosol to activate signalling cascades. Some Ca<sup>2+</sup> is taken into mitochondria, and (4) surplus Ca<sup>2+</sup> is extruded from the cytosol by plasma membrane Ca<sup>2+</sup>-ATPase 1 (PMCA1) and/or the NCX Na<sup>+</sup>/Ca<sup>2+</sup> exchanger. (<b>B</b>) (5) Ca<sup>2+</sup> depletion in the endoplasmic reticulum (ER) activates the Ca<sup>2+</sup> sensor STIM1, and (6) activated STIM1 binds to Orai1 in the plasma membrane, opening the channel for Ca<sup>2+</sup> entry. (7) Activated STIM1 can also associate with transient receptor C (TRPC) proteins, opening their channels for Ca<sup>2+</sup> entry. (8) Imported Ca<sup>2+</sup> is transported into the ER by sarcoendoplasmic reticulum Ca<sup>2+</sup> ATPase (SERCA) to replenish the ER stores.</p> ">
Review Reports Versions Notes

Abstract

:
From fertilisation to delivery, calcium must be transported into and within the foetoplacental unit for intracellular signalling. This requires very rapid, precisely located Ca2+ transfers. In addition, from around the eighth week of gestation, increasing amounts of calcium must be routed directly from maternal blood to the foetus for bone mineralisation through a flow-through system, which does not impact the intracellular Ca2+ concentration. These different processes are mediated by numerous membrane-sited Ca2+ channels, transporters, and exchangers. Understanding the mechanisms is essential to direct interventions to optimise foetal development and postnatal bone health and to protect the mother and foetus from pre-eclampsia. Ethical issues limit the availability of human foetal tissue for study. Our insight into the processes of placental Ca2+ handling is advancing rapidly, enabled by developing genetic, analytical, and computer technology. Because of their diverse sources, the reports of new findings are scattered. This review aims to pull the data together and to highlight areas of uncertainty. Areas needing clarification include trafficking, membrane expression, and recycling of channels and transporters in the placental microvilli; placental metabolism of vitamin D in gestational diabetes and pre-eclampsia; and the vascular effects of increased endothelial Orai expression by pregnancy-specific beta-1-glycoproteins PSG1 and PSG9.

1. Introduction

Three essential specifications for a system that delivers Ca2+ to the foetoplacental unit are as follows: First, it must be able to provide Ca2+ in a stringently titred amount for Ca2+ signalling in the placenta and foetus, starting at fertilisation and continuing to term. Second, it must be able to simultaneously provide a flow-through system to transfer increasing amounts of Ca2+ safely across the placenta to mineralise foetal bones without increasing the cytosolic Ca2+ to toxic levels. By approximately 35 weeks of gestation, around 300 mg of Ca2+ is transferred daily [1]. Third, it must match the supply to the changing and increasing needs of the foetus [2,3,4]. The source of Ca2 is the mother. Before around 16 weeks of gestation, Ca2+ is provided by uterine fluid, which contains secretions from decidual glands in the uterine endometrium [5,6]. The Ca2+ content of this fluid in humans is unknown; however, in sheep, it increased from 14 days postfertilisation (pf.) [7]. As the mature placenta forms and begins to function, the source is maternal blood bathing the foetal placental villi. With an average ionised Ca2+ concentration of approximately1.18 mM [8], this is around 11,800 times higher than the normal resting intracellular cytosolic concentration (<100 nmol/L) [9]. To meet these demands, the surface of the foetal placenta is covered by a very large syncytium of trophoblasts, covering around 5 m2 at 28 weeks of gestation and increasing to 11–12 m2 at term, which is folded into a myriad of branched free-floating villi [10,11]. Embedded in the villus membranes are a body of Ca2+ channels, which collectively can respond to a full range of stimuli from growth factors, cytokines, hormones, and mechanical stress [12,13]. Failure to meet the above requirements may result in ectopic pregnancy [14], growth restriction with risks to the foetus in utero and postnatally, early pre-eclampsia (PET) with risks to mother and foetus [3,15,16], and possibly long-term risks for osteoporosis in adult life [17,18].
We need to understand the mechanisms involved to intervene appropriately to reduce the risks from malfunction and to protect the bones of infants born extremely preterm and/or with birth weights < 1000 g [19,20]. Because of ethical issues, data on primary human tissues are scarce. Investigation requires a multi-pronged approach, drawing in data from human genetic disorders; an increasing range of animal models; physiological studies of trophoblast cell lines; tumour tissues; basic science; and, increasingly, gene expression studies using RNA sequencing (RNA-Seq) and chromatin immunoprecipitation-sequencing ChIP-Seq studies to investigate gene regulation. Extrapolation of findings from animal models may be problematic because of differences in placental structure and gene expression, notably between two commonly used animals, namely, sheep for physiological experiments [21,22,23] and mice for genetic manipulation [24,25]. Further, experimental procedures to collect samples for analysis influence the results [1]. RNA-Seq and microarrays only show genes expressed at a single time point and will differ according to epigenetic regulation. This accounts for some of the wide intra- and inter-individual variation and smaller racial differences observed [26,27]. Reported cell studies have used cell lines from choriocarcinomas, except for HTRB/SV neo, a human embryonic villus cell line transfected with simian virus 40 antigen. However, a new human trophoblast line has been produced [28], and others must follow. The use of 3D organoids that resemble the structure and physiology of the human placenta will enable studies of developmental events in human implantation and placentation [29,30,31,32,33].
Our insight into the mechanisms of placental Ca2+ transport has extended rapidly because of recent advances in analytical, genetic, and computer technology [34,35,36,37]. Coming from a variety of different sources, the published data are scattered. The aims here are to look for common threads and to highlight deficiencies in our knowledge. This may help to guide future research and policies for interventions to improve bone health. The items covered were selected because of their known involvement in Ca2+ turnover postnatally or because they have been shown experimentally to affect placental Ca2+ transport. There are four sections: Section A describes placental development, and Section B covers calcium carriers and transporters that control calcium. The transient receptor potential vanilloid 6 (TRPV6) channel is covered in depth. This has emerged as the major Ca2+ import channel in the placental villi, and its importance is manifest in human inherited TRPV6 deficiency. Section C covers the effects of peptides and hormones on placental calcium transport, and Section D is a brief overview of the developmental origins of health and disease (DOHaD) proposal in relation to postnatal bone development.
  • Section A: The Placenta: Laying the Foundation

2. Placental Development

A primitive placenta forms by approximately 10–12 weeks pf., but the villi have a low surface area and are poorly vascularised [38]. The definitive placenta has developed and functions by approximately 20 weeks pf. and grows exponentially until term [5,6]. The very early processes through which the embryo transfers from the uterine fallopian tube and becomes engulfed (implanted) within the endometrium of the body of the uterus are of paramount importance for normal development of the placenta. Figure 1 shows some key intermediates [6,39,40,41]. These are summarised briefly in Table 1, which also indicates the likely sources of calcium at each stage. For detailed information, refer to the literature [5,6,41,42,43,44].
At term, the placenta is 15–20 cm in diameter, weighs around 500 g, and normally has a centrally placed umbilical cord [45]. Septa arising from the basal layer of the uterine decidua grow into the foetal placenta, dividing it into 15–20 cotyledons, which each contain an anchoring (stem) villus and the villi that branch out from it—the intermediate and terminal villi. The highly branched terminal villi are the placental workforce (Figure 2).
Spiral branches of the maternal uterine arteries empty into the intervillous spaces. Blood percolates around the villi and then leaves via uterine veins. Two foetal arteries carry blood from the villi to the foetus via the umbilical cord, where they lie alongside a single vein carrying blood from the foetus to the maternal venous system. There is no continuity between maternal and foetal circulations in an undamaged placenta [40,41].

Terminal Villi: Actin, Ezrin, and NHERFI/EBP50

At term, the terminal villi are mobile, branched, and dissimilar from the rigid, unbranched microvilli of the duodenal mucosa. They have a core of bundled actin filaments, which are cross-linked and extend only a short distance into the cell cytoplasm. They are attached to the plasma membrane directly or indirectly, in part by ezrin [46,47,48,49]. At low cytoplasmic Ca2 concentrations, a villus protein, namely, α-actinin, probably crosslinks the actin fibres, reinforcing the cytoskeleton. At Ca2+ concentrations > 0.3 μM, the release of α-actinin would weaken the structure [48]. Two abundant villus proteins, namely, ezrin and Na+/H+ exchange regulatory cofactor 1 (NHERF1), interact and bind with actin. Figure 3 is a hypothetical model of the terminal villus core based on the better-defined structure of duodenal villi [50].
Ezrin, belonging to the ezrin, radixin, and myosin (ERM) protein family, accounted for approximately 5% of the total mass of protein isolated from placental syncytiotrophoblast and was present mainly as noncovalent dimers and higher-order oligomers [51]. ERMs regulate microvillus formation in tissue culture epithelial cells. Cells lacking ERMs have reduced numbers of microvilli [52]. Ezrin contains an NH2-terminal (N) domain of around 300 residues and a 100-residue COOH-terminal (C) domain. In a dormant form in the cytoplasm, the C-terminal tail binds to the N-terminal FERM domain and closes the molecule [53,54]. Phosphorylation of critical residues at the interface between the N- and C-terminal regions blocks this association and stabilises ezrin in an open state. The freed C-terminal binds to actin [55], and the N-terminal is directed towards ERM binding proteins associated with the apical membranes of the microvilli.
A search for ERM binding partners in placental microvilli identified a protein named as ezrin binding protein 50 kDa (EBP50) [56]. This colocalised with actin and ezrin and specifically associated with the microvilli of the placental syncytiotrophoblast. EBP50 was also found in cultured JEG-3 human choriocarcinoma cells [51,57]. Because EBP50 has two postsynaptic density 95/discs large/zonula occludens-1 (PDZ)-binding domains, which associate with integral membrane proteins, it was proposed that EBP50 might mediate the membrane attachment of ezrin. The microvilli of EBP50-null mice are short and abnormal, like those of ezrin-null mice, suggesting that the two proteins function together in microvillus structure or regulation [58,59]. However, this possibility does not appear to have been investigated. Independently, others identified the same protein through its involvement with regulation of the rabbit renal brush border Na+/H+ ion exchanger, NHE3 [60,61]. This protein was named NHERF1, and EBP50 has been renamed. NHERF1 has two NH2-terminal PDZ domains, PDZ1 and PDZ2, and a carboxyterminal ERM binding domain (Figure 4) [62,63]. The ERM binding site and a cholesterol-binding site in PDZ1 promote its close association with the villus membrane.
NHERF1 (solute carrier family 9 member A3 regulator 1, SLC9A3 regulator 1, gene NHERF1) is a cytoplasmic multifunctional scaffolding protein. It scaffolds membrane-bound proteins to the sub-apical actin cytoskeleton and stabilises them at the cell surface [62,63,64]. However, its main role is to connect membrane-associated proteins with transiently assembled cytosolic complexes, including kinases, phosphatases, and trafficking proteins, to direct cell signalling or transport activities [56,62]. Four members of the NHERF family are expressed in the kidneys (NHERF1-4). Whether the placenta expresses all forms is unknown. Amongst numerous membrane proteins bound by NHERF1 are the sodium–phosphate cotransporters NaPi-2a and NaPi-2c, and NHE3, which are expressed in placental microvilli. NHERF1 is phosphorylated by several kinases, which alter its binding activity and downstream signalling events [65,66]. Its role in the placenta is unknown.
  • Section B: Tools for Controlling intracellular Ca2+: Ca2+ Channels and Transporters

3. Sources and Removal of Cytosolic Ca2+

Ca2+ enters the cytosol through two routes: (i) from the extracellular space via channels or transporters in the plasma membrane, (Section 4) or (ii) by release from stores in the endoplasmic reticulum (ER). Rapid increases in Ca2+ trigger signalling, which is then cascaded through the cell along specified pathways to generate an appropriate response [67]. Surplus Ca2+ must be cleared rapidly to avoid a global increase in cytosolic Ca2+ and widespread uncontrolled stimulation of signalling (Section 6). A collection of Ca2+ entry channels is expressed variably on placental membranes across gestation. These include members of the transient receptor potential (TRP) family [12,68]. Current evidence indicates that the TRPV6 channel is the major channel mediating the vast influx of Ca2+ through terminal microvilli in the third trimester (Section 4.2). Surges of incoming Ca2+ through plasma membrane channels can increase the Ca2+ concentration across the cell to 0.5 to 1 μM, but concentrations may exceed 100 μM at the channel mouth [67]. It is essential that this is directed directly to the relevant signalling complex assembled close to the channel and that channel opening and closure are tightly regulated to respond promptly to changes in stimulation. Intracellular Ca2+ release from ER stores is mediated by inositol trisphosphate (IP3), generally generated by phospholipase (PLC) activation in response to stimulation of receptors in the plasma membrane (Section 3.1).

3.1. Phospholipase C

The ER acts as an intracellular store of Ca2+, which is readily available for a rapid signalling response to stimulation of membrane receptors. ER Ca2+ concentration in resting cells is normally around 400 μM [69], ranging from 200 μM to 650 μM [9]. Ligand binding to G-protein-coupled receptors, tyrosine kinase receptors, or other plasma membrane receptors activates phospholipase C (PLC) [70]. There are six PLC families, β, γ, δ, ε, η, and ζ, which all have four Ca2+-binding EF hands and catalytic domains. They are all targeted to membrane-bound phosphatidyl inositol 4,5 bisphosphate (PI (4,5) P2, alias PIP2). Except for PLCζ, this is via an amino-terminal pleckstrin homology (PH) domain [70,71,72,73]. PLC cleaves PIP2, releasing IP3 (Figure 5).
IP3 activates the inositol trisphosphate receptor (IP3R) on the surface of the ER, leading to release of Ca2+ into the cytosol. There are three IP3R isoforms, denoted IP3R1–3, that function as tetramers. Each monomer has a cytosolic amino-terminal domain that binds IP3; a regulatory domain that binds Ca2+, ATP, and other modulatory molecules/proteins; and a carboxy-terminal channel that contains six transmembrane domains and a short cytosolic tail. The activation and opening of the IP3R require binding by both Ca2+ and IP3. Dual regulation enables the channel to support long-lasting Ca2+ oscillations [73].

3.2. PLCζ (Zeta)

PLCζ is only expressed in sperm heads in mammals. In humans, PLCζ localises to three distinct regions [74,75,76,77]. Following the fusion of sperm with the egg, PLCζ releases IP3 from PIP2. This mediates Ca2+ release from the ER, causing oscillations in cytosolic Ca2+ that activate the ovum via numerous pathways, concluding in cortical granule exocytosis, resumption of meiosis II, and pronuclear formation. Cortical granule exocytosis, mediated principally by protein kinase C (PKC), releases enzymes that modify the zona pellucida and prevent further sperm entry. PLCζ is the only protein known so far that initiates Ca2+ oscillations during human fertilisation [73]. Ca2+-bound Ca2+-calmodulin-stimulated protein kinase II (CAMKII) ensures proper cell cycle progression. In the absence of Ca2+ at fertilisation, embryos consistently display a reduced inner cell mass at the blastocyst stage [78]. Oocyte activation deficiency (OAD) is the basis of total fertilisation failure (TFF) and is attributed to mutations in the PLCζ gene, termed male factor infertility [70,79].

4. Ca2+ Importation Across the Plasma Membrane

4.1. Transient Receptor Potential Channels (TRPs)

TRPs are widely distributed ion channels, which are permeable to monovalent and divalent cations, including Ca2+, Mg2+, Na+, and K+, with highly variable Ca2+/Na+ permeability ratios [67,80,81,82,83]. They act as sensors of chemically toxic and physical stimuli [84]. Mammals express 28 TRP channels that compose the TRP superfamily. There are six subfamilies based on amino acid sequence homology [85], designated as canonical (TRPC), vanilloid (TRPV), melastatin (TRPM), ankyrin (TRPA), mucolipin (TRPML), and polycystin (TRPP). TRPCs fall into four subsets: TRPC1, TRPC2, TRPC3/6/7, and TRPC4/5 [83,86]. TRPC2 is present in rodents but is a pseudogene in humans [87]. TRPV channels were named for their sensitivity to vanilloid and capsaicin [88,89]. TRPV5 and TRPV6 differ from all other TRP channels because of their high Ca2+ selectivity (permeability of Ca2+/Na+ > 100) [90,91,92,93]. TRPM6 and TRPM7 are permeable to both Ca2+ and Mg2+, but their main role may be to regulate Mg2+ import because they are sensitive to physiological Mg2+-ATP concentrations. Unlike the other TRPs, they have a kinase domain [94,95]. Although TRPV5/6 are reported to be co-expressed in the human placenta [96,97,98], expression of TRPV5 is negligible and 1000-fold lower than that of TRPV6 [82,99]. TRPV6 is expressed in the human placenta at term [16,100], in human trophoblasts and syncytiotrophoblasts [101], and in the mouse placenta [102].
TRPs assemble as homotetramers or frequently heterotetramers to form a cation-conducting pore [12,80,83,103,104]. Each monomer has six transmembrane domains (S1–S6) and intracellular NH2 and COOH (C) termini of variable length. Members of the TRPC, TRPM, and TRPV subfamilies have a conserved TRP domain comprising approximately 25 amino acids in the C-terminal. The function of the TRP domain is uncertain, but it may be involved in PIP2 binding [105] or subunit assembly [106]. Multiple regulatory elements, interaction sites, and enzymatic domains are present in the intracellular termini [80]. Diacylglycerol (DAG) is a direct activator of TRPC3, TRPC6, and TRPC7 channels [83,107]. Ankyrin repeat domains are present on the NH2 terminus of TRPC, TRPV, and TRPA subunits. The COOH terminus of TRPC channels has calmodulin (CaM)/IP3R-binding (CIRB) domains; Ca2+-binding EF hands; and in the case of TRPC4, a PDZ domain [81]. TRPs regulate many signalling pathways, including the mitogen-activated protein kinase (MAPK), transforming growth factor (TGF)-β, nuclear factor kappa-B (NF-κB), and AMP-activated protein kinase (AMPK) pathways. TRP channel activation induces multiple immunomodulatory effects [108,109].

4.2. Transient Receptor Potential Vanilloid 6 (TRPV6)

TRPV6 (alias CAT1) was identified in rat duodenal cells and shown to mediate intestinal Ca2+ absorption [110,111]. A related channel, TRPV5 (alias CATII), is expressed predominantly in the kidney and mediates trans-cellular Ca2+ reabsorption by a well-defined pathway [112]. Its primary function is to regulate urinary Ca2+ excretion [113,114]. They share 75% sequence homology and, unique for TRPs, have very high selectivity for Ca2+ ions. Unlike TRPV5, TRPV6 is expressed widely in normal epithelial tissues. Vitamin D, oestrogen, and dietary Ca2+ regulate the abundance of TRPV6 in the duodenum [115]. There is mounting evidence that TRPV6 expression is increased, or suppressed, by epigenetic mechanisms [116,117]. Two initiation sites for TRPV6 transcription have been identified: (i) a classic AUG methionine start codon and (ii) an unusual, non-canonical ACG codon, located 120 base pairs upstream of the AUG start site [118], which is decoded as methionine and not threonine as predicted. This produces a protein transcript with a 40-amino-acid extension on the N terminus, which may be important for attachment to the plasma membrane. TRPV6 expression in the placenta is driven largely by oestrogens [111,115]. TRPV6 is expressed by syncytiotrophoblasts from human term placenta [111,118,119,120,121,122]. Unlike the duodenum [116], 1,25(OH)2D3 did not increase TRPV6 expression significantly [100]. In mice, trpv6 was expressed at E10 pf. and increased by 14-fold during the last 4 days of gestation [123]. In rats, expression increased steadily and peaked at E20.5 pf. Expression was found to be progesterone- and oestrogen-receptor dependent [124] and increased in hypoxia [125]. Trpv6 mRNA and protein were also expressed in mouse bone, with a gradual increase from E9.5 pf. to E15.5–17.5 pf. but a marked decrease at parturition [123]. Unlike human and rodent placentas, in ungulates the cotyledons (referred to as placentomes) are not aggregated but are implanted separately in the endometrium and are connected by foetal membranes [126]. Trpv6 mRNA was found to increase gradually in the placentomes of cows through gestation but without an observable increase in proteins. In contrast, there was a large increase in both mRNA and protein of TRPV6 in the membranes, evidence that maternofoetal Ca2+ transport mainly occurs in intraplacentomal regions [126]. It seems that the processes of Ca2+ transfer for signalling in the placenta and provision for foetal bones are largely segregated in cows. Kogel, Fecher-Trost, Wissenbach, et al. (2022) [127] observed that in living cells, TRPV6 had a very short residence time in the plasma membrane and was barely discernible. This was explained by clathrin-mediated internalisation triggered by locally released Ca2+. TRPV6 was demonstrated in Rab7-positive endosomes targeted for recycling to the membrane or (most) in Rab11a-positive late endosomes for lysosomal degradation.

4.2.1. Structure and Operation

TRPV6 channels are constitutively open, allowing electro-chemical gradient-driven Ca2+ entry into the cell [128,129], but they are inactivated reversibly by Ca2+ to protect from a toxic influx. The Km value for Ca2+ uptake was 0.25 mmol/L for human TRPV6 [130]. Mg2+, gadolinium Gd3+, ruthenium red, low extracellular pH, 2-aminoethoxydiphenyl borate (2-APB), and soricidin derivatives decrease flow through the channel [82,115,131]. Like the other TRPs, the TRPV6 channel is tetrameric, with the pore located centrally and formed by helix 6 of all four monomers, as described in detail in the literature [132,133,134]. Figure 6 illustrates the important domains. Features of relevance to channel closure are the intracellular “skirt” formed by the interlocking N and C termini, which encloses a 50 Å × 50 Å wide cavity below the pore of the ion channel; the molecular cage at the pore orifice formed by four TRPV6 W583 residues, one from each monomer; and binding sites for Ca2+-CaM in the C-terminal tails.
Closure of the channel by Ca2+ inactivation has two components. An initial fast inactivation is attributed to amino acid sequences in the intracellular loop between trans-membrane helices H2 and H3 [135] and in the helix–loop–helix (HLH) domain of the TRPV6 N-terminal [136]. A slower process involves Ca2+/CaM binding to the TRPV6 C-terminal. CaM is a monomeric protein comprising two structurally similar globular lobes at the N- and C-terminals connected by a linker helix. Each lobe has two Ca2+-binding EF hands. The C-lobe binds Ca2+ more tightly (Kd around 0.2 μM) than the N-lobe (Kd around 2 μM) [137,138,139]. The N-lobe only becomes Ca2+ loaded when the intracellular Ca2+ increases significantly above the resting concentrations (upper limit approximately100 nM [140]), and hence, it can act as a Ca2+ sensor. In the absence of Ca2+, apo-CaM adopts a closed conformation. When fully loaded, the inter-lobe linker is extended so that the molecule opens and becomes flexible, and the N- and C-lobes are distanced [140,141,142]. The lobes bind at separate sites on the TRPV6 carboxy terminal. This causes the CaM C-lobe to rotate towards the channel pore, and CaM K115 engages with the W583 molecular cage at the pore orifice. This interaction not only blocks the channel exit but also initiates a sequence of intramolecular events that prevent Ca2+ entry into the TRPV6 channel [115,140,142,143,144,145].

4.2.2. Associated Proteins

In addition to CaM, other proteins that bind or associate with TRPV6 are S100 calcium-binding protein A10 (S100A10) [142,146], Na+/H+ exchange regulatory cofactor 4 (NHERF4) [147], Ras-related protein Rab-11A (Rab11a) [148], TRPM8 channel-associated factor 2 TCAF2 [149], cyclophylin B [150,151], klotho [98], and TRPC1 [152,153]. S100A10 increases expression of TRPV6 at the cell surface (Section 4.2.4). Rab 11a is required for endosomal membrane fusion and may have a role in recycling TRPV6 at the cell surface [154]. Cyclophilin B associates with TRPV6 in human syncytiotrophoblasts and increases TRPV6 activity in vitro [82]. It regulates the activation of interferon-regulatory factor 3 [IRF3]. Klotho has low expression in the placenta. TRPC1 and TRPV6 interact via their N-terminal ankyrin repeats, resulting in downregulation of TRPV6 expression in the plasma membrane and decreased TRPV6 Ca2+ currents, perhaps by retaining TRPV6 in intracellular compartments [153].

4.2.3. TRPV6 Deficiency in Pregnancy

Trpv6 deficiency occurs in human foetuses with inherited mutations of the TRPV6 genes and has also been associated with pre-eclampsia (foetal and placental abnormalities were observed in mouse models with inactivating mutations of the trpv6 gene).
Inherited TRPV6 deficiency, namely, transient hyperparathyroidism of the newborn, OMIM 618188, has been reported in 12 individuals [115,155,156,157,158,159]. Inheritance was autosomal recessive, and there was similar sex distribution. Most were born at term, and five had low birth weight. Eleven had a gross disturbance of bone turnover (bone dysplasia), which was evident in utero in the third trimester. At birth, this manifests as a narrow, funnel-/bell-shaped chest; soft abnormal ribs with fractures; short, sometimes bowed, femora; and a generalised deficiency of mineralised bone (osteopenia). There were no other physical abnormalities. The phenotype was normal in the twelfth case. The chest restriction caused severe life-threatening respiratory problems and feeding difficulties requiring intensive support, often for several months. At birth, five babies had a low serum-ionised or corrected calcium. All had raised parathyroid hormone (PTH) levels. The serum 25(OH)D of seven newborns indicated vitamin D insufficiency. Remarkably, after the correction of serum Ca2+, the skeletal abnormalities resolved over months, with normal oral intakes of calcium and vitamin D supplements for infancy [155,156,157,158,159,160]. Most mutations were inactivating mutations in ankyrin repeats, transmembrane helices or their linkers, or the C-terminal hook but not in the channel pore. However, two were gain of function mutations located at fast Ca2+ inactivation sites in TRPV6 [135,136].
Additional observations relevant to the pathogenesis of this defect were as follows: (i) A bone abnormality was detected at 20 weeks of gestation in one foetus, which progressed during the third trimester [155]. This early manifestation might be explained by local effects of bone TRPV6 deficiency, rather than poor placental Ca2+ provision. (ii) Four of the five affected newborns were Japanese. One contributory factor might have been a low maternal dietary Ca2+ intake [161,162]. The average Ca2+ intake in Japan has been estimated at approximately 480 mg/day, which is below the recommended intake for Asia of 700 mg/day [162]. The individual with a normal phenotype was Japanese and had the same mutations as an older severely affected sibling. It was suggested that the calcium status of the mother might have differed between pregnancies [158]. (iii) In one affected newborn, serum 25(OH)D indicated frank vitamin D deficiency, as did the concentrations of her mother and dizygous twin sibling. The twin had normal TRPV6 alleles. His serum Ca2+ was low, but he had no phenotypic abnormalities [157]. Hence, frank vitamin D deficiency alone did not cause overt bone abnormalities in utero. (iv) The placenta collected at term delivery of one of the affected foetuses [155] had a normal weight and macroscopic appearance [160]. However, compared with a control placenta from an unaffected pregnancy, 15 proteins were significantly increased, and 4 were decreased. Two proteases, HTRA1 (high-temperature requirement A serine peptidase 1) and cathepsin G, were only found in the affected placenta [163].
Pre-eclampsia is associated with foetal growth restriction [16]. Ca2+ transport by cultured primary syncytiotrophoblasts from placentas of women with pre-eclampsia was significantly lower than that of matched controls, as were mRNA and protein expression of genes encoding TRPV5, TRPV6, calbindin D9k, calbindin D28K, and plasma membrane Ca2+ ATPase-1 and -4 (PMCA1 and PMCA4). mRNA expression of Ip3R and the ryanodine receptor (Ryr) was also decreased, but mRNA of sarcoendoplasmic reticulum ATPase (SERCA 1, 2 and 3); the mitochondrial voltage-dependent anion channel 1 (VDAC); and the DNA repair gene 8-oxoguanine glycosylase (OGG1) were increased. The findings indicated a disturbance in placental Ca2+ homeostasis and transport, possibly attributable to ATP depletion and oxidative stress [16].
In mice, TRPV6 was inactivated by knock-down of the gene or by introducing a mutation at the TRPV6 channel pore [102,164]. TRPV6-deficient foetuses accumulated less Ca2+ [164]. Placentas from pregnancies in which both the mother and the foetus had homozygous mutations were structurally abnormal. The foetal labyrinth was thicker and less dense and had much larger cell-free spaces than in placentas from wild-type (WT) pregnancies [102]. Whereas WT-cultured placental trophoblasts connected with many adjacent cells and built a tight cellular network, placental cells from mutant animals had less contact with neighbouring cells, and their extracellular matrix (ECM) was less dense. Mutant trophoblasts had higher expression of some proteases, including HTRA1 and granzymes (cytotoxic serine proteases), and decreased fibronectin, a component of the ECM [164]. Zebrafish with a naturally occurring loss-of-function trpv6 mutation [R304Stop] had a 68% reduction in total body calcium content, bone mineral defects, and a marked reduction in Ca2+ uptake by the yolk sac and gills [165]. Zebrafish TRPV6 protein is expressed in yolk sac cells, and long stretches of the amino acid sequence are closely homologous with the human TRPV6 sequence [166].

4.2.4. TRPV6 Membrane Expression: S100A10 and Annexin A2

S100A10 (p11, annexin II light chain, calpactin light chain) is a multifunctional protein with a wide range of physiological activity. Unlike other S100 family members, it cannot bind Ca2+ but is locked in the Ca2+-loaded conformation. S100A10 interacts with several proteins but predominantly exists in a heterotetrameric complex with Annexin A2 (ANXA2) [114,167,168]. ANXA2, S100A10, and the heterotetramer were expressed in human syncytiotrophoblast, mainly in areas with active placentation [169]. Both ANXA2 and S100A10 were demonstrated in brush-border membrane vesicles from human placenta, and both increased progressively during gestation [170]. ANXA2 and S100A10 were among 115 genes in sequenced transcriptomes from the placentas of 14 mammalian species and were expressed in all and classed as core genes [25]. This suggests that they may have important roles.
Annexins are Ca2+-binding proteins that interact with biological membranes. The ANXA2 monomer is an intracellular protein with a core Ca2+- and phospholipid-binding domain and a small N-terminal tail domain [168,169]. The Ca2+-binding sites are distinct from EF-hand sites [171]. Ca2+ liganded by carboxyl and carbonyl oxygens located on the ANXA2 surface binds with negatively charged phospholipids, including PIP2 and phosphatidylserine with high affinity [172], and bridges ANXA2 to the cell membrane. In the presence of Ca2+, ANXA2 also binds to F-actin [168]. The first 14 N-terminal ANX2 residues bind with S100A10, forming a tight dimer, and two dimers combine to form a heterotetramer that interacts with actin and membrane phospholipids [114,168]. The S100A10 subunits face the cytosol and can interact with additional proteins, including ion channels such as TRPV6 [168]. The membrane binding function of ANXA2 in the complex may actuate endocytosis or exocytosis and/or the regulation of intracellular transport of various ion channels and transporters [167,168,169,173,174]. S100A10 associates with a conserved sequence in the C-terminal tails of TRPV5 and TRPV6 in a Ca2+-independent manner. The complex plays a crucial role in routing TRPV5 and TRPV6 to the plasma membrane. In the kidney, a significant subset of TRPV5 channels is localised subapically in the distal renal tubules, suggesting that TRPV channels may be shuttled to the plasma membrane [146,175]. Downregulation of ANXA2 inhibited TRPV5- and TRPV6-mediated currents in transfected HEK293 cells [114].

5. Store-Operated Ca2+ Entry (SOCE)

Recurrent stimulation of membrane receptors that activates PLC and generates IP3 drains Ca2+ from the ER stores (Section 3). This must be replenished from extracellular sources. A major route demonstrated in non-excitable tissues was via plasma membrane Ca2+ channels, which were activated by the depleted stores: store-operated Ca2+ entry (SOCE) channels, also termed capacitive Ca2+ entry channels [176,177]. These were characterised electrophysiologically and named the Ca2+ release-activated Ca2+ (CRAC) channel, but the protein responsible was unknown. They had a high selectivity for Ca2+ versus Na+, displayed an inwardly rectifying current–voltage relationship, and responded to a decrease in ER Ca2+ [69]. Clarson, Roberts, Hamark, et al. (2003) [178] demonstrated for the first time that SOCE occurred in term human placenta. Ca2+ depletion of term villi by incubation in a Ca2+-free buffer with thapsigargin to inhibit Ca uptake into the ER, followed by superfusion with Ca2+-containing buffer, resulted in a rapid rise in [Ca2+]i. This was inhibited by GdCl3, NiCl2, CoCl2, or the channel inhibitor MSKF96365 but not by nifedipine. This was not demonstrable in fragments of first-trimester placenta. mRNA encoding TRPC1, TRPC3, TRPC4, TRPC5, and TRPC6 was identified in both first-trimester and term placentas.
The human CRAC ortholog was identified in two independent studies using unbiased genome-wide RNA interference screens in Drosophila cells [179,180]. This was characterised as a plasma membrane–resident protein encoded by gene FLJ14466 [179]. Feske, Müller, Graf, et al. (1996) [181] found that T lymphocytes from two siblings with one form of hereditary severe combined immune deficiency (SCID) syndrome were defective in store-operated Ca2+ entry and CRAC channel function. This rare disorder can be caused by mutations in a variety of genes that lead to impaired function of T, B, or natural killer cells. Using a combination of a modified linkage analysis with single-nucleotide polymorphism arrays and a Drosophila RNA interference screen, they subsequently identified a point mutation in the CRAC gene, which they named Orai1 [182]. Later, mutations of stromal interaction molecule 1 (STIM1) (Section 5.2) were identified in two siblings with SCID with defective store-operated Ca2+ entry [183]. The CRAC channels mediate SOCE through the partnership of Orai and STIM1 [184,185,186]. Several non-selective Ca2+-permeable TRPC channels (Section 5) have also been proposed to act as store-operated channels [185,186,187,188], although this is controversial [69]. TRPC1 was thought to be incorporated into a combined “ISOC” channel with Orai1 and STIM1. Although both TRPC1 and Orai1 interact with STIM1, there are no data to show direct interaction of TRPC1 with Orai1. The available evidence indicates that TRPC1 and Orai1 are distinct channels [185,186].

5.1. CRAC/Orai1 Entry Channel

Human Orai1 (CRAC) channels are around 55 Å long and narrow in the closed conformation [184]. They assemble and function as hexamers. Each subunit comprises four transmembrane helices, (TM1, TM2, TM3, and TM4). The hexameric channel contains a single ion conduction pore lined by the TM1 helices. The TM4 domains form an outer coat that interacts with the plasma membrane and STIM [189]. Incoming Ca2+ ions are concentrated by a ring of acidic E106 side chains. Below this, two rings of residues, V102 and F99, form a hydrophobic gate. Activation of the channel by STIM1 causes the pore helices to rotate, moving the F99 residues away from the pore axis and enabling Ca2+ ions to flow through [189].
The Orai1 channel is highly Ca2+-selective with a Ca2+:Na+ permeability ratio of >1000, but in the closed state, the unitary Ca2+ conductance is extremely low, estimated at around 10–35 fS [190]. Activation of the channels is not voltage-dependent, and they are insensitive to most of the common voltage Ca2+ channel inhibitors. Extracellular Ca2+ potentiates channel activity by several fold [69], and associated Orai-binding proteins including CaM may fine tune activity by unknown mechanisms [69]. CRAC activity is modulated directly by lipids and posttranslational modifications or indirectly by accessory proteins at the ER–plasma membrane contact site. The channels undergo fast Ca2+-dependent inactivation (milliseconds) due to binding of Ca2+ close to the intracellular face of the pore [69]. Over a longer timescale (seconds to tens of seconds), intracellular Ca2+ accumulation causes slow inactivation [69]. Refer to the literature [191,192,193,194] for more detail.

5.2. Stromal Interaction Molecule 1 and 2 (STIM1 and STIM2)

STIM1 and STIM2 are single-pass ER membrane proteins with >74% sequence similarity (54% sequence identity). Their functions are to sense the ER Ca2+ concentration and to respond to Ca2+ depletion by activating Orai1 and opening the channel to allow a large influx of Ca2+ into the cell. STIM1 and STIM2 work in concert and can form heterooligomers. STIM2 has a lower affinity for Ca2+ than STIM1 and is therefore more sensitive to small changes in ER Ca2+ concentration and can initiate an earlier response (reviewed [69,186,191,195]). STIM1 and STIM2 have a luminal NH2 terminus and a cytoplasmic COOH terminus, are glycosylated, and have a series of structural modules with defined functions [69,186,191,196,197]. Figure 7 shows the main domains for STIM1 and their roles.
In the N-terminal within the ER, there are two Ca2+-binding EF-hand domains, EF1 and EF2, and a sterile alpha motif (SAM), which is essential for the dimerization/multimerization of the STIM monomers and activation of the protein [191]. The STIM-Orai activating region (named SOAR) is located at two of the three highly conserved coiled-coil (CC) domains, CC2/CC3, on the cytosolic side of the ER membrane [69,189,198]. Positively charged lysine residues in the polybasic domain (PBD) at the C-terminal interact electrostatically with acidic phospholipids and acyl chains in the plasma membrane, including PIP2, and localise the activated STIM proteins at the ER–plasma membrane (PM) [186].
In the resting state, the two monomeric EF-hand domains bind five or six free Ca2+ atoms within the ER [189]. The STIM monomers are not associated, and their N-terminals float freely. The CC1 coil binds with the CC2/CC3 SOAR domain, blocking the access of the STIM to Orai1 and the attachment of the PBD to the plasma membrane. When ER Ca2+ is depleted and the EF-binding sites are unoccupied, there is a dramatic change in the STIM conformation, initiated by the dimerization/polymerization of the STIM monomers, followed by rotation of the multimers in the ER membrane [191]. The SOAR domain is released from CC1, the bundled STIM multimers spread out, and their C-terminal PBDs associate with the plasma membrane at the ER/PM junctions. Orai1 channels are drawn closer to the STIM, facilitating interaction of the N-terminal of Orai with the SOAR domain of the STIM. This promotes rotation of the Orai channel helices, which opens the channel pore for Ca2+ entry (Figure 8) [69,186,191].
There are many predicted/putative Ca2+/CaM-binding sites in the cytoplasmic domains of both STIM1 and STIM2 [199]. The lysine-rich PBD binds CaM with very high affinity in the presence of Ca2+ (Kd: 0.8 μM for STIM1 and 0.9 μM for STIM2). In its absence, the affinity is lower (Kd: 55 μM and 150 μM for STIM1 and STIM2, respectively) [200]. These levels indicate that cytosolic Ca2+ may modulate the STIM–CaM interaction. One proposal was that CaM might destabilise the ER–PM contacts at high Ca2+ concentrations by competing with phosphoinositides for binding to the PBD.
STIM1 has been shown to activate TRPC channels but through a mechanism not requiring the STIM1 N terminus or transmembrane domain. It was proposed that the C–C domains in the STIM1 C terminus bind to TRPC1, 4, and 5 and that the terminal PBD region gates the TRPC channels [201,202].

5.3. SOCE in the Placenta

In the Protein Atlas (https://www.proteinatlas.org/, accessed on 29 December 2024) seven entries reported expression of small amounts of mRNA of Orai1 and eight of STIM1. These genes or their proteins were not highlighted in array studies of placental villi, syncytiotrophoblasts, or trophoblasts [24,25,26,100,164]. From this lack of data, it seems that Orai and STIM are not produced by the brush-border syncytium. This requires confirmation. If true, ER stores must be replenished with extracellular Ca2+ imported by TRP Ca2+ channels, possibly with a minor contribution by voltage-gated channels. TRPC3, TRPC4, and TRPC6 mRNA and proteins were expressed in the syncytiotrophoblasts of human term placentas, and mRNA of these channels and of TRPC1 and TRPC6 was detected in first-trimester placentas [178].
Two studies demonstrated that pregnancy-specific beta-1-glycoproteins from the immunoglobulin superfamily, PSG1 [203] and PSG9 [204], increased production of Orai in foetal cell cultures. Both PSGs are secreted by syncytiotrophoblasts into the maternal circulation, and serum concentrations increase with gestational age. PSG1 may have a role in placental vascular development [205]. PSG9 has important roles in immune regulation, thrombosis regulation, and angiogenesis during pregnancy [204]. Compared with concentrations in healthy pregnant women, PSG1 levels were lower in the serum of individuals with early-onset pre-eclampsia (EOPE) [203,206], and serum PSG9 levels were significantly decreased in patients with pre-eclampsia [204].
PSG1 treatment of cultured human trophoblast HTR-8/SVneo cells upregulated the expression of Orai1 protein and phosphorylation of Akt triggered by Ca2+ entry through the Orai1 channel. The selective inhibitor of Orai1 (MRS1845) suppressed Akt signalling and decreased the migration of trophoblasts in response to PSG1. It was proposed that downregulated PSG1 may reduce the Orai1/Akt signalling pathway, thereby inhibiting trophoblast migration, and suggested that PSG1 may serve as a potential target for the treatment and diagnosis of EOPE. It is probable that this was a response of stromal cells to PSG1 [28]. HTR-8/SVneo cells are first-trimester extravillous trophoblasts, which were infected with simian virus 40 large T antigen (SV40). Cultured HTR-8/SVneo cells were shown to generate a predominance of stromal cells and relatively few trophoblasts, whereas choriocarcinoma cell lines (BeWo, JEG-3, and Jar) generated only trophoblasts [207]. Further investigation using trophoblast cell lines derived from human stem cells that develop into extravillous trophoblasts may help to resolve this issue [29].
PSG9 treatment (0.1 µg/mL) of human umbilical vein endothelial cells (HUVECs) significantly enhanced the expression levels of Orai1 and Orai2 and store-operated calcium entry (SOCE) and the expression of endothelial nitric oxide synthase (eNOS) and NO production. Pretreatment with an inhibitor of SOCE (3,5-bis(trifluoromethyl) pyrazole derivative, BTP2) abolished these responses to PSG9. eNOS is synthesised primarily by endothelial cells and regulates vascular tension by promoting NO synthesis. Ca2+ bound to calmodulin increases eNOS activity [208]. The study indicates that PSG9 may regulate the function of vascular endothelial cells by increasing Ca2+ entry via the Orai1 channel. The same mechanism may apply to PSG1. These findings may be relevant to pre-eclampsia, could have therapeutic potential, and should be pursued.

6. Ca2+ Clearance from the Cytosol

Excess Ca2+ may be removed from the cytosol by transfer to the ER, mitochondria, or extracellular space, as shown in Figure 9A.

6.1. Sarcoplasmic/Endoplasmic Reticulum Ca 2+-ATPase, (SERCA)

SERCA is a membrane transport protein found ubiquitously in the ER of all eukaryotic cells. Its major function is to transport Ca2+ from the cytosol into the ER to restore basal Ca2+ concentrations following oscillations to μM concentrations elicited by cell stimulation [209]. There are three members of the SERCA family (SERCA1–3). The genes are ATP2a1/2/3. mRNA of ATP2a2 and ATP2a3 was expressed in mouse trophoblasts at E14.5 pf., but ATP2a1 was not detectable [102]. SERCA is expressed in oocytes [210,211,212]. However, only a fraction of Ca2+ from each Ca2+ oscillation is recycled into the ER Ca2+ pool by SERCA. A small amount of Ca2+ is transported into mitochondria by the electrophoretic pump [209], but most Ca2+ is extruded from the cytosol into the extracellular space via two plasma membrane transporters: plasma membrane calcium ATPase (PMCA) and Na+/Ca2+ exchanger (NCX).

6.2. Plasma Membrane Calcium ATPase (PMCA)

In mammals, four separate genes code for the major PMCA isoforms 1–4 (genes Atp2b1–Atp2b4) [213]. Differentiated trophoblasts from human term placenta expressed mRNA of PMCA 1–4 [214]. Mouse trophoblast homogenates expressed mRNA for atpb1 (PMCA1) and atpb4 (PMCA4) at E14.5 pf. but very little atpb2 or atpb3 [102]. Homozygous atp2b1−/− mice die in early embryonic life [215]. PMCA has 10 transmembrane domains, two large intracellular loops, and amino- and carboxy-terminal cytoplasmic tails. The carboxy tail contains the main regulatory sites for the activity of the pump: a CaM-binding domain, phosphorylation sites for protein kinases A (PKA) and C (PKC), and a high-affinity allosteric Ca2+-binding site. The carboxy terminal of full-length splice isoforms interacts with the PDZ domains of a variety of proteins, notably NHERF2 [209,213]. PMCA operates as a Ca2+:H+ exchanger with a 1:1 Ca2+/ATP stoichiometry. It has a high Ca2+ affinity, and many agents modulate its activity. The first loop contains sites for activation by phospholipids and autoinhibitory interaction with the CaM-binding domain. The second loop includes the binding sites for ATP, the acyl phosphate intermediate, and the second binding site for the carboxy-terminal CAM-binding domain [209].
The main regulator of PMCA function is Ca2+/CaM [216,217]. In the absence of CaM, the pumps are autoinhibited by the C-terminal tail, which binds to the two major intracellular loops. The release of autoinhibition requires binding of Ca2+/CaM to the PMCA C-terminal. This induces a conformational change in PMCA, with displacement of the C-terminal tail from the major catalytic domain [214,216]. Acidic phospholipids or phosphorylation of Ser/Thr residues in the C-terminal by PKC and/or PKA may facilitate these events. Inhibition of CaM prevents PMCA stimulation [106]. PMCA2b interacts with NHERF2 and may link the transporter to the underlying actin cytoskeleton and stabilise its local retention [218]. Whether PMCA1 interacts with NHERF1 is unknown.

6.3. Sodium–Calcium Exchangers (NCX)

Mammals express three NCX isoforms (NCX1, NXC2, and NXC3), the genes being SLC8A1, SLC8A2, and SLC8A3, respectively. They are low-affinity, high-capacity Na+/Ca2+ antiporters sited in the plasma membrane, with a central role in maintaining cellular calcium homeostasis for cell signalling. NCX1 is expressed widely, predominantly in the heart, kidney, and brain. NXC2 and NXC3 have a more restricted distribution [219,220,221].
NCX has a transmembrane (TM) domain with 10 TM helices and a large intracellular loop that separates the TM domain into two halves containing TMs 1–5 and TMs 6–10, respectively, and short extracellular N and intracellular C termini [219,222,223,224]. The TM domain mediates ion exchange, and the intracellular loop controls allosteric regulation by cytosolic Ca2+ and Na+ [219,225]. NCX catalyses the exchange of Na+ and Ca2+ with a 3:1 stoichiometry and is electrogenic [209,219]. The intracellular loop contains two Ca2+-binding domains (CBDs). CBD1 is the primary Ca2+ sensor. It detects small increases in cytosolic Ca2+ and undergoes large structural changes that activate the exchanger. CBD2 only binds Ca2+ at high concentrations and undergoes modest structural alterations [209]. Many other agents participate in regulation [220]. NCX1 is expressed in human placental tissue [226], and both NCX1and NCX2 isoforms in trophoblasts from human term placenta and in BeWo (choriocarcinoma) cells [214,227]. Under basal conditions, NCX does not have a major role in placental Ca2+ efflux [214].

7. Transcellular Calcium Transport to the Foetus

As the foetal organs grow and the skeleton develops, there is an increasing transcellular flow of Ca2+ across the placenta from the mother to the foetus, which rises exponentially in the third trimester. This must be accomplished without causing an increase in the syncytial cytosolic Ca2+ concentration and cell toxicity. It requires that Ca2+ is chaperoned throughout its passage through the placental villi and then rapidly discharged across the basal membrane to the extracellular space and foetal circulation. In contrast to the duodenum, Ca2+ is not absorbed additionally by the paracellular route [228,229]. In the duodenum, transcellular absorption is driven by 1,25(OH)2D and has been investigated intensively. Few data are available for the placenta.

7.1. The Working Model Proposed for Postnatal Intestinal Ca2+ Absorption

A working model [114,230] shows that Ca2 enters the microvillus through the TRPV6 calcium channel and is promptly taken up by CaM bound to brush-border myosin 1A (myo1A) [231]. Binding to myo1A may facilitate the movement of the Ca2+/CaM complex into the terminal web (shown in Figure 3). Here, Ca2+ transfers to calbindin-D9K (CaBP-9k) [232,233], which has a higher affinity for Ca2+ than CaM [234], and is transported through the cytoplasm in endocytic vesicles. At the basolateral membrane, Ca2+ is discharged from the cell by the Ca-ATPase PMCA1b. In support of the proposed model, the following can be noted: (i) CaM is the major Ca2+-binding protein in the microvillus [234]. Inhibitors of CaM block 1,25(OH)2D-stimulated calcium uptake by brush-border membrane vesicles (BBMV) [235]. (ii) The 1,25(OH)2D increases the concentration of CaM in the microvilli. (iii) CaM/myo1A is demonstrable at the highest concentration in intestinal brush-border membrane cells with high capacity to transport Ca2 [236]. (iv) The 1,25(OH)2D enhances intestinal calcium transport by inducing the expression of trpv6, cabp-9k, and pmca1b [114,126,220,230,237]. (v) In vitamin-D-deficient animals, Ca2+ microanalysis showed Ca2+ accumulation along the inner surface of the plasma membrane of microvilli [230]. (vi) Following vitamin D or 1,25(OH)2D administration, Ca2+ is seen in mitochondria and vesicles within the terminal web [230]. (vii) Numerous tubules and vesicles containing Ca2+ were seen in the terminal web region of enterocytes from normal and vitamin-D-treated chicks. These were smaller and depleted in vitamin D deficiency [238].
Two CaBPs, calbindin-D9K (CaBP-9k, gene S100G, locus Xp22.2) and calbindin-D28K (gene CALB1, locus 8q21.3), act as cytosolic Ca2+ buffers to maintain low intracellular Ca2+ levels during changes in transcellular Ca2+ transport. Both bind Ca2+ with high affinity and are increased by 1,25(OH)2D [114]. Calbindin-D9K is highly expressed in the small intestine and mediates Ca2+ transport [114,126,237]. In the small intestine, PMCa1b is the vitamin-D-regulated extrusion system and not NXC1 [230,239]. The deletion of pmca1b in mice reduces calcium absorption and growth and bone mineralization [240]. PMCA activation is dependent on CaM, which is its main regulator (Section 6.2) [114,216,240]. CaBP-9k may directly enhance PMCa1b activity [241]. Problems with the proposal are as follows: (i) Mice null for s100g have normal intestinal calcium transport, serum Ca2+ concentration, and bone mineralisation [242]. Another protein may act as a chaperone [243]. (ii) Ca2+ transport of myosin1-deficient mice is not reduced [244].

7.2. Transcellular Ca2+ Transport in the Placenta

The transcellular pathway for Ca2+ in the placenta similarly requires that Ca2+ is chaperoned in transit and is extruded rapidly. Data are limited. As in the duodenum, Ca2+ enters the villi of the human placenta via TRPV6 channels (Section 4), and PMCA1b has the principal role in extruding Ca2+, with a minimal contribution from NCX [214]. In rodents, pmca gene expression doubles during the last 7 days of gestation [233]. In humans, PMCA protein expression did not change during the third trimester of pregnancy, but the activity of the transporter increased linearly during this period [245], indicating activation by other agents such as CaM, PKC, and acidic phospholipids [246].
A major difference between placental and duodenal transcellular Ca2+ transport is that it is driven by oestrogens and not by 1,25(OH)2D [100]. In addition, there is uncertainty about whether CaBP-9k is the chaperone for placental transport in humans. There are species differences in calbindin expression, explained by complex regulation by sex steroids [247,248]. s100g was expressed in rat placenta and increased dramatically in late gestation [233]. Structural studies demonstrated a difference in two essential nucleotides in the oestrogen response element [ERE] between the rat and human genes. The human ERE failed to bind the oestrogen receptor. This could prevent CaBP-9k expression in the human uterus and possibly the placenta [247]. In the mouse uterus, s100g is mainly regulated by progesterone and not oestrogen, explained by a single-base difference in the mouse ERE compared with that of rats [248]. Nonetheless, mRNA and protein for CaBP-9k were expressed in human placental tissue collected at term [16]. RNA and protein of calbindin-D28K were expressed in cytotrophoblasts and syncytiotrophoblasts from human term placenta [16,249], and calbindin-D28K might potentially contribute to transport.
  • Section C: Effects of Peptides and Hormones on Placental Ca2+ Transport
The provision of an adequate mineral supply to the foetus is critically dependent on a normally formed and functional villous system. The foundations for this are laid very early in embryonic life at the time of implantation. Numerous factors are required for this process and continued placental development. Two factors that disturb calcium transport in animal models with genetic deficiencies are insulin-like growth factor 2 (IGF2) and parathyroid hormone-related peptide (PTHrP).

8. Insulin-like Growth Factor 2 (IGF2)

IGF2 is a growth factor expressed in the placenta in many species, including humans and rodents. The IGF2 gene has a complex genomic and transcriptional organisation with five promoters [3,250,251,252,253,254], which is normally expressed by only the paternal allele [250]. The P0 promoter operates specifically in the placenta, leading to high levels of IGF2 expression in placental tissues during gestation. It is abundant in foetal tissues and circulation and in the maternal circulation [3,4,250,255,256]. IGF2 is expressed from very early in embryonic life and increases progressively to term and has been shown to be crucially involved in blastocyst implantation and for normal foetal and placental growth throughout pregnancy. Foetoplacental endothelial cells are a significant source [4]. Levels increased progressively in human cord blood from the 21st week of gestation to delivery, with the main increase after the 32nd week [255]. IGF2 binds with high affinity to insulin-like growth-factor-binding proteins (IGFBPs). A metalloproteinase, pregnancy-associated plasma protein-A2 (PAPP-A2), cleaves bioactive IGF from IGFBP-3 and -5 [256,257]. Free IGF2 binds principally to the type 1 IGF receptor (IGF1R) with high affinity [258]. This is a tyrosine kinase receptor widely expressed in the placenta, which activates mitogen-activated protein kinase (MAPK) and PI3 kinase signalling pathways [3,250,253,254,259,260]. IGF2 also binds with IGF receptor 2 (IGF2R), which has been shown to mediate IGF2 clearance through lysosomal degradation, and it can bind to the type A insulin receptor INSR-A [250]. In vitro, IGF2 stimulates the survival, proliferation, and differentiation of human placental trophoblasts [3,253] and, in mouse studies, the expansion of the placental vasculature to support foetal growth [4].

Pathophysiology

Complete ablation of Igf2 in mice (Igf2 null) results in poor growth of the placenta, defective vascularization, increased barrier thickness and reduced surface area of the villous labyrinth, impaired amino acid transport capacity, and foetal growth restriction [261]. Deletion of the placental-specific Igf2 PO transcript in mice similarly causes placental growth restriction and compromised labyrinthine zone formation. However, adaptive upregulation of glucose, glutamine, system A amino acid, and Ca2+ transport [251] moderates the effects [254]. At embryonic day 17 (E17) pf., the foetal and placental weights of P0 knockout (P0) mice were reduced when compared with WT (wild type) but the ratios of foetal to placental weight were significantly increased. Despite having similar rates of placental Ca2+ transport, P0 foetuses had a lower capacity for Ca2+ transport, and blood and body Ca2+ and bone mineralisation were reduced. By E19 pf., placental Ca2+ transport and bone mineralisation had increased, demonstrating an adaptive response [251]. Despite this, mRNA expression of placental Ca2+ pathway genes was generally reduced compared with WT [262].
The importance of IGF2 for foetal and placental growth is manifest in two human disorders [250]. Patients with mutations causing Silver–Russell syndrome have low levels of IGF2 expression, foetal growth restriction, and hypoplasia of the placenta and chorionic villi [263,264]. In contrast, individuals with Beckwith–Wiedemann syndrome with increased IGF2 due to bi-allelic IGF2 expression caused by loss of imprinting have enlarged placentas and somatic overgrowth [265,266,267].

9. PTH and PTH-Related Peptide (PTHrP and PTHLH)

PTH-related peptide (PTHrP, alias PTH-like hormone, PTHLH, Online Mendelian Inheritance in Man (OMIM) entry 168470, gene PTLH) is a multifunctional protein that acts as a paracrine, autocrine, and intracrine factor [268,269,270] to regulate diverse physiological processes. These include cell proliferation and differentiation in bone and epithelial Ca2+ transport [271]. PTHrP is expressed from the very early stages of embryogenesis [268] by human and mouse trophoblasts and by immature chondrocytes in foetal bone and is abundant in the placenta and foetal membranes [268,272,273,274]. Human PTHrP is encoded by a single gene, namely, PTHLH, which undergoes complex translational and posttranslational processing. There are three principal secretory forms, namely, PTHrP (1–36), PTHrP (38–94), and PTHrP (107–139, osteostatin). PTHrP 1–36 shares homology with human PTH, with 8 of the first 13 residues being identical, and binds the peptide to the PTH receptor [275]. Hence, PTHrP can stimulate most of the actions of PTH. The peptide sequence from residue 14 has little similarity to PTH. Surprisingly, mid-region peptides, including amino acids 38–94, 38–95, and 38–101, stimulate placental Ca2+ transport and not the N-terminal as might be predicted [276,277]. A lysine/arginine-rich sequence in the mid-region may directly import PTHrP into mitochondria via importin β/Ran GTPase, enabling intracrine signalling [274,278]. The C-terminal fragment, namely, residues 107–139, has actions on skin, heart, and bone cells [274].

9.1. The PTH/PTHrP Receptor (PTH1R)

The PTH/PTHrP receptor, PTHR1, is also expressed in the placenta. It is a class B1 G-protein-coupled receptor (GPCR) [41]. There are three domains: a large extracellular N-terminal, ligand-binding domain (ECD) [279,280,281]; a transmembrane domain with seven helices linked by three extracellular loops and three intracellular loops [279]; and an intracellular C-terminal [279,280,281]. PTH1R signals primarily via Gs, which stimulates adenylyl cyclase activity but can also couple to Gq/11, which activates phospholipase C (PLC), G12/13, and Gi/o. The Gγ subunit can couple to diverse transducer proteins. PTH 1–34 and PTHrP 1–36 molecular fragments both bind to the ECD of the PTHR. However, PTH fits more tightly into the receptor’s binding cleft [275]. Binding of the PTH or PTHrP peptides to the ECD can bias signalling [279,280]. One important manifestation of this is that while both PTH and PTHrP signal from the cell surface, PTH can also induce prolonged signalling from early endosomes [115]. The PTH1R has two distinct active conformations (RG and R0). The RG conformation is G protein dependent and is associated with transient cAMP responses from the plasma membrane. It is stabilised by PTH1–34 and PTHrP1–36 indistinguishably. The R0 state is stabilised preferentially by PTH and determined by PTH activation of Gq and formation of a Gβγ, PTH/PTH1R, β-arrestin complex. This is internalised and promotes endosomal signalling [282].

9.2. Roles in the Embryo and Foetus

PTHrP is detected in mouse embryonic and extraembryonic tissues from the late morula stage onwards [268,283] and is required for blastocyst formation [284]. Furthermore, it acts as a potent vasodilator of the foetoplacental vasculature to maintain low-resistance blood flow [285]. PTHrP has essential roles in regulating normal placental morphogenesis and functional development, attainment of normal foetal growth, regulating foetal Ca2+ homeostasis and placental Ca2+ transport [286,287], and skeletal development [286,288,289]. PTHrP is increased in cord blood in foetal growth restriction [262]. The mechanism is unknown but may be an adaptive response of the vascular endothelium to hypoxia and oxidative stress to increase nitric oxide (NO) and blood vessel dilatation as has been suggested for pregnancy-specific beta-1-glycoprotein 9 (refer to Discussion). In a study of 78 pregnancies complicated by gestational diabetes, immunochemical analysis demonstrated PTHrP and PTHR1 expression in extravillous cytotrophoblast and in the uterine decidua, with small amounts in the syncytiotrophoblast and villous cytotrophoblast and none in the villous stroma. Placental PTHrP and PTHR1 expression in the extravillous cytotrophoblast was associated with a higher incidence of maternal abnormal fasting glycemia in an oral glucose tolerance test at 24–28 weeks of gestation (indicating insulin resistance), compared with a normal fasting but abnormal 60′ or 120′ glycaemia (indicating insulinopaenia). PTHR1 expression in the extravillous cytotrophoblast was associated with a lower foetal weight in the third trimester estimated by ultrasound scanning and a lower foetoplacental weight ratio. PTH-rP expression in the syncytiotrophoblast was associated with higher incidence of maternal obesity and neonatal Apgar score at 1 min <7 [290].
Pthlp-null mice die in mid-gestation [288,289], and mice expressing a truncated form of PTHrP die in the early postnatal period [291]. The placental weight of PTHrP-null mutants was reduced and associated with a reduced foetal to placental weight ratio [291]. The placental labyrinths were fissured, and the cells were disorganised and showed increased apoptosis [268]. Disrupted cell–matrix interactions resulting from changes in the profile of integrins expressed by PTHrP may have been contributory [292,293]. The blood Ca2+ concentration of null mice was significantly reduced but was corrected by foetal injection of hPTHrP (1–86) and hPTHrP (67–86) [287]. Placental Ca2+ transport and Ca2+ accretion were increased, indicating adaptation to Ca2+ depletion, but expression of trpv6 and PMCA isoforms 1 and 4 were unaltered [286].

10. The Calcium-Sensing Receptor (CaSR)

The CaSR is a class C GPCR that couples to multiple G-protein subtypes to activate intracellular signalling pathways [294]. It is expressed in the parathyroid glands [295], as well as by chondrocytes and osteoblasts in bone, with its highest expression seen in the hypertrophic chondrocytes [296]. In postnatal life, the CaSR in the parathyroid glands is activated by an increased serum Ca2+ concentration and suppresses PTH secretion. This in turn leads to decreases in renal 1,25(OH)2D3 synthesis and in bone turnover. In the foetus, the CaSR is probably activated by the raised serum Ca2+ and suppresses PTH secretion, accounting for the low observed serum PTH levels [1]. The CaSR promotes chondrocyte differentiation in foetal bone [296].
The CaSR primarily exists at cell surfaces as a disulphide-linked homodimer. The extracellular domain comprises a bi-lobed Venus flytrap domain and a cysteine-rich domain [297,298]. In the inactive state, the two CaSR protomers interact primarily at the lobe 1–lobe interface. They rotate on activation, and this extends the interface between them [297]. The ligand-bound CaSR receptor predominantly uses the Gi/o pathway to suppress cAMP and activate MAPK signalling cascades and the Gq/11 pathway, which activates Ca2+ mobilisation and MAPK signalling [294,295].

10.1. Mutations of the CaSR in Mice

In mice, ablation of one (Casr+/−) or both (Casr null) alleles of CaSR increased the foetal blood Ca2+ above the WT value and increased circulating PTH and 1,25(OH)2D3 [299]. The placenta might have been the source of the increased 1,25(OH)2D3 [100]. The circulating plasma PTHrP was significantly lower in Casr-null foetal mice compared with WT siblings, perhaps indicating the effects of the CaSR to downregulate PTHrP expression [300]. Placental transport of 45Ca was significantly, and dose-dependently, reduced in Casr+/− and Casr-null foetuses compared with WT siblings [299]. The explanation is unknown. One possibility might be that TRPV6 and/or other Ca 2+ entry channels in the placental villi were inactivated by the increased Ca 2+ (Section 4). Bone resorption was increased in global Casr-null foetuses in association with increased circulating PTH levels, increased excretion of Ca2+ and the bone resorption marker deoxypyridinoline into amniotic fluid, and a significant reduction in the mineral content of the foetal skeleton [1,299]. Chondrocyte-specific ablation of Casr (CasrChon-null mice) caused severely delayed cartilage and bone development and embryonic death [295,296].

10.2. Mutations of the CaSR in Humans

More than 400 different germline loss- and gain-of-function CaSR mutations giving rise to disordered Ca2+ homeostasis were identified by 2019 [294]. The mutations may be inherited or occur de novo. CaSR-inactivating mutations cause neonatal severe hyperparathyroidism (NSHPT). The median age of the diagnosis of NSHPT is 14 days, ranging from 2 days to several months of age. No case reports have described blood Ca2+ or PTH values in foetuses or newborns [1]. Presentation is with symptoms of hypercalcaemia and skeletal demineralization with fractures. The parathyroid glands are grossly enlarged, and serum PTH is high. Without urgent parathyroidectomy, this condition is almost invariably fatal [294,301]. Most reported cases are due to homozygous (around 75%) or compound heterozygous mutations (around 10%). A minority of patients had heterozygous mutations of CaSR located in regions that are critical for receptor activation [297,298]. Individuals heterozygous for dominant CaSR mutations (familial hypocalciuric hypercalcaemia) have mild or moderately raised serum Ca2+, which is often asymptomatic and detected as an incidental finding in later life [295] and has not been associated with foetal Ca2+ disturbances. Mutations in the Gα11 protein and the adaptor protein-2 sigma subunit (AP2σ), by which the CaSR is internalised, cause CaSR signalling disturbances, demonstrating novel mechanisms by which CaSR is internalised and that CaSR can signal by an endosomal pathway [302].

11. Calcitonin and Calcitonin Gene-Related Peptide (CGRP)

Identifying the roles of calcitonin in the placenta is proving difficult for four reasons: (i) the calcitonin gene (CALCA, alias Ctcgrp1) encodes two distinct proteins, calcitonin and calcitonin gene-related peptide (CGRP), with multiple isoforms produced by gene splicing; (ii) the two proteins bind to different receptors, CALCR and CRLR; (iii) the gene is expressed by both the uterine endometrial decidua and the foetal trophoblast; and (iv) the nomenclature is confusing. The gene has six exons, which are all included in the primary RNA transcript. Calcitonin or CGRP mRNA is formed subsequently by posttranscriptional processing. Transcription of CGRP is downregulated by vitamin D. A second CGRP homolog, β-CGRP, is produced by a separate gene [303,304,305,306,307]. Human CGRP and β-CGRP differ by three amino acids [308].

11.1. Calcitonin

Calcitonin is a 32-amino-acid peptide calcium-lowering hormone. It is synthesised in the parafollicular cells or “C cells” of the thyroid gland [304] and is liberated by endocrine secretion. It regulates Ca2+ levels in bone and kidney cells. The placenta has been shown to express calcitonin mRNA and protein and the calcitonin receptor [309]. Calcitonin may have a critical role during blastocyst implantation. There was a sharp burst of calcitonin mRNA and protein expression in the endometrium of pregnant rats from day 2 to day 5 pf., encompassing the time of blastocyst implantation, which ended abruptly by day 6 pf., when implantation was completed [310,311]. Expression was localised to the endometrial glandular epithelial cells and stimulated by progesterone. Oestrogen in a low dosage synergised with progesterone but had no effect alone [310,311]. Attenuation of calcitonin expression blocked the implantation of the early embryo [311]. There are no reports of human genetic calcitonin deficiency.

11.2. The Calcitonin Receptor (CALCR, CTR)

The calcitonin receptor is a class B GPCR. Numerous alternatively spliced transcripts of the human CT receptor gene (CALCR alias CTR) have been reported. In contrast to the CGRP receptor (Section 11.3), the calcitonin receptor does not require receptor-activity-modifying protein (RAMP1) to bind and respond to calcitonin. Two splice variants of the human CT receptor differ by the presence (CT(b)) or absence (CT(a)) of 16 amino acids in the first intracellular loop. The CT(a) receptor is expressed widely. In most tissues, CT(b) receptor expression is low [312] but is reported to be significant in the ovary and placenta [305]. The two variants have similar affinity for peptide ligands, but unlike the CT(a) receptor, the CT(b) receptor is poorly internalised and has altered coupling to the G protein in response to stimulation, with no Gq-mediated responses and attenuated Gs-mediated signalling [305]. CALCR ablation causes embryonic death at mid-gestation for unknown reasons [1].

11.3. CGRP, the Calcitonin Receptor-like Receptor (CRLR), and the Functional CGRP Receptor

Published data for foetal expression of CGRP are scant. In three very premature foetuses, CGRP was expressed in the endothelium of human umbilical cord blood vessels at delivery. The % methylation of the promoter of CGRP was higher in cord blood DNA from preterm than term infants [313]. Progesterone stimulated CGRP expression, while oestrogen inhibited it [314]. CGRP is a potent vasodilator [315]. It was suggested that CGRP may have a role in increasing blood flow through the foetoplacental unit in late gestation [314].
The functional CGPR receptor is a complex of CGRP, the calcitonin receptor-like receptor (CRLR), RAMP1, and an additional intracellular protein, namely, the receptor component protein (RCP) [303,305,307,316]. The CRLR by itself is not expressed significantly at the cell surface and does not respond to any known ligand. With RAMP1 it becomes the CGRP receptor (i.e., CRLR/RAMP1). RAMP1 is an intrinsic protein with an extracellular N terminus, a single transmembrane domain, and a short intracellular domain [305,307,317]. Immunoreactive CGRP receptors were demonstrated in the placental labyrinth of foetal rats in the trophoblasts and in the endothelium and underlying smooth muscle cells of villus blood vessels. Levels increased 10-fold from day 17 pf. to a peak before labour on day 22 pf. [314].

11.4. Calcitonin and CGRP in Regulation of Mineral Status

A small number of calca (ctgrp)-null mice died. The survivors had normal lifespans and were fertile. The serum Ca2+, placental transport of Ca2+, and skeletal Ca2+ content were not significantly different from littermate controls [318], demonstrating clearly that foetal-placental Ca2+ transfer does not require calcitonin or CGRP. Notable observations were that Mg2+ was significantly reduced in serum of calca-null mothers and foetuses, as well as in the foetal skeleton. There was no obvious explanation [318]. One possibility might be that calcitonin or CGRP interacts with the TRP melastatin cation transporters TRPM7 and/or TRPM6, which interact and are both expressed by placental trophoblasts (Section 4.1). TRPM7 forms homotetrameric channels that are highly permeable to Ca2+, Mg2+, and Zn2 and are regulated by Mg2+. Depletion of intracellular Mg2+ and Mg-ATP promotes TRPM7-mediated uptake of extracellular Mg2+ [319,320]. The interaction of TRPM6 with TRPM7 was proposed to increase epithelial Mg2+ transport [321].

12. Vitamin D and Placental Ca2+ Transport

12.1. Importation of 25OHD and Activation to 1,25(OH)2D

Serum ionised Ca2+ concentrations in human cord blood are uniformly around 0.30–0.50 mM higher than in maternal serum [1]. A summary of data for serum vitamin D metabolites in foetal rodents and lambs showed that circulating concentrations of 1,25-dihydroxyvitamin D3 (1,25(OH)2D3) were low and that in foetal mice, 24,25-dihydroxyvitamin D3 (24,25(OH)2D3) was 54-fold higher than 1,25(OH)2D3 and three times higher than maternal 24,25(OH)2D3 [322]. Published data for human foetuses are sparse. In one study, cord sera from 21 preterm foetuses were compared with paired maternal sera. The cord and maternal serum Ca2+, Mg2+, and phosphorus concentration correlated closely, but levels of all three analytes were significantly higher in the cord samples. Cord serum 25-hydroxyvitamin D3 (25(OH)D3), 24,25(OH)2D3, and 1,25(OH)2D3 levels were significantly lower than those observed for the mothers. Whereas the concentrations of 25(OH)D3) and 24,25(OH)2D3 in cord blood both correlated directly with maternal blood levels, 1,25(OH)2D3 levels showed no association. Maternal 1,25(OH)2D3 levels were not related to the gestational age or to maternal Ca2+, Mg2, inorganic phosphate, or 25(OH)D3 concentrations. Cord 1,25(OH)2D3 correlated significantly only with cord Ca2+ levels [323]. Other studies found that (i) the average ratio of foetal serum 25(OH)D3 at 39.5 ± 1.2 weeks of gestation and maternal serum at 34–35 weeks of gestation was 0.67 (interquartile range, 0.56–0.84), with a correlation of r = 0.89 (p < 0.0001) [324], and (ii) cord blood concentrations of 25(OH)D3 and 1,25(OH)2D were tightly correlated (r = 0.78, p = 0.0001), in contrast to adult blood, in which 1,25(OH)2D remained relatively constant over a range of 25(OH)D concentrations (r = 0.02, p = 0.89) [325].
More than 99% of serum vitamin D metabolites are bound to carrier proteins, 85–88% with high affinity to D-binding protein (DBP) and 12–15% to albumin [326,327]. In the kidneys, both DBP and albumin with their vitamin D ligands bind to megalin and are internalised into apical endosomes by a clathrin-dependent endocytosis [100,328,329,330]. They are then transported via the endosome/lysosome pathway to lysosomes, where the binding proteins are degraded, and the ligands are released [327]. The 25(OH)2D3 is hydroxylated to inactive 24,25(OH)2D3 by CYP24A1 or taken into mitochondria for hydroxylation to 1,25(OH)2D3 by CYP27B1. It has now been shown that 25(OH)2D3 is taken up actively from the blood by trophoblasts from human term placentas into sub-membranous vesicles, probably by clathrin-dependent endocytosis [100]. The trophoblast cell surface is actively endocytic and has numerous coated pits and vesicles located between the microvilli and in the apical cytoplasm [331,332,333]. Placental syncytial trophoblasts express megalin [334]. The 1α-hydroxylase (CYP27B1) and 24-hydroxylase (CYP24A1) are both expressed in syncytiotrophoblasts in human placenta [100]. The mechanism by which 25(OH)2D3 reaches the mitochondria for activation to 1,25(OH)2D3 is uncertain. Intracellular vitamin-D-binding proteins (IDBPs) from the heat shock protein 70 family have been identified [327]. IDBP was found to bind 25(OH)2D3 and other steroid hormones. One proposal is that IDBP may chaperone 25(OH)2D3 for delivery to the mitochondria for activation [335]. An alternative proposal is that megalin may shuttle 25(OH)2D3 with its binding protein to the mitochondria via the retrograde early endosome to the Golgi pathway [330].

12.2. From Generation of 1,25(OH)2D3 to Activation of Gene Translation

The 1,25(OH)2D3 binds to the vitamin D receptor (VDR) with a very high affinity of 0.1 nM [336]. Ligand-bound VDR heterodimerises with the nuclear receptor retinoid X receptor (RXR). This increases the affinity of the VDR for the vitamin D response element (VDRE) in the promoter region of vitamin-D-responsive genes. A complex of supportive nuclear proteins is recruited to the site. These include possible co-receptors, pioneer factors, co-factors, members of the Mediator complex [337], chromatin modifiers, and chromatin remodellers [336]. Corepressors typically work by recruiting histone deacetylases (HDACs) or methyl transferases (MTs) to the gene. These coregulators can be specific for different genes and may be expressed differentially in different cells, thus providing some specificity for the actions of 1,25(OH)2D3/VDR. In addition, 1,25(OH)2D3 can affect the epigenome via direct interaction of the VDR/RXR complex with chromatin-modifying enzymes or indirectly by regulating genes that encode chromatin modifiers [336].

12.3. 1,25(OH)2D/VDR in Ca2+ Regulation in Pregnancy

The 1,25(OH)2D3 has a central role in maintaining normal Ca2+ status postnatally by promoting transcellular and paracellular Ca2+ absorption in the duodenum and kidneys. This is largely through its actions to coordinate transcription of the numerous proteins involved in operating the Ca2+ transport pathways. This is not the case for placental transport, which is driven by oestrogens and not by vitamin D. A review of evidence from numerous sources found that vitamin D, 1,25(OH)2D3, and the VDR are not required for the foetus to maintain normal serum mineral concentrations [1]. This conclusion is substantiated by observations of inherited disorders of vitamin D metabolism in humans that are not manifest at birth but present later in infancy. Individuals with inherited deficiencies of CYP27B1 or of the VDR appear normal at birth and generally present with hypocalcaemia and/or rickets at 2 to 24 months [338,339] and 2–8 months of life [338], respectively. Infants with biallelic loss-of-function mutations in CYP24A1 (idiopathic infantile hypercalcemia) presented with severe hypercalcemia at 6 to 8 months of age [340].
A small study found increased expression of hOGG1 (human 8-oxoguanine DNA glycosylase) mRNA in placentas from pregnancies with pre-eclampsia (n = 20) or gestational diabetes (GDM, n = 20) indicative of oxidative stress, compared with 20 controls. The bone mineral content (BMC) of newborns was reduced. Both groups had lower concentrations of cord blood 25(OH)2D3; significantly lower placental mRNA expression of CYP2R1 (cytochrome P4502R, product vitamin D 25-hydroxylase), VDR, TRPV6, TRPV5, CABP9k, CaBP28k PMCA1,2,3, IP3R, and the ryanodine receptors RyR1,2,3; and significantly higher expression of CYP27B1 and CYP24A. There was a negative correlation between BMC and CYP2R1, CYP24A1, VDR, CABP28K, and PMCA2. It was proposed that hypoxia disturbed vitamin D metabolism and calcium transport [341]. Another study, similarly, found that compared with controls (n = 40), serum 25(OH)2D3 was lower in maternal and cord blood from 41 pregnancies with gestational GDM. Placenta and umbilical cord tissues from GDM pregnancies had significantly higher mRNA and protein expression of CYP24A1, but unlike the above study, the expression of VDR was higher, and CYP27B1 was significantly lower [342]. Larger studies are required to establish whether vitamin D metabolism is disturbed in GDM and pre-eclampsia and to investigate the functional significance if confirmed.

13. Genetic Disorders of Placental Ca2+ Transport

Observations of inherited defects demonstrate the roles of Ca2+ trafficking in vivo. Table 2 lists genetic disorders of calcium transport that are manifest in the human foetus in utero and/or at delivery.
Other disorders of the calcium supply chain present with a variety of clinical problems from early infancy to adult life [356]. Biallelic loss of Orai1 causes severe combined immunodeficiency (SCID) presenting with recurrent infections and autoimmunity from a few weeks of birth [189,357,358]. Heterozygous autosomal dominant mutations of Orai1 or STIM1 cause an overlapping spectrum of disorders of skeletal muscle contraction and platelet function [359], which include tubular aggregate myopathy (TAM) [189,357], Stormarken syndrome [360], and York platelet syndrome [189,357]. Deficiency of calsequestrin, an ER Ca2+-binding protein, causes myopathy [361]. Inherited deficiencies of CYP27B1, VDR, and CYP24a1 generally present after 2 months of age (Section 12) [338,339,340]. TRPC6 deficiency causes renal focal glomerulosclerosis in childhood [362]. Inherited syndromic defects of other TRPCs, SERCA, IP3/IP3R, or calbindin D9k have not been reported.
  • Section D: The Developmental Origins of Health and Disease (DOHaD) and Postnatal Bone Development

14. DOHaD and Postnatal Bone Development

David Barker [363] first suggested that adverse nutritional and environmental exposures during pregnancy may programme a foetus to have a higher risk for common chronic diseases in adult life. This concept of developmental plasticity, now termed the developmental origins of health and disease (DOHaD), has gained momentum and stimulated extensive research [2,17,364,365]. Epidemiologic observations that smaller size or relative thinness at birth and during infancy is associated with increased rates of coronary heart disease, stroke, type 2 diabetes mellitus, adiposity, metabolic syndrome, and osteoporosis [366] in adult life have been extensively replicated [17]. Developmental plasticity requires stable modulation of gene expression, and this appears to be mediated, at least in part, by epigenetic processes such as DNA methylation and histone modification [17,367]. Such modifications may subsequently promote chronic diseases under favourable enriched environmental conditions [365].

14.1. Evidence for an Association of Intrauterine Factors with Bone Health in Later Life

14.1.1. Observational Epidemiologic Studies

One small study found that a low level of maternal serum 25(OH)D during pregnancy was associated with a small, significant reduction in bone calcium content at 9 years of age, which was not evident at birth [368]. No association was observed in a larger study [369]. High maternal intakes of Ca2+ and folate [370] and of phosphorus and protein, as well as maternal serum homocysteine and vitamin B12 concentrations in the first trimester [371], were associated with an increase in bone mineral at 6 years of age. Roseboom et al. (2019) [365] and Vaiserman and Lushchak (2021) [372] reviewed studies of health outcomes of in utero famine exposure [373,374,375] and findings from birth cohorts from Hertfordshire, UK; South Africa; Finland; the United States; and Brazil [365]. None of these studies appear to have addressed effects on postnatal bone calcification. However, prenatal malnutrition in Dutch [376,377] and Gambian [375,378] famine exposure was associated with alterations in methylation patterns in adult life, which correlated with phenotypic outcomes. This suggested that early exposure to nutrient restriction may influence epigenetic control of metabolic processes later in life. Observations from two studies of a population-based mother–offspring cohort [379,380] provided evidence for this. One study found a significant negative correlation between the percent methylation of RXRA at five of six CpG sites analysed in cord blood with bone mineral content at 4 years of age [379]. The other study demonstrated a significant inverse association of methylation at sites in the promoter of CDKN2A, which encodes long non-coding RNA ANRIL in cord blood with bone mineralisation at 4 and 6 years of age [380]. Differential DNA methylation of this gene at birth was shown previously to correlate with childhood adiposity [381].

14.1.2. Intervention Studies in Which Women Received a Vitamin D Supplement

Two prospective controlled studies of children in the first 6–7 years of life indicate that high doses of antenatal vitamin D supplementation have beneficial effects on offspring skeletal mineralisation. In the Maternal Vitamin D Osteoporosis Study (MAVIDOS), 1000 IU of cholecalciferol was prescribed daily from 14 weeks of gestation to delivery [382,383]. There was no difference from placebo-treated controls at birth in dual-energy X-ray absorptiometry (DXA) bone scans. At 4 years of age a small increase in whole-body-less-head (WBLH) bone mineral density (BMD) was observed. Scans of 447 children followed up at ages 6–7 y demonstrated that offspring of gestationally supplemented mothers had higher WBLH bone mineral content (BMC) [0.15 SD, 95% confidence interval (CI): 0.04, 0.26], bone mineral density (BMD) (0.18 SD, 95% CI: 0.06, 0.31), and bone mineral apparent density (BMAD) (0.18 SD, 95% CI: 0.04, 0.32) compared with the controls. In the Copenhagen Prospective Studies on Asthma in Childhood 2010 (COPSAC2010) trial [384], high-dose vitamin D supplementation with 2800 IU/d from pregnancy week 24 to 1 week postpartum was compared with a standard dose of 400 IU/d. A combined analysis of DXA scans at age 3 and 6 years of 517 children showed that children in the vitamin D vs. placebo group had higher whole-body BMC, where the mean difference adjusted (aMD) for age, sex, height, and weight was 11.5 g (95% CI, 2.3–20.7; p = 0.01); higher WBLH BMC aMD, 7.5 g (95% CI, 1.5–13.5; p = 0.01); and higher head BMD aMD, 0.023 g/cm2 (95% CI, 0.003–0.004; p = 0.03). The largest effect was in children from vitamin-D-insufficient mothers and among winter births. There was a small decrease in the incidence of fractures in the vitamin D group (n = 23 vs. n = 36; incidence rate ratio, 0.62 (95% CI, 0.37–1.05; p = 0.08)). A review of 10 randomised and quasi-randomised placebo-controlled trials evaluated the effect of vitamin D supplementation alone or combined with calcium or other micronutrients during pregnancy [385]. None addressed foetal calcium and bone status. The evidence of benefit for pre-eclampsia, preterm birth, and birthweight was uncertain.

14.2. Epigenetic Processes and DOHaD

Epigenetic modifications of DNA in the foetus in response to adverse conditions in utero are proposed to contribute to poor metabolic health later in life [2,17,79,364,386,387,388]. Epigenetic activities modulate gene expression in numerous signalling pathways controlling key functions of the human placenta. Mechanisms include DNA methylation, histone tail posttranslational modifications (PTMs), and non-coding RNAs (ncRNAs). DNA methylation and histone PTMs regulate chromatin accessibility to transcription factors to facilitate or repress gene expression. ncRNAs are strong posttranscriptional regulators (reviewed [389,390]). DNA is methylated by covalent modification of carbon-5 of cytosine, catalysed by DNA methyltransferases (DMNTs). Methylation is erased by ten-eleven translocases (TETs) [391,392,393,394,395,396]. DNA methylation patterns in 12 cancer cell types showed that genes regulating Ca2+ transport are major targets of hypermethylation and downregulation [397]. In mice, knock-down of tet2 decreased muscle stem cell proliferation and differentiation and disrupted calcium homeostasis. Muscle cell Ca2+ concentration was lower in tet2-null cells than in wild-type cells, and RNA expression of calcium-pathway-related genes was drastically reduced. Hypermethylated genes were predominantly enriched in the calcium pathway [396].
A critical risk period for introducing persisting disturbances of DNA methylation must be in very early embryonic life at around the time the blastocyst is implanted in the uterine decidua and embryogenesis is commencing [79]. Prior to fertilisation, genes of male and female gametes are selectively methylated. Starting immediately after fertilisation, both paternal and maternal DNA is demethylated via TETs. Simultaneously, phospholipase C PLCζ is activated, resulting in the release of Ca2+ from the ER stores and Ca2+ oscillations, which persist for about 2–4 h postgamete fusion [79]. Demethylation is essential to abolish gamete identity and confer cells of the growing embryo with pluripotent potential [79,392,394,398,399]. DNA methylation is regained following implantation and proceeds in a lineage-specific pattern. Histone-posttranslational modifications and epigenetic modifications are also reprogrammed during early embryogenesis [400]. Abnormal postfertilization Ca2+ profiles disrupt gene expression [79,401]. Dampened Ca2+ oscillations jeopardise preimplantation development, with RNA processing and polymerase II transcription genes particularly affected [402]. Hyperstimulation of Ca2+ oscillations compromises postimplantation development. These observations indicate that interventions to reduce the risk of adverse epigenetic changes should probably be well established before pregnancy.

15. Discussion

It has not been too long since it was commonly thought and taught that the main requirement for Ca2+ in utero is to mineralise the foetal skeleton and that to meet this need, Ca2+ is transferred from maternal blood to the foetus by 1,25(OH)2D3 synthesised in the placenta. The mechanisms were unclear but were probably, like intestinal Ca2+ absorption, promoted by circulating 1,25(OH)2D3. Over the last two or three decades, we have learnt that Ca2+ transport in the foetoplacental unit is far more complex than could have been envisaged. This is because of the power of the Ca2+ ion to orchestrate the activity of numerous signalling pathways and to direct/change cellular processes within milliseconds [67]. This additional role must be tightly regulated and matched to need and requires the interaction of numerous proteins, carriers, and transporters. The processes involved are under intensive investigation on a broad front. The aim here was to pull the data together to obtain an overview of our current understanding of the mechanisms and to highlight areas needing more clarification.
A normally formed and functional placental villous system with a large covering of terminal villi—the workforce of the mature placenta—is of paramount importance for providing an adequate mineral supply to the foetus. This was demonstrable for two proteins, IGF2 (Section 8) and PTHrP (Section 9), for which a deficiency in rodent models was associated with abnormal placental structure and reduced Ca2+ delivery in late gestation. Both were expressed very early in embryonic life, at the critical time for blastocyst implantation. In humans with Silver–Russell syndrome with low levels of IGF2 expression, the placenta and chorionic villi are hypoplastic, and foetal growth is restricted (Section 8 and Section 13). It is important to note that IGF2, PTHrP, and TRPV6 are expressed in bone as well as the placenta and appear to operate independently at these sites. How much the local effects of deficiencies of these proteins in bone contribute to the skeletal changes in global defects is unknown.
The placental villi must be equipped with a variety of Ca2+-importing channels, which can collectively respond to a range of different stimuli. There is abundant evidence that TRPV6 is the principal Ca2+ channel of the terminal microvilli, which are operative from around 16 weeks of gestation (Section 4.2). TRPV6 expression rises in synchrony with expansion of the placental syncytium and increases exponentially in the third trimester. Identification of human inherited TRPV6 deficiency in 2017 was a clear demonstration of its physiological importance. The expression of TRPV6 is low at earlier stages in gestation when the requirement for bones is also low. Other channels from the TRP family expressed in the placenta probably meet the Ca2+ requirement for signalling, despite their lower Ca2+ selectivity (Section 4.1).
What is not clear is how channels embedded in the apical membranes of the terminal microvilli are transported in and out of the villi and secured in the membranes. Like duodenal microvilli, they have a central actin core, but the transport mechanisms are enigmatic. The scaffolding protein EBP50, renamed NHERF1, was first isolated from placental villi (refer to Section 2: Terminal villi), but this appears to have been largely ignored, and its role in the placenta has not been pursued. In the microvilli of the renal proximal tubules, NHERF1 is an extremely versatile scaffold located close to the villus membranes. It provides support for transporters at the cell surface; links them to the cell cytoskeleton; and, importantly, connects them with large transiently assembled signalling complexes. In the kidneys, NHERF1 has an important role in the regulation of the sodium phosphate transporters NaPi-2a and NaPi-2c and the operation of the Na+/H+ exchanger NHE3 [115]. These proteins are all active in placental transport, and it seems likely that NHERF1 similarly has a role in the function of placental transporters. This merits investigation, as does the role of ezrin, which accounts for 5% of proteins in villus extracts and is another “neglected” placental protein (Section 2). Ezrin is important for linking membrane proteins with the actin cytoskeleton and may have a role in the transport of villus membrane channels and receptors, and possibly of the Ca2+ bound to CaM in the microvilli.
Ca2+ absorbed by the placental villi for export to the foetus, and not for local use in the placenta, must be chaperoned during transport through the cytosol to the basal membrane for extrusion by Ca2+-ATPase. In duodenal villi, calbindin D9k (CaBP-9k) is the principal carrier, and its expression is increased by 1,25(OH)2D3. The identity of the chaperone in the placenta is uncertain, and data are lacking. Placental Ca2+ absorption is driven by oestrogens and not by 1,25(OH)2D3. There are differences between species in placental expression of CaBP-9k, which may be attributable to small differences in the nucleotide sequence of the oestrogen response element on the CaBP-9k gene s100g. It is not known for certain whether the human placenta expresses CaBP-9k. This is another issue that requires clarification.
Cells maintain a store of Ca2+ in the ER, which is immediately available for signalling when membrane receptors are stimulated. Clearly, when the stores are getting low, they must be topped up with Ca2+ brought into the cells from the extracellular space. In many tissues, this is achieved by a partnership between a protein located in the ER, STIM1, and a Ca2+ channel, Orai1, located in the plasma membrane. STIM1 is a sensor that monitors the ER Ca2+ stores and increases Ca2+ entry into the cell by activating its effector Orai1 when stores are low. This store-operated Ca2+ entry (SOCE) system could be a mechanism that matches Ca2+ supply to need in the foetoplacental unit. However, there is no available evidence to date that STIM1 and Orai1 are expressed by the villus syncytiotrophoblast (Section 5.3); therefore, this requires confirmation. Because STIM1/Orai1 are widely expressed in organs of mammals postnatally, it would be expected that they are also expressed in the developing foetal organs. The observation that Orai1 was generated in vitro by cells likely to be embryonic stromal cells [203], and by foetal endothelial cells [204], may be preliminary supportive evidence warranting further exploration. Two studies demonstrated that pregnancy-specific beta-1-glycoproteins from the immunoglobulin superfamily, PSG1 [203] and PSG9 [204], increased production of Orai1 and intracellular Ca2+ entry (Section 5.3). In response, PSG1 increased trophoblast migration, and PSG9 increased the expression of endothelial nitric oxide synthase (eNOS) and NO production. The suggestion that downregulation of PSG1, PSG9, and Orai1 channel function may be relevant to pre-eclampsia and could have therapeutic potential merits investigation. There is emerging evidence that Orai and STIM may contribute to cardiovascular disease [403], pulmonary hypertension [404], and diabetes [405,406], and the possibilities of pharmacological manipulation are being explored. The findings of these studies may provide insight into placental function.
Because STIM1 and Orai1 are apparently not the SOCE system in the placental syncytium, what is? A strong contender is calmodulin (CaM), which is abundant in terminal microvilli [234]. CaM is not just a Ca2+ binder/transporter but also both a Ca2+ sensor and effector. CaM loaded with Ca2+ controls the two key processes in intracellular Ca2+ regulation: Ca2+ entry by regulating closure of TRPV6, the principal Ca2+ import channel, and Ca2+ export by increasing the activity of PMCA1b, the major Ca2+-exporting enzyme. CaM is a sensor for increased cytosolic Ca2+ because it is activated by Ca2+ loading. Activation extends and opens the CaM molecule and separates its N- and C-lobes. These can attach to two sites on the targets TRPV6 or PMCA1b and alter their molecular conformation, in both cases by displacing flexible unstructured peptide sequences in their protein tails (Section 4.2.1 and Section 6.2). It is notable that there are multiple CaM-binding sites both on the polybasic domain of STIM1 downstream of a long unstructured peptide sequence and on the N-terminal of Orai1. It may be that in cells that express STIM1 and Orai1, Ca2+-CAM could pull the two proteins together and participate in their interaction. Ca2+-CAM may also regulate activation of TRPC channels, which, like TRPV6, have CaM-binding domains in their carboxyl terminals [81].
TRPV6 must be expressed on the villus surface membrane to import Ca2+ into the trophoblasts. One mechanism for this is through interaction with cytosolic S100A10/annexinA2 (ANXA2) heterotetramers (Section 4.2.4). The ANXA2 moiety links with phospholipids in the plasma membrane and to the actin cytoskeleton. The S100A10 moiety attaches to a conserved binding site in the TRPV6 C-terminal. Association with the heterotetramer stabilises the channel at the cell surface. It is notable that both S100A10 and ANXA2 were among the 115 core genes expressed in term placentas across 14 mammalian species [25], suggesting that the membrane-binding function of the ANXA2/S100A2 heterotetramer has a much wider role than regulating TRPV6 expression. It may be involved in endocytosis or exocytosis and/or the regulation of intracellular transport of other ion channels and transporters.
The review considered evidence for involvement of four peptides and vitamin D in placental Ca2+ transport, presented in Section 8, Section 9, Section 10, Section 11 and Section 12. Small studies reported differences in the expression of genes that regulate vitamin D metabolism and Ca2+ transport in placentas from pregnancies complicated by pre-eclampsia [341] and gestational diabetes mellitus [341,342] (Section 12.3). Larger studies are required to establish whether vitamin D metabolism is disturbed in GDM and pre-eclampsia and to investigate the functional significance if confirmed. IGF2 and PTHrP are clearly important for normal blastocyst implantation and placental development (paragraph 2 above). Some additional observations are highlighted here.
Unlike PTH, PTHrP is not normally a circulating hormone but has paracrine, autocrine, and intracrine actions. It is abundant in the placenta, expressed very early in embryonic life and needed for blastocyst, placental, and skeletal development but normally not for regulating foetal serum Ca2+. PTHrP stimulates placental Ca2+ transport but, surprisingly, not through binding to the PTH1R receptor through its amino terminal, which has close homology with the PTH N-terminal, but instead through amino acid sequences in the middle of the molecule with no PTH counterparts. It is speculated that these bind with a receptor that is distinct from the PTH receptor; however, its identity is unknown. Of possible relevance is that there is a lysine-rich sequence in the middle of PTHrP, which may direct PTHrP into the nucleus [270,274]. Perhaps the activation of nuclear transcription factors accounts for many of the activities of PTHrP, which may include activation of placental Ca2+ transport.
Calcitonin is the product of the CALCA gene. Another protein, CGRP, is also generated from this gene by posttranscriptional modification of the gene transcript. The two proteins have separate receptors, and there is considerable uncertainty about the roles of these proteins in the placenta. The facts and proposals so far are that neither calcitonin nor CGRP is required for foetal–placental Ca2+ transfer and that calcitonin produced by uterine endometrial glands from 2 to 6 days of gestation may have a role in blastocyst implantation [1]. An intriguing unanticipated finding was that in mice, Mg2+ was significantly reduced in the serum of calca-null mothers and foetuses, as well as in the foetal skeleton. There is no obvious link between calcitonin and Mg2+ metabolism. One possibility, which has not been considered, is that one of the CALCA gene proteins interacts with the TRP carriers TRPM6 and TRPM7 (Section 4.1). Together, these regulate the cellular importation of Mg2+ and Mg2+-ATP; perhaps the function of calcitonin at implantation is to supply Mg2+-ATP. In one published study, the % methylation of the promoter of CGRP was higher in cord blood DNA from preterm than term infants (Section 11), but published data on the role of this peptide in the placenta are lacking. Postnatally, CGRP is a vasodilator, and roles in the foetoplacental vasculature were suggested [314] but without any supportive evidence.
Observations on mouse foetuses with CaSR ablation confirm the normal role of the CaSR to suppress PTH secretion by the parathyroid glands. Without CaSR, foetal serum PTH and Ca2+ and bone turnover are markedly increased (Section 10). The increased Ca2+ must be from bone resorption because placental Ca2+ transport was low; no explanation was offered for this observation. It may be that TRPV6 and other Ca2+ import channels were inactivated by the raised Ca2+ (Section 4.2.1). From case reports, CaSR deficiency was not evident at birth in humans but presented very early in neonatal life with symptomatic hypercalcemia. Already, there were marked parathyroid gland enlargement and bone demineralisation, probably indicating that the problems started in foetal life. The ion and fluid disturbances may be partially corrected in utero via the foetomaternal circulation.
The DOHaD proposal [363] has generated intense research. There is high-quality evidence for associations with increased rates of coronary heart disease, stroke, type 2 diabetes mellitus, adiposity, and metabolic syndrome and early evidence for a link with osteoporosis (Section 14). The sticking block has been to explain the mechanism for the association. As the extent and diversity of epigenetic modifications have become evident, these have emerged as likely mediators. The contribution of transposons to regulation is another aspect to explore [407,408,409]. The genes regulating Ca2+ transport are under strong epigenetic control. This underlies attempts to decrease the risk of osteoporosis in later life by manipulating vitamin D and dietary mineral intakes, in addition to lifestyle changes of mothers in pregnancy. There is early evidence that these may succeed, but more large-scale studies that incorporate recent scientific observations are required. The most critical period for adverse modifications of methylation and histones is probably early in embryogenesis (Section 14.2.), suggesting that educational and dietary interventions should be in place antenatally to prepare for pregnancy to reduce risk.
From the above summary, areas of current uncertainty requiring clarification are as follows:
  • The mechanisms of trafficking transporters and ion channels to and from the apical membranes of placental villi.
  • Whether NHERF1 is expressed in placental villi and, if expression is confirmed, the roles of this scaffold in villi.
  • The roles of ezrin in placental villi.
  • Whether calbindin D9k is expressed in the villus syncytium.
  • How Ca2+ absorbed for export to the foetus is chaperoned or transported across the trophoblast safely for extrusion.
  • Whether Orai1 and STIM1 are expressed by syncytiotrophoblasts.
  • The roles of Ca2+-CaM in mediating SOCE in the syncytiotrophoblasts.
  • The roles of PTHrP mid-molecular residues in regulating DNA transcription, particularly of genes involved in Ca2+ transport.
  • Whether CGRP is expressed in the placenta and, if so, its likely roles.
  • Although not relevant to Ca2+ transport, it might be of wider interest to find out whether calcitonin or CGRP interacts with TRPM7 or TRPM6.
  • Whether placental metabolism of vitamin D is disturbed in gestational diabetes and pre-eclampsia.
  • Whether pregnancy-specific beta-1-glycoproteins PSG1 and PSG9 increase endothelial NO production and placental blood flow by increasing Orai1 expression.
For the future. The generation of functional multicellular placental organoids from human placental villi or naïve human pluripotential stem cells is a major achievement that will enable studies of normal placental development and villus function and identification of pathologic disturbances, which may lead to pre-eclampsia or IUGR [13,29,32,33,410]. They may also be useful for generating new non-malignant human trophoblast cell lines [411]. Rigorously undertaken analyses of extracellular vesicles derived from the foetal placenta and circulating in maternal blood may uncover predictive markers for placental pathology, enabling early clinical intervention [412,413]. Continued expansion of our knowledge of epigenetic regulation of expression of genes controlling placental Ca2+ transfer will guide interventions to protect bone health in later life. If the greatest risk for acquiring epigenetic disorder is in early embryonic life, then dietary and lifestyle improvements should be well established before conception.

16. Conclusions

Recent studies have advanced our insight into the mechanisms of placental Ca2+ transport considerably, although there are still many gaps to fill. What has emerged is that meeting the high Ca2+ demand in the third trimester depends on normal implantation of the blastocyst and placental development in early embryonic life. Ca2+ signalling plays an integral part in these processes, and a complex network of Ca2+ channels and transporters regulate its availability.

Funding

This review received no external funding.

Acknowledgments

I am grateful to Geerten Vuister, Structural Biology, University of Leicester, UK, for discussions about TRPV6 and calmodulin and for designing and producing Figure 6. This was reproduced from our review, Biochemistry and pathophysiology of the Transient Potential Receptor Vanilloid 6 (TRPV6) calcium channel. Adv Clin Chem. 2023; 113:43–100. I thank Elsevier, who hold the copyright, for granting me permission to use this figure here and Wiley-Blackwell, John Wiley and Sons for permission to use Figure 2. I also thank Graham Burton, Centre for Trophoblast Research, University of Cambridge, for sourcing electron microscopy references and encouragement.

Conflicts of Interest

The author declares no conflicts of interest.

Definitions and Abbreviations

Definitions
Embryo (human)A conceptus from fertilisation to the end of 8 weeks pf. (postfertilisation) approximately10 weeks following the last menstrual period (LMP).
Foetus (human)A conceptus from 9 weeks pf. to term at 38 weeks pf. (approximately 11 to 40 weeks post-LMP).
Preterm birth20 weeks to 37 weeks following LMP: term birth, 37 weeks to 41 weeks 6 days.
TrimestersFirst, up to 13 weeks following LMP; second, 13 to 28 weeks; third, 29 weeks to 40 weeks or until birth.
Intrauterine growth restriction (IUGR)Birth weight < 10th percentile for gestational age and below the genetic growth potential.
LabyrinthThe inner compartment of rodent placenta that contains the highly branched microvilli, which undertake nutrient exchange.
Abbreviations
ANXA2Annexin A2
BBMBrush-border membrane
CaBPCa2+-binding protein
CaBP-9kCalbindin-D9K
CGRPCalcitonin gene-related peptide
CaMCalmodulin
CALCR (CTR)The calcitonin receptor
CaSRThe calcium-sensing receptor
CRACCa2+ release-activated Ca2+ channels
CRLRThe calcitonin gene-related peptide receptor
DOHaDDevelopmental origins of health and disease
ERMEzrin, radixin, myosin protein family
EBP50Ezrin-binding protein 50 kDa
EREndoplasmic reticulum
IGF1Insulin-like growth factor 1
IGF2Insulin-like growth factor 2
IP3Inositol 1,4,5-trisphosphate
NaPi-2aSodium-dependent phosphate transporter-2a SLC34A1
NCXNa+/Ca2+ exchanger
NHE3Sodium/hydrogen exchange factor3
NHERF1Na+/H+ exchange regulatory cofactor-1
Orai1Calcium release-activated calcium channel protein 1
PDZPSD-95/Discs-large/ZO1 domain PKA protein kinase A
PETPre-eclamptic toxaemia
pf.Postfertilisation (=postconception)
PIP2Phosphatidylinositol 4,5 bisphosphate
PLAPhospholipase A
PLCPhospholipase C
PMCAPlasma membrane Ca2+ ATPase
PTHParathyroid hormone
PTHrPParathyroid hormone related protein
PTH1RPTH/PTHrP 1 receptor
S100A10Annexin II light chain (calpactin light chain, p11)
S100G (S100 Calcium-Binding Protein G)Gene for calbindin-D9K
SERCASarcoendoplasmic reticulum ATPase
SOCEStore-operated Ca2+ entry
STIM1; STIM2Stromal interaction molecule 1 and 2
TRPTransient receptor potential
TRPCTransient receptor potential canonical
TRPMTransient receptor potential melastatin
TRPVTransient receptor potential vanilloid

References

  1. Kovacs, C.S. Bone development and mineral homeostasis in the fetus and neonate: Roles of the calciotropic and phosphotropic hormones. Physiol. Rev. 2014, 94, 1143–1218. [Google Scholar] [CrossRef] [PubMed]
  2. Hoffman, D.J.; Powell, T.L.; Barrett, E.S.; Hardy, D.B. Developmental origins of metabolic diseases. Physiol. Rev. 2021, 101, 739–795. [Google Scholar] [CrossRef] [PubMed] [PubMed Central]
  3. Sferruzzi-Perri, A.N.; Sandovici, I.; Constancia, M.; Fowden, A.L. Placental phenotype and the insulin-like growth factors: Resource allocation to fetal growth. J. Physiol. 2017, 595, 5057–5093. [Google Scholar] [CrossRef] [PubMed] [PubMed Central]
  4. Sandovici, I.; Georgopoulou, A.; Pérez-García, V.; Hufnagel, A.; López-Tello, J.; Lam, B.Y.H.; Schiefer, S.N.; Gaudreau, C.; Santos, F.; Hoelle, K.; et al. The imprinted Igf2-Igf2r axis is critical for matching placental microvasculature expansion to fetal growth. Dev. Cell 2022, 57, 63–79.e8. [Google Scholar] [CrossRef] [PubMed] [PubMed Central]
  5. Burton, G.J.; Jauniaux, E.; Charnock-Jones, D.S. Human early placental development: Potential roles of the endometrial glands. Placenta 2007, 28 (Suppl. A), S64–S69. [Google Scholar] [CrossRef] [PubMed] [PubMed Central]
  6. Burton, G.J.; Jauniaux, E. Placentation in the Human and Higher Primates. Adv. Anat. Embryol. Cell Biol. 2021, 234, 223–254. [Google Scholar] [CrossRef] [PubMed]
  7. Gao, H.; Wu, G.; Spencer, T.E.; Johnson, G.A.; Li, X.; Bazer, F.W. Select nutrients in the ovine uterine lumen. I. Amino acids, glucose, and ions in uterine lumenal flushings of cyclic and pregnant ewes. Biol. Reprod. 2009, 80, 86–93. [Google Scholar] [CrossRef] [PubMed]
  8. Moniz, C.F.; Nicolaides, K.H.; Tzannatos, C.; Rodeck, C.H. Calcium homeostasis in second trimester fetuses. J. Clin. Pathol. 1986, 39, 838–841. [Google Scholar] [CrossRef] [PubMed] [PubMed Central]
  9. Panda, S.; Behera, S.; Alam, M.F.; Syed, G.H. Endoplasmic reticulum & mitochondrial calcium homeostasis: The interplay with viruses. Mitochondrion 2021, 58, 227–242. [Google Scholar] [CrossRef] [PubMed]
  10. Ellery, P.M.; Cindrova-Davies, T.; Jauniaux, E.; Ferguson-Smith, A.C.; Burton, G.J. Evidence for transcriptional activity in the syncytiotrophoblast of the human placenta. Placenta 2009, 30, 329–334. [Google Scholar] [CrossRef]
  11. Cell types of the placenta. In Vascular Biology of the Placenta; Wang, Y., Zhao, S., Eds.; Chapter 4; Morgan & Claypool Life Sciences: San. Rafael, CA, USA, 20 January 2010. Available online: https://www.ncbi.nlm.nih.gov/books/NBK53245 (accessed on 29 December 2024).
  12. van Goor, M.K.C.; Hoenderop, J.G.J.; van der Wijst, J. TRP channels in calcium homeostasis: From hormonal control to structure-function relationship of TRPV5 and TRPV6. Biochim. Biophys. Acta Mol. Cell Res. 2017, 1864, 883–893. [Google Scholar] [CrossRef] [PubMed]
  13. Miura, S.; Sato, K.; Kato-Negishi, M.; Teshima, T.; Takeuchi, S. Fluid shear triggers microvilli formation via mechanosensitive activation of TRPV6. Nat. Commun. 2015, 6, 8871. [Google Scholar] [CrossRef] [PubMed] [PubMed Central]
  14. Demir, R.; Kaufmann, P.; Castellucci, M.; Erbengi, T.; Kotowski, A. Fetal vasculogenesis and angiogenesis in human placental villi. Acta Anat. 1989, 136, 190–203. [Google Scholar] [CrossRef]
  15. Fowden, A.L.; Sferruzzi-Perri, A.N.; Coan, P.M.; Constancia, M.; Burton, G.J. Placental efficiency and adaptation: Endocrine regulation. J. Physiol. 2009, 587 Pt 14, 3459–3472. [Google Scholar] [CrossRef] [PubMed] [PubMed Central]
  16. Haché, S.; Takser, L.; LeBellego, F.; Weiler, H.; Leduc, L.; Forest, J.C.; Giguère, Y.; Masse, A.; Barbeau, B.; Lafond, J. Alteration of calcium homeostasis in primary preeclamptic syncytiotrophoblasts: Effect on calcium exchange in placenta. J. Cell Mol. Med. 2011, 15, 654–667. [Google Scholar] [CrossRef] [PubMed] [PubMed Central]
  17. Gluckman, P.D.; Hanson, M.A.; Cooper, C.; Thornburg, K.L. Effect of in utero and early-life conditions on adult health and disease. N. Engl. J. Med. 2008, 359, 61–73. [Google Scholar] [CrossRef] [PubMed] [PubMed Central]
  18. Cooper, C.; Westlake, S.; Harvey, N.; Javaid, K.; Dennison, E.; Hanson, M. Review: Developmental origins of osteoporotic fracture. Osteoporos. Int. 2006, 17, 337–347. [Google Scholar] [CrossRef] [PubMed]
  19. Chacham, S.; Pasi, R.; Chegondi, M.; Ahmad, N.; Mohanty, S.B. Metabolic Bone Disease in Premature Neonates: An Unmet Challenge. J. Clin. Res. Pediatr. Endocrinol. 2020, 12, 332–339. [Google Scholar] [CrossRef] [PubMed] [PubMed Central]
  20. Lee, H.C.; Liu, J.; Profit, J.; Hintz, S.R.; Gould, J.B. Survival Without Major Morbidity Among Very Low Birth Weight Infants in California. Pediatrics 2020, 146, e20193865. [Google Scholar] [CrossRef] [PubMed] [PubMed Central]
  21. Wooding, F.B.; Morgan, G.; Jones, G.V.; Care, A.D. Calcium transport and the localisation of calbindin-D9k in the ruminant placenta during the second half of pregnancy. Cell Tissue Res. 1996, 285, 477–489. [Google Scholar] [CrossRef] [PubMed]
  22. Walker, V.; Bennet, L.; Mills, G.A.; Green, L.R.; Gnanakumaran, K.; Hanson, M.A. Effects of hypoxia on urinary organic acid and hypoxanthine excretion in fetal sheep. Pediatr. Res. 1996, 40, 309–318. [Google Scholar] [CrossRef] [PubMed]
  23. Walker, V.; Gentry, A.J.; Green, L.R.; Hanson, M.A.; Bennet, L. Effects of hypoxia on plasma amino acids of fetal sheep. Amino Acids 2000, 18, 147–156. [Google Scholar] [CrossRef] [PubMed]
  24. Rossant, J.; Cross, J.C. Placental development: Lessons from mouse mutants. Nat. Rev. Genet. 2001, 2, 538–548. [Google Scholar] [CrossRef] [PubMed]
  25. Armstrong, D.L.; McGowen, M.R.; Weckle, A.; Pantham, P.; Caravas, J.; Agnew, D.; Benirschke, K.; Savage-Rumbaugh, S.; Nevo, E.; Kim, C.J.; et al. The core transcriptome of mammalian placentas and the divergence of expression with placental shape. Placenta 2017, 57, 71–78. [Google Scholar] [CrossRef] [PubMed] [PubMed Central]
  26. Hughes, D.A.; Kircher, M.; He, Z.; Guo, S.; Fairbrother, G.L.; Moreno, C.S.; Khaitovich, P.; Stoneking, M. Evaluating intra- and inter-individual variation in the human placental transcriptome. Genome Biol. 2015, 16, 54. [Google Scholar] [CrossRef] [PubMed] [PubMed Central]
  27. Sood, R.; Zehnder, J.L.; Druzin, M.L.; Brown, P.O. Gene expression patterns in human placenta. Proc. Natl. Acad. Sci. USA 2006, 103, 5478–5483. [Google Scholar] [CrossRef] [PubMed] [PubMed Central]
  28. Liu, H.; Wang, L.; Wang, Y.; Zhu, Q.; Aldo, P.; Ding, J.; Mor, G.; Liao, A. Establishment and characterization of a new human first trimester Trophoblast cell line, AL07. Placenta 2020, 100, 122–132. [Google Scholar] [CrossRef] [PubMed] [PubMed Central]
  29. Abbas, Y.; Turco, M.Y.; Burton, G.J.; Moffett, A. Investigation of human trophoblast invasion in vitro. Hum. Reprod. Update 2020, 26, 501–513. [Google Scholar] [CrossRef] [PubMed] [PubMed Central]
  30. Huang, C.C.; Hsueh, Y.W.; Chang, C.W.; Hsu, H.C.; Yang, T.C.; Lin, W.C.; Chang, H.M. Establishment of the fetal-maternal interface: Developmental events in human implantation and placentation. Front. Cell Dev. Biol. 2023, 11, 1200330. [Google Scholar] [CrossRef] [PubMed] [PubMed Central]
  31. Hadjantonakis, A.K.; Siggia, E.D.; Simunovic, M. In vitro modeling of early mammalian embryogenesis. Curr. Opin. Biomed. Eng. 2020, 13, 134–143. [Google Scholar] [CrossRef] [PubMed] [PubMed Central]
  32. Rabussier, G.; Bünter, I.; Bouwhuis, J.; Soragni, C.; van Zijp, T.; Ng, C.P.; Domansky, K.; de Windt, L.J.; Vulto, P.; Murdoch, C.E.; et al. Healthy and diseased placental barrier on-a-chip models suitable for standardized studies. Acta Biomater. 2023, 164, 363–376. [Google Scholar] [CrossRef] [PubMed]
  33. Lermant, A.; Rabussier, G.; Lanz, H.L.; Davidson, L.; Porter, I.M.; Murdoch, C.E. Development of a human iPSC-derived placental barrier-on-chip model. iScience 2023, 26, 107240. [Google Scholar] [CrossRef] [PubMed] [PubMed Central]
  34. Cox, B.; Leavey, K.; Nosi, U.; Wong, F.; Kingdom, J. Placental transcriptome in development and pathology: Expression, function, and methods of analysis. Am. J. Obstet. Gynecol. 2015, 213 (Suppl. 4), S138–S151. [Google Scholar] [CrossRef] [PubMed]
  35. Rahnavard, A.; Chatterjee, R.; Wen, H.; Gaylord, C.; Mugusi, S.; Klatt, K.C.; Smith, E.R. Molecular epidemiology of pregnancy using omics data: Advances, success stories, and challenges. J. Transl. Med. 2024, 22, 106. [Google Scholar] [CrossRef] [PubMed] [PubMed Central]
  36. Gál, L.; Fóthi, Á.; Orosz, G.; Nagy, S.; Than, N.G.; Orbán, T.I. Exosomal small RNA profiling in first-trimester maternal blood explores early molecular pathways of preterm preeclampsia. Front. Immunol. 2024, 15, 1321191. [Google Scholar] [CrossRef] [PubMed] [PubMed Central]
  37. Stevens, A.; Khashkhusha, T.; Sharps, M.; Garner, T.; Ruane, P.T.; Aplin, J.D. The Human Early Maternal–Embryonic Interactome. Reprod. Med. 2023, 4, 40–56. [Google Scholar] [CrossRef]
  38. Jackson, M.R.; Mayhew, T.M.; Boyd, P.A. Quantitative description of the elaboration and maturation of villi from 10 weeks of gestation to term. Placenta 1992, 13, 357–370. [Google Scholar] [CrossRef]
  39. Mitchell, B.; Sharma, R. How does an embryo form? In Embryology: An Illustrated Colour Text; Mitchell, B., Sharma, R., Eds.; Elsevier/Churchill Livingstone: Amsterdam, The Netherlands, 2005; pp. 9–14. ISBN 0443073988. [Google Scholar]
  40. Harrison, R.G. The development of the placenta. In RG Harrison: A Textbook of Human Embryology, 2nd ed.; Blackwell Scientific Publications Ltd.: Oxford, UK, 1963; pp. 60–67. [Google Scholar]
  41. Sadler, T.W. Bilaminar Germ Disc (Second Week of Development). In Langman’s Medical Embryology, 15th ed.; Sadler, T.W., Ed.; Lippincott Williams & Wilkins a Wolters Kluwer Business: Alphen aan den Rijn, The Netherlands, 2023; pp. 51–59. Available online: https://integratedsciences.lwwhealthlibrary.com/book.aspx?bookid=3221&sectionid= (accessed on 29 December 2024).
  42. Burton, G.J.; Jauniaux, E. The human placenta: New perspectives on its formation and function during early pregnancy. Proc. Biol. Sci. 2023, 290, 20230191. [Google Scholar] [CrossRef] [PubMed] [PubMed Central]
  43. Herrick, E.J.; Bordoni, B. Embryology, Placenta. [Updated 2023 May 1]. In StatPearls [Internet]; StatPearls Publishing: Treasure Island, FL, USA, 2024. Available online: https://www.ncbi.nlm.nih.gov/books/NBK551634/ (accessed on 29 December 2024).
  44. Ruane, P.T.; Garner, T.; Parsons, L.; Babbington, P.A.; Wangsaputra, I.; Kimber, S.J.; Stevens, A.; Westwood, M.; Brison, D.R.; Aplin, J.D. Trophectoderm differentiation to invasive syncytiotrophoblast is promoted by endometrial epithelial cells during human embryo implantation. Hum. Reprod. 2022, 37, 777–792. [Google Scholar] [CrossRef] [PubMed] [PubMed Central]
  45. Sathasivam, R.; Selliah, P.; Sivalingarajah, R.; Mayorathan, U.; Munasinghe, B.M. Placental weight and its relationship with the birth weight of term infants and body mass index of the mothers. J. Int. Med. Res. 2023, 51, 3000605231172895. [Google Scholar] [CrossRef] [PubMed] [PubMed Central]
  46. King, B.F. The organization of actin filaments in the brush border of yolk sac epithelial cells. J. Ultrastruct. Res. 1983, 85, 329–337. [Google Scholar] [CrossRef] [PubMed]
  47. Ockleford, C.D.; Wakely, J.; Badley, R.A. Morphogenesis of human placental chorionic villi: Cytoskeletal, syncytioskeletal and extracellular matrix proteins. Proc. R. Soc. Lond. B Biol. Sci. 1981, 212, 305–316. [Google Scholar] [CrossRef] [PubMed]
  48. Vanderpuye, O.A.; Edwards, H.C.; Booth, A.G. Proteins of the human placental microvillar cytoskeleton. Alpha-Actinin. Biochem. J. 1986, 233, 351–356. [Google Scholar] [CrossRef] [PubMed] [PubMed Central]
  49. Viswanatha, R.; Bretscher, A.; Garbett, D. Dynamics of ezrin and EBP50 in regulating microvilli on the apical aspect of epithelial cells. Biochem. Soc. Trans. 2014, 42, 189–194. [Google Scholar] [CrossRef] [PubMed] [PubMed Central]
  50. Walker, V. The Intricacies of Renal Phosphate Reabsorption—An Overview. Int. J. Mol. Sci. 2024, 25, 4684. [Google Scholar] [CrossRef] [PubMed] [PubMed Central]
  51. Berryman, M.; Gary, R.; Bretscher, A. Ezrin oligomers are major cytoskeletal components of placental microvilli: A proposal for their involvement in cortical morphogenesis. J. Cell Biol. 1995, 131, 1231–1242. [Google Scholar] [CrossRef] [PubMed] [PubMed Central]
  52. Bonilha, V.L.; Finnemann, S.C.; Rodriguez-Boulan, E. Ezrin promotes morphogenesis of apical microvilli and basal infoldings in retinal pigment epithelium. J. Cell Biol. 1999, 147, 1533–1548. [Google Scholar] [CrossRef] [PubMed] [PubMed Central]
  53. Reczek, D.; Bretscher, A. The carboxyl-terminal region of EBP50 binds to a site in the amino-terminal domain of ezrin that is masked in the dormant molecule. J. Biol. Chem. 1998, 273, 18452–18458. [Google Scholar] [CrossRef] [PubMed]
  54. Li, Q.; Nance, M.R.; Kulikauskas, R.; Nyberg, K.; Fehon, R.; Karplus, P.A.; Bretscher, A.; Tesmer, J.J. Self-masking in an intact ERM-merlin protein: An active role for the central alpha-helical domain. J. Mol. Biol. 2007, 365, 1446–1459. [Google Scholar] [CrossRef] [PubMed] [PubMed Central]
  55. Turunen, O.; Wahlström, T.; Vaheri, A. Ezrin has a COOH-terminal actin-binding site that is conserved in the ezrin protein family. J. Cell Biol. 1994, 126, 1445–1453. [Google Scholar] [CrossRef] [PubMed] [PubMed Central]
  56. Reczek, D.; Berryman, M.; Bretscher, A. Identification of EBP50: A PDZ-containing phosphoprotein that associates with members of the ezrin-radixin-moesin family. J. Cell Biol. 1997, 139, 169–179. [Google Scholar] [CrossRef] [PubMed] [PubMed Central]
  57. Berryman, M.; Franck, Z.; Bretscher, A. Ezrin is concentrated in the apical microvilli of a wide variety of epithelial cells whereas moesin is found primarily in endothelial cells. J. Cell Sci. 1993, 105 Pt 4, 1025–1043. [Google Scholar] [CrossRef] [PubMed]
  58. Saotome, I.; Curto, M.; McClatchey, A.I. Ezrin is essential for epithelial organization and villus morphogenesis in the developing intestine. Dev. Cell. 2004, 6, 855–864. [Google Scholar] [CrossRef] [PubMed]
  59. Morales, F.C.; Takahashi, Y.; Kreimann, E.L.; Georgescu, M.M. Ezrin-radixin-moesin (ERM)-binding phosphoprotein 50 organizes ERM proteins at the apical membrane of polarized epithelia. Proc. Natl. Acad. Sci. USA 2004, 101, 17705–17710. [Google Scholar] [CrossRef] [PubMed] [PubMed Central]
  60. Weinman, E.J.; Shenolikar, S. The Na-H exchanger regulatory factor. Exp. Nephrol. 1997, 5, 449–452. [Google Scholar] [PubMed]
  61. Yun, C.H.; Oh, S.; Zizak, M.; Steplock, D.; Tsao, S.; Tse, C.M.; Weinman, E.J.; Donowitz, M. cAMP-mediated inhibition of the epithelial brush border Na+/H+ exchanger, NHE3, requires an associated regulatory protein. Proc. Natl. Acad. Sci. USA 1997, 94, 3010–3015. [Google Scholar] [CrossRef]
  62. Zhang, Q.; Xiao, K.; Paredes, J.M.; Mamonova, T.; Sneddon, W.B.; Liu, H.; Wang, D.; Li, S.; McGarvey, J.C.; Uehling, D.; et al. Parathyroid hormone initiates dynamic NHERF1 phosphorylation cycling and conformational changes that regulate NPT2A-dependent phosphate transport. J. Biol. Chem. 2019, 294, 4546–4571. [Google Scholar] [CrossRef] [PubMed] [PubMed Central]
  63. Bhattacharya, S.; Stanley, C.B.; Heller, W.T.; Friedman, P.A.; Bu, Z. Dynamic structure of the full-length scaffolding protein NHERF1 influences signaling complex assembly. J. Biol. Chem. 2019, 294, 11297–11310. [Google Scholar] [CrossRef] [PubMed] [PubMed Central]
  64. Hernando, N.; Wagner, C.A.; Gisler, S.M.; Biber, J.; Murer, H. PDZ proteins and proximal ion transport. Curr. Opin. Nephrol. Hypertens. 2004, 13, 569–574. [Google Scholar] [CrossRef] [PubMed]
  65. Murer, H. Functional domains in the renal type IIa Na/P(i)-cotransporter. Kidney Int. 2002, 62, 375–382. [Google Scholar] [CrossRef] [PubMed]
  66. Shenolikar, S.; Voltz, J.W.; Cunningham, R.; Weinman, E.J. Regulation of ion transport by the NHERF family of PDZ proteins. Physiology 2004, 19, 362–369. [Google Scholar] [CrossRef] [PubMed]
  67. Bootman, M.D.; Bultynck, G. Fundamentals of Cellular Calcium Signaling: A Primer. Cold Spring Harb. Perspect. Biol. 2020, 12, a038802. [Google Scholar] [CrossRef] [PubMed] [PubMed Central]
  68. Vangeel, L.; Voets, T. Transient Receptor Potential Channels and Calcium Signaling. Cold Spring Harb. Perspect. Biol. 2019, 11, a035048. [Google Scholar] [CrossRef] [PubMed] [PubMed Central]
  69. Prakriya, M.; Lewis, R.S. Store-Operated Calcium Channels. Physiol. Rev. 2015, 95, 1383–1436. [Google Scholar] [CrossRef] [PubMed] [PubMed Central]
  70. Bill, C.A.; Vines, C.M. Phospholipase C. Adv. Exp. Med. Biol. 2020, 1131, 215–242. [Google Scholar] [CrossRef] [PubMed] [PubMed Central]
  71. Rhee, S.G. Regulation of phosphoinositide-specific phospholipase C. Annu. Rev. Biochem. 2001, 70, 281–312. [Google Scholar] [CrossRef] [PubMed] [PubMed Central]
  72. Kouchi, Z.; Shikano, T.; Nakamura, Y.; Shirakawa, H.; Fukami, K.; Miyazaki, S. The role of EF-hand domains and C2 domain in regulation of enzymatic activity of phospholipase Czeta. J. Biol. Chem. 2005, 280, 21015–21021. [Google Scholar] [CrossRef] [PubMed]
  73. Wakai, T.; Mehregan, A.; Fissore, R.A. Ca2+ Signaling and Homeostasis in Mammalian Oocytes and Eggs. Cold Spring Harb. Perspect. Biol. 2019, 11, a035162. [Google Scholar] [CrossRef] [PubMed] [PubMed Central]
  74. Grasa, P.; Coward, K.; Young, C.; Parrington, J. The pattern of localization of the putative oocyte activation factor, phospholipase Czeta, in uncapacitated, capacitated, and ionophore-treated human spermatozoa. Hum. Reprod. 2008, 23, 2513–2522. [Google Scholar] [CrossRef] [PubMed]
  75. Heytens, E.; Parrington, J.; Coward, K.; Young, C.; Lambrecht, S.; Yoon, S.Y.; Fissore, R.A.; Hamer, R.; Deane, C.M.; Ruas, M.; et al. Reduced amounts and abnormal forms of phospholipase C zeta (PLCzeta) in spermatozoa from infertile men. Hum. Reprod. 2009, 24, 2417–2428. [Google Scholar] [CrossRef] [PubMed]
  76. Kashir, J.; Jones, C.; Mounce, G.; Ramadan, W.M.; Lemmon, B.; Heindryckx, B.; de Sutter, P.; Parrington, J.; Turner, K.; Child, T.; et al. Variance in total levels of phospholipase C zeta (PLC-ζ) in human sperm may limit the applicability of quantitative immunofluorescent analysis as a diagnostic indicator of oocyte activation capability. Fertil. Steril. 2013, 99, 107–117.e3. [Google Scholar] [CrossRef] [PubMed]
  77. Yoon, S.Y.; Jellerette, T.; Salicioni, A.M.; Lee, H.C.; Yoo, M.S.; Coward, K.; Parrington, J.; Grow, D.; Cibelli, J.B.; Visconti, P.E.; et al. Human sperm devoid of PLC, zeta 1 fail to induce Ca(2+) release and are unable to initiate the first step of embryo development. J. Clin. Investig. 2008, 118, 3671–3681. [Google Scholar] [CrossRef] [PubMed] [PubMed Central]
  78. Rogers, N.T.; Halet, G.; Piao, Y.; Carroll, J.; Ko, M.S.; Swann, K. The absence of a Ca(2+) signal during mouse egg activation can affect parthenogenetic preimplantation development, gene expression patterns, and blastocyst quality. Reproduction 2006, 132, 45–57. [Google Scholar] [CrossRef] [PubMed]
  79. Shafqat, A.; Kashir, J.; Alsalameh, S.; Alkattan, K.; Yaqinuddin, A. Fertilization, Oocyte Activation, Calcium Release and Epigenetic Remodelling: Lessons from Cancer Models. Front. Cell Dev. Biol. 2022, 10, 781953. [Google Scholar] [CrossRef] [PubMed] [PubMed Central]
  80. Zhang, M.; Ma, Y.; Ye, X.; Zhang, N.; Pan, L.; Wang, B. TRP (transient receptor potential) ion channel family: Structures, biological functions and therapeutic interventions for diseases. Signal Transduct. Target. Ther. 2023, 8, 261. [Google Scholar] [CrossRef] [PubMed] [PubMed Central]
  81. Earley, S.; Brayden, J.E. Transient receptor potential channels in the vasculature. Physiol. Rev. 2015, 95, 645–690. [Google Scholar] [CrossRef] [PubMed] [PubMed Central]
  82. Peng, J.-B.; Suzuki, Y.; Gyimesi, G.; Hediger, M.A. TRPV5 and TRPV6 calcium selective channels. In Calcium Entry Channels in Non-Excitable Cells; Kozak, J.A., Putney, J.W., Jr., Eds.; CRC Press/Taylor & Francis: Boca Raton, FL, USA, 2018; pp. 242–274. [Google Scholar] [CrossRef]
  83. Chen, X.; Sooch, G.; Demaree, I.S.; White, F.A.; Obukhov, A.G. Transient Receptor Potential Canonical (TRPC) Channels: Then and Now. Cells. 2020, 9, 1983. [Google Scholar] [CrossRef] [PubMed] [PubMed Central]
  84. Chen, Y.; Mu, J.; Zhu, M.; Mukherjee, A.; Zhang, H. Transient Receptor Potential Channels and Inflammatory Bowel Disease. Front. Immunol. 2020, 11, 180. [Google Scholar] [CrossRef] [PubMed] [PubMed Central]
  85. Clapham, D.E.; Julius, D.; Montell, C.; Schultz, G. International Union of Pharmacology. XLIX. Nomenclature and structure-function relationships of transient receptor potential channels. Pharmacol. Rev. 2005, 57, 427–450. [Google Scholar] [CrossRef] [PubMed]
  86. Yildirim, E.; Dietrich, A.; Birnbaumer, L. The mouse C-type transient receptor potential 2 (TRPC2) channel: Alternative splicing and calmodulin binding to its N terminus. Proc. Natl. Acad. Sci. USA 2003, 100, 2220–2225. [Google Scholar] [CrossRef] [PubMed] [PubMed Central]
  87. Venkatachalam, K.; Montell, C. TRP channels. Annu. Rev. Biochem. 2007, 76, 387–417. [Google Scholar] [CrossRef] [PubMed]
  88. Seebohm, G.; Schreiber, J.A. Beyond hot and spicy: TRPV channels and their pharmacological modulation. Cell Physiol. Biochem. 2021, 55, 108–130. [Google Scholar] [CrossRef] [PubMed]
  89. Tomohiro, D.; Mizuta, K.; Fujita, T.; Nishikubo, Y.; Kumamoto, E. Inhibition by capsaicin and its related vanilloids of compound action potentials in frog sciatic nerves. Life Sci. 2013, 92, 368–378. [Google Scholar] [CrossRef] [PubMed]
  90. Yue, L.; Peng, J.B.; Hediger, M.A.; Clapham, D.E. CaT1 manifests the pore properties of the calcium-release-activated calcium channel. Nature 2001, 410, 705–709. [Google Scholar] [CrossRef] [PubMed]
  91. Liu, C.; Xue, L.; Song, C. Calcium binding and permeation in TRPV channels: Insights from molecular dynamics simulations. J. Gen. Physiol. 2023, 155, e202213261. [Google Scholar] [CrossRef] [PubMed]
  92. De Clercq, K.; Vriens, J. Establishing life is a calcium-dependent TRiP: Transient receptor potential channels in reproduction. Biochim. Biophys. Acta Mol. Cell Res. 2018, 1865 Pt B, 1815–1829. [Google Scholar] [CrossRef]
  93. Flockerzi, V.; Nilius, B. TRPs: Truly remarkable proteins. Handb. Exp. Pharmacol. 2014, 222, 1–12. [Google Scholar] [CrossRef] [PubMed]
  94. Chubanov, V.; Köttgen, M.; Touyz, R.M.; Gudermann, T. TRPM channels in health and disease. Nat. Rev. Nephrol. 2024, 20, 175–187. [Google Scholar] [CrossRef] [PubMed]
  95. Zou, Z.G.; Rios, F.J.; Montezano, A.C.; Touyz, R.M. TRPM7, magnesium, and signaling. Int. J. Mol. Sci. 2019, 20, 1877. [Google Scholar] [CrossRef] [PubMed] [PubMed Central]
  96. Moreau, R.; Hamel, A.; Daoud, G.; Simoneau, L.; Lafond, J. Expression of calcium channels along the differentiation of cultured trophoblast cells from human term placenta. Biol. Reprod. 2002, 67, 1473–1479. [Google Scholar] [CrossRef] [PubMed]
  97. Chang, Q.; Hoefs, S.; van der Kemp, A.W.; Topala, C.N.; Bindels, R.J.; Hoenderop, J.G. The beta-glucuronidase klotho hydrolyzes and activates the TRPV5 channel. Science 2005, 310, 490–493. [Google Scholar] [CrossRef] [PubMed]
  98. Lu, P.; Boros, S.; Chang, Q.; Bindels, R.J.; Hoenderop, J.G. The beta-glucuronidase klotho exclusively activates the epithelial Ca2+ channels TRPV5 and TRPV6. Nephrol. Dial. Transplant. 2008, 23, 3397–3402. [Google Scholar] [CrossRef] [PubMed]
  99. Peng, J.B.; Brown, E.M.; Hediger, M.A. Structural conservation of the genes encoding CaT1, CaT2, and related cation channels. Genomics 2001, 76, 99–109. [Google Scholar] [CrossRef] [PubMed]
  100. Ashley, B.; Simner, C.; Manousopoulou, A.; Jenkinson, C.; Hey, F.; Frost, J.M.; Rezwan, F.I.; White, C.H.; Lofthouse, E.M.; Hyde, E.; et al. Placental uptake and metabolism of 25(OH)vitamin D determine its activity within the fetoplacental unit. eLife 2022, 11, e71094. [Google Scholar] [CrossRef] [PubMed] [PubMed Central]
  101. Wissenbach, U.; Niemeyer, B.A.; Fixemer, T.; Schneidewind, A.; Trost, C.; Cavalie, A.; Reus, K.; Meese, E.; Bonkhoff, H.; Flockerzi, V. Expression of CaT-like, a novel calcium-selective channel, correlates with the malignancy of prostate cancer. J. Biol. Chem. 2001, 276, 19461–19468. [Google Scholar] [CrossRef] [PubMed]
  102. Fecher-Trost, C.; Lux, F.; Busch, K.M.; Raza, A.; Winter, M.; Hielscher, F.; Belkacemi, T.; van der Eerden, B.; Boehm, U.; Freichel, M.; et al. Maternal Transient Receptor Potential Vanilloid 6 (Trpv6) Is Involved in Offspring Bone Development. J. Bone Miner. Res. 2019, 34, 699–710. [Google Scholar] [CrossRef] [PubMed]
  103. Cheng, W.; Sun, C.; Zheng, J. Heteromerization of TRP channel subunits: Extending functional diversity. Protein Cell. 2010, 1, 802–810. [Google Scholar] [CrossRef]
  104. Hellmich, U.A.; Gaudet, R. Structural biology of TRP channels. Handb. Exp. Pharmacol. 2014, 223, 963–990. [Google Scholar] [CrossRef] [PubMed] [PubMed Central]
  105. Rohács, T.; Lopes, C.; Michailidis, I.; Logothetis, D.E. PI(4,5)P2 regulates the activation and desensitization of TRPM8 channels through the TRP domain. Nat. Neurosci. 2005, 8, 626–634. [Google Scholar] [CrossRef]
  106. García-Sanz, N.; Fernández-Carvajal, A.; Morenilla-Palao, C.; Planells-Cases, R.; Fajardo-Sánchez, E.; Fernández-Ballester, G.; Ferrer-Montiel, A. Identification of a tetramerization domain in the C terminus of the vanilloid receptor. J. Neurosci. 2004, 24, 5307–5314. [Google Scholar] [CrossRef] [PubMed] [PubMed Central]
  107. Hofmann, T.; Obukhov, A.G.; Schaefer, M.; Harteneck, C.; Gudermann, T.; Schultz, G. Direct activation of human TRPC6 and TRPC3 channels by diacylglycerol. Nature 1999, 397, 259–263. [Google Scholar] [CrossRef] [PubMed]
  108. Froghi, S.; Grant, C.R.; Tandon, R.; Quaglia, A.; Davidson, B.; Fuller, B. New Insights on the Role of TRP Channels in Calcium Signalling and Immunomodulation: Review of Pathways and Implications for Clinical Practice. Clin. Rev. Allergy Immunol. 2021, 60, 271–292. [Google Scholar] [CrossRef] [PubMed] [PubMed Central]
  109. Khalil, M.; Alliger, K.; Weidinger, C.; Yerinde, C.; Wirtz, S.; Becker, C.; Engel, M.A. Functional Role of Transient Receptor Potential Channels in Immune Cells and Epithelia. Front. Immunol. 2018, 9, 174. [Google Scholar] [CrossRef] [PubMed] [PubMed Central]
  110. Zhuang, L.; Peng, J.B.; Tou, L.; Takanaga, H.; Adam, R.; Hediger, M.; Freeman, M. Calcium-Selective Ion Channel, CaT1, Is Apically Localized in Gastrointestinal Tract Epithelia and Is Aberrantly Expressed in Human Malignancies. Lab. Investig. 2002, 82, 1755–1764. [Google Scholar] [CrossRef]
  111. Fecher-Trost, C.; Wissenbach, U.; Weissgerber, P. TRPV6: From identification to function. Cell Calcium. 2017, 67, 116–122. [Google Scholar] [CrossRef] [PubMed]
  112. Hoenderop, J.G.; van der Kemp, A.W.; Hartog, A.; van de Graaf, S.F.; van Os, C.H.; Willems, P.H.; Bindels, R.J. Molecular identification of the apical Ca2+ channel in 1, 25-dihydroxyvitamin D3-responsive epithelia. J. Biol. Chem. 1999, 274, 8375–8378. [Google Scholar] [CrossRef] [PubMed]
  113. Peng, J.B.; Chen, X.Z.; Berger, U.V.; Vassilev, P.M.; Brown, E.M.; Hediger, M.A. A rat kidney-specific calcium transporter in the distal nephron. J. Biol. Chem. 2000, 275, 28186–28194. [Google Scholar] [CrossRef] [PubMed]
  114. Hoenderop, J.G.; Nilius, B.; Bindels, R.J. Calcium absorption across epithelia. Physiol. Rev. 2005, 85, 373–422. [Google Scholar] [CrossRef] [PubMed]
  115. Walker, V.; Vuister, G.W. Biochemistry and pathophysiology of the Transient Potential Receptor Vanilloid 6 (TRPV6) calcium channel. Adv. Clin. Chem. 2023, 113, 43–100. [Google Scholar] [CrossRef] [PubMed]
  116. Lee, S.M.; Riley, E.M.; Meyer, M.B.; Benkusky, N.A.; Plum, L.A.; DeLuca, H.F.; Pike, J.W. 1,25-Dihydroxyvitamin D3 Controls a Cohort of Vitamin D Receptor Target Genes in the Proximal Intestine That Is Enriched for Calcium-regulating Components. J. Biol. Chem. 2015, 290, 18199–18215. [Google Scholar] [CrossRef] [PubMed] [PubMed Central]
  117. Pomerantz, M.M.; Li, F.; Takeda, D.Y.; Lenci, R.; Chonkar, A.; Chabot, M.; Cejas, P.; Vazquez, F.; Cook, J.; Shivdasani, R.A.; et al. The androgen receptor cistrome is extensively reprogrammed in human prostate tumorigenesis. Nat. Genet. 2015, 47, 1346–1351. [Google Scholar] [CrossRef] [PubMed] [PubMed Central]
  118. Fecher-Trost, C.; Wissenbach, U.; Beck, A.; Schalkowsky, P.; Stoerger, C.; Doerr, J.; Dembek, A.; Simon-Thomas, M.; Weber, A.; Wollenberg, P.; et al. The in vivo TRPV6 protein starts at a non-AUG triplet, decoded as methionine, upstream of canonical initiation at AUG. J. Biol. Chem. 2013, 288, 16629–16644. [Google Scholar] [CrossRef] [PubMed] [PubMed Central]
  119. Moreau, R.; Daoud, G.; Bernatchez, R.; Simoneau, L.; Masse, A.; Lafond, J. Calcium uptake and calcium transporter expression by trophoblast cells from human term placenta. Biochim. Biophys. Acta 2002, 1564, 325–332. [Google Scholar] [CrossRef] [PubMed]
  120. Bernucci, L.; Henríquez, M.; Díaz, P.; Riquelme, G. Diverse calcium channel types are present in the human placental syncytiotrophoblast basal membrane. Placenta 2006, 27, 1082–1095. [Google Scholar] [CrossRef] [PubMed]
  121. Khattar, V.; Wang, L.; Peng, J.B. Calcium selective channel TRPV6: Structure, function, and implications in health and disease. Gene 2022, 817, 146192. [Google Scholar] [CrossRef] [PubMed] [PubMed Central]
  122. Suzuki, Y.; Kovacs, C.S.; Takanaga, H.; Peng, J.B.; Landowski, C.P.; Hediger, M.A. Calcium channel TRPV6 is involved in murine maternal-fetal calcium transport. J. Bone Miner. Res. 2008, 23, 1249–1256. [Google Scholar] [CrossRef] [PubMed] [PubMed Central]
  123. Li, S.H.; Yin, H.B.; Ren, M.R.; Wu, M.J.; Huang, X.L.; Li, J.J.; Luan, Y.P.; Wu, Y.L. TRPV5 and TRPV6 are expressed in placenta and bone tissues during pregnancy in mice. Biotech. Histochem. 2019, 94, 244–251. [Google Scholar] [CrossRef] [PubMed]
  124. Lee, B.M.; Lee, G.S.; Jung, E.M.; Choi, K.C.; Jeung, E.B. Uterine and placental expression of TRPV6 gene is regulated via progesterone receptor- or estrogen receptor-mediated pathways during pregnancy in rodents. Reprod. Biol. Endocrinol. 2009, 7, 49. [Google Scholar] [CrossRef] [PubMed] [PubMed Central]
  125. Yang, H.; An, B.S.; Choi, K.C.; Jeung, E.B. Change of genes in calcium transport channels caused by hypoxic stress in the placenta, duodenum, and kidney of pregnant rats. Biol. Reprod. 2013, 88, 30. [Google Scholar] [CrossRef] [PubMed]
  126. Sprekeler, N.; Kowalewski, M.P.; Boos, A. TRPV6 and Calbindin-D9k-expression and localization in the bovine uterus and placenta during pregnancy. Reprod. Biol. Endocrinol. 2012, 10, 66. [Google Scholar] [CrossRef] [PubMed] [PubMed Central]
  127. Kogel, A.; Fecher-Trost, C.; Wissenbach, U.; Flockerzi, V.; Schaefer, M. Ca2+ transport via TRPV6 is regulated by rapid internalization of the channel. Cell Calcium 2022, 106, 102634. [Google Scholar] [CrossRef] [PubMed]
  128. Trofimov, Y.A.; Krylov, N.A.; Minakov, A.S.; Nadezhdin, K.D.; Neuberger, A.; Sobolevsky, A.I.; Efremov, R.G. Dynamic molecular portraits of ion-conducting pores characterize functional states of TRPV channels. Commun. Chem. 2024, 7, 119. [Google Scholar] [CrossRef] [PubMed] [PubMed Central]
  129. Xin, Y.; Malick, A.; Hu, M.; Liu, C.; Batah, H.; Xu, H.; Duan, C. Cell-autonomous regulation of epithelial cell quiescence by calcium channel Trpv6. eLife 2019, 8, e48003. [Google Scholar] [CrossRef] [PubMed] [PubMed Central]
  130. Peng, J.B.; Chen, X.Z.; Berger, U.V.; Weremowicz, S.; Morton, C.C.; Vassilev, P.M.; Brown, E.M.; Hediger, M.A. Human calcium transport protein CaT1. Biochem. Biophys. Res. Commun. 2000, 278, 326–332. [Google Scholar] [CrossRef] [PubMed]
  131. Bhardwaj, R.; Lindinger, S.; Neuberger, A.; Nadezhdin, K.D.; Singh, A.K.; Cunha, M.R.; Derler, I.; Gyimesi, G.; Reymond, J.L.; Hediger, M.A.; et al. Inactivation-mimicking block of the epithelial calcium channel TRPV6. Sci. Adv. 2020, 6, eabe1508. [Google Scholar] [CrossRef] [PubMed] [PubMed Central]
  132. Yelshanskaya, M.V.; Nadezhdin, K.D.; Kurnikova, M.G.; Sobolevsky, A.I. Structure and function of the calcium-selective TRP channel TRPV6. J. Physiol. 2021, 599, 2673–2697. [Google Scholar] [CrossRef] [PubMed] [PubMed Central]
  133. McGoldrick, L.L.; Singh, A.K.; Saotome, K.; Yelshanskaya, M.V.; Twomey, E.C.; Grassucci, R.A.; Sobolevsky, A.I. Opening of the human epithelial calcium channel TRPV6. Nature 2018, 553, 233–237. [Google Scholar] [CrossRef] [PubMed] [PubMed Central]
  134. Singh, A.K.; McGoldrick, L.L.; Twomey, E.C.; Sobolevsky, A.I. Mechanism of calmodulin inactivation of the calcium-selective TRP channel TRPV6. Sci. Adv. 2018, 4, eaau6088. [Google Scholar] [CrossRef] [PubMed] [PubMed Central]
  135. Nilius, B.; Prenen, J.; Hoenderop, J.G.; Vennekens, R.; Hoefs, S.; Weidema, A.F.; Droogmans, G.; Bindels, R.J. Fast and slow inactivation kinetics of the Ca2+ channels ECaC1 and ECaC2 (TRPV5 and TRPV6). Role of the intracellular loop located between tranmembrane segments 2 and 3. J. Biol. Chem. 2002, 277, 30852–30858. [Google Scholar] [CrossRef] [PubMed]
  136. Flores-Aldama, L.; Vandewege, M.W.; Zavala, K.; Colenso, C.K.; Gonzalez, W.; Brauchi, S.E.; Opazo, J.C. Evolutionary analyses reveal independent origins of gene repertoires and structural motifs associated to fast inactivation in calcium-selective TRPV channels. Sci. Rep. 2020, 10, 8684. [Google Scholar] [CrossRef] [PubMed] [PubMed Central]
  137. Linse, S.; Helmersson, A.; Forsén, S. Calcium binding to calmodulin and its globular domains. J. Biol. Chem. 1991, 266, 8050–8054. [Google Scholar] [CrossRef] [PubMed]
  138. Pumroy, R.A.; Fluck, E.C., 3rd; Ahmed, T.; Moiseenkova-Bell, V.Y. Structural insights into the gating mechanisms of TRPV channels. Cell Calcium 2020, 87, 102168. [Google Scholar] [CrossRef] [PubMed] [PubMed Central]
  139. Andrews, C.; Xu, Y.; Kirberger, M.; Yang, J.J. Structural Aspects and Prediction of Calmodulin-Binding Proteins. Int. J. Mol. Sci. 2020, 22, 308. [Google Scholar] [CrossRef] [PubMed] [PubMed Central]
  140. Kovalevskaya, N.V.; van de Waterbeemd, M.; Bokhovchuk, F.M.; Bate, N.; Bindels, R.J.; Hoenderop, J.G.; Vuister, G.W. Structural analysis of calmodulin binding to ion channels demonstrates the role of its plasticity in regulation. Pflugers Arch. 2013, 465, 1507–1519. [Google Scholar] [CrossRef] [PubMed]
  141. Zhang, M.; Jang, H.; Gaponenko, V.; Nussinov, R. Phosphorylated Calmodulin Promotes PI3K Activation by Binding to the SH2 Domains. Biophys. J. 2017, 113, 1956–1967. [Google Scholar] [CrossRef] [PubMed] [PubMed Central]
  142. van de Graaf, S.F.; Hoenderop, J.G.; Bindels, R.J. Regulation of TRPV5 and TRPV6 by associated proteins. Am. J. Physiol. Renal Physiol. 2006, 290, F1295–F1302. [Google Scholar] [CrossRef] [PubMed]
  143. de Groot, T.; Kovalevskaya, N.V.; Verkaart, S.; Schilderink, N.; Felici, M.; van der Hagen, E.A.; Bindels, R.J.; Vuister, G.W.; Hoenderop, J.G. Molecular mechanisms of calmodulin action on TRPV5 and modulation by parathyroid hormone. Mol. Cell Biol. 2011, 31, 2845–2853. [Google Scholar] [CrossRef] [PubMed] [PubMed Central]
  144. Bokhovchuk, F.M.; Bate, N.; Kovalevskaya, N.V.; Goult, B.T.; Spronk, C.A.E.M.; Vuister, G.W. The Structural Basis of Calcium-Dependent Inactivation of the Transient Receptor Potential Vanilloid 5 Channel. Biochemistry 2018, 57, 2623–2635. [Google Scholar] [CrossRef] [PubMed]
  145. Bate, N.; Caves, R.E.; Skinner, S.P.; Goult, B.T.; Basran, J.; Mitcheson, J.S.; Vuister, G.W. A Novel Mechanism for Calmodulin-Dependent Inactivation of Transient Receptor Potential Vanilloid 6. Biochemistry 2018, 57, 2611–2622. [Google Scholar] [CrossRef] [PubMed]
  146. van de Graaf, S.F.; Hoenderop, J.G.; Gkika, D.; Lamers, D.; Prenen, J.; Rescher, U.; Gerke, V.; Staub, O.; Nilius, B.; Bindels, R.J. Functional expression of the epithelial Ca(2+) channels (TRPV5 and TRPV6) requires association of the S100A10-annexin 2 complex. EMBO J. 2003, 22, 1478–1487. [Google Scholar] [CrossRef] [PubMed] [PubMed Central]
  147. Kim, H.J.; Yang, D.K.; So, I. PDZ domain-containing protein as a physiological modulator of TRPV6. Biochem. Biophys. Res. Commun. 2007, 361, 433–438. [Google Scholar] [CrossRef] [PubMed]
  148. van de Graaf, S.F.; Chang, Q.; Mensenkamp, A.R.; Hoenderop, J.G.; Bindels, R.J. Direct interaction with Rab11a targets the epithelial Ca2+ channels TRPV5 and TRPV6 to the plasma membrane. Mol. Cell Biol. 2006, 26, 303–312. [Google Scholar] [CrossRef] [PubMed] [PubMed Central]
  149. Gkika, D.; Lemonnier, L.; Shapovalov, G.; Gordienko, D.; Poux, C.; Bernardini, M.; Bokhobza, A.; Bidaux, G.; Degerny, C.; Verreman, K.; et al. TRP channel-associated factors are a novel protein family that regulates TRPM8 trafficking and activity. J. Cell Biol. 2015, 208, 89–107. [Google Scholar] [CrossRef] [PubMed] [PubMed Central]
  150. Stumpf, T.; Zhang, Q.; Hirnet, D.; Lewandrowski, U.; Sickmann, A.; Wissenbach, U.; Dörr, J.; Lohr, C.; Deitmer, J.W.; Fecher-Trost, C. The human TRPV6 channel protein is associated with cyclophilin B in human placenta. J. Biol. Chem. 2008, 283, 18086–18098. [Google Scholar] [CrossRef] [PubMed]
  151. Fecher-Trost, C.; Weissgerber, P.; Wissenbach, U. TRPV6 channels. Handb. Exp. Pharmacol. 2014, 222, 359–384. [Google Scholar] [CrossRef] [PubMed]
  152. Raphaël, M.; Lehen’kyi, V.; Vandenberghe, M.; Beck, B.; Khalimonchyk, S.; Vanden Abeele, F.; Farsetti, L.; Germain, E.; Bokhobza, A.; Mihalache, A.; et al. TRPV6 calcium channel translocates to the plasma membrane via Orai1-mediated mechanism and controls cancer cell survival. Proc. Natl. Acad. Sci. USA 2014, 111, E3870–E3879. [Google Scholar] [CrossRef]
  153. Schindl, R.; Fritsch, R.; Jardin, I.; Frischauf, I.; Kahr, H.; Muik, M.; Riedl, M.C.; Groschner, K.; Romanin, C. Canonical transient receptor potential (TRPC) 1 acts as a negative regulator for vanilloid TRPV6-mediated Ca2+ influx. J. Biol. Chem. 2012, 287, 35612–35620. [Google Scholar] [CrossRef] [PubMed] [PubMed Central]
  154. Stenmark, H. Rab GTPases as coordinators of vesicle traffic. Nat. Rev. Mol. Cell Biol. 2009, 10, 513–525. [Google Scholar] [CrossRef] [PubMed]
  155. Burren, C.P.; Caswell, R.; Castle, B.; Welch, C.R.; Hilliard, T.N.; Smithson, S.F.; Ellard, S. TRPV6 compound heterozygous variants result in impaired placental calcium transport and severe undermineralization and dysplasia of the fetal skeleton. Am. J. Med. Genet. A 2018, 176, 1950–1955. [Google Scholar] [CrossRef] [PubMed] [PubMed Central]
  156. Suzuki, Y.; Chitayat, D.; Sawada, H.; Deardorff, M.A.; McLaughlin, H.M.; Begtrup, A.; Millar, K.; Harrington, J.; Chong, K.; Roifman, M.; et al. TRPV6 Variants Interfere with Maternal-Fetal Calcium Transport through the Placenta and Cause Transient Neonatal Hyperparathyroidism. Am. J. Hum. Genet. 2018, 102, 1104–1114. [Google Scholar] [CrossRef] [PubMed] [PubMed Central]
  157. Yamashita, S.; Mizumoto, H.; Sawada, H.; Suzuki, Y.; Hata, D. TRPV6 Gene Mutation in a Dizygous Twin with TransientNeonatal Hyperparathyroidism. J. Endocr. Soc. 2019, 3, 602–606. [Google Scholar] [CrossRef] [PubMed] [PubMed Central]
  158. Suzuki, Y.; Sawada, H.; Tokumasu, T.; Suzuki, S.; Ninomiya, S.; Shirai, M.; Mukai, T.; Saito, C.T.; Nishimura, G.; Tominaga, M. Novel TRPV6 mutations in the spectrum of transient neonatal hyperparathyroidism. J. Physiol. Sci. 2020, 70, 33. [Google Scholar] [CrossRef] [PubMed] [PubMed Central]
  159. Almidani, E.; Elsidawi, W.; Almohamedi, A.; Bin Ahmed, I.; Alfadhel, A. Case Report of Transient Neonatal Hyperparathyroidism: Medically Free Mother. Cureus 2020, 12, e7000. [Google Scholar] [CrossRef] [PubMed] [PubMed Central]
  160. Mason, A.E.; Grier, D.; Smithson, S.F.; Burren, C.P.; Gradhand, E. Post-mortem histology in transient receptor potential cation channel subfamily V member 6 (TRPV6) under-mineralising skeletal dysplasia suggests postnatal skeletal recovery: A case report. BMC Med. Genet. 2020, 21, 64. [Google Scholar] [CrossRef] [PubMed] [PubMed Central]
  161. Ohta, H.; Uenishi, K.; Shiraki, M. Recent nutritional trends of calcium and vitamin D in East Asia. Osteoporos. Sarcopenia 2016, 2, 208–213. [Google Scholar] [CrossRef] [PubMed] [PubMed Central]
  162. Shlisky, J.; Mandlik, R.; Askari, S.; Abrams, S.; Belizan, J.M.; Bourassa, M.W.; Cormick, G.; Driller-Colangelo, A.; Gomes, F.; Khadilkar, A.; et al. Calcium deficiency worldwide: Prevalence of inadequate intakes and associated health outcomes. Ann. N. Y. Acad. Sci. 2022, 1512, 10–28. [Google Scholar] [CrossRef] [PubMed] [PubMed Central]
  163. Fecher-Trost, C.; Wolske, K.; Wesely, C.; Löhr, H.; Klawitter, D.S.; Weissgerber, P.; Gradhand, E.; Burren, C.P.; Mason, A.E.; Winter, M.; et al. Mutations That Affect the Surface Expression of TRPV6 Are Associated with the Upregulation of Serine Proteases in the Placenta of an Infant. Int. J. Mol. Sci. 2021, 22, 12694. [Google Scholar] [CrossRef] [PubMed] [PubMed Central]
  164. Winter, M.; Weissgerber, P.; Klein, K.; Lux, F.; Yildiz, D.; Wissenbach, U.; Philipp, S.E.; Meyer, M.R.; Flockerzi, V.; Fecher-Trost, C. Transient Receptor Potential Vanilloid 6 (TRPV6) Proteins Control the Extracellular Matrix Structure of the Placental Labyrinth. Int. J. Mol. Sci. 2020, 21, 9674. [Google Scholar] [CrossRef] [PubMed] [PubMed Central]
  165. Vanoevelen, J.; Janssens, A.; Huitema, L.F.; Hammond, C.L.; Metz, J.R.; Flik, G.; Voets, T.; Schulte-Merker, S. Trpv5/6 is vital for epithelial calcium uptake and bone formation. FASEB J. 2011, 25, 3197–3207. [Google Scholar] [CrossRef] [PubMed]
  166. Pan, T.C.; Liao, B.K.; Huang, C.J.; Lin, L.Y.; Hwang, P.P. Epithelial Ca (2+) channel expression and Ca (2+) uptake in developing zebrafish. Am. J. Physiol. Regul. Integr. Comp. Physiol. 2005, 289, R1202–R1211. [Google Scholar] [CrossRef] [PubMed]
  167. Okura, G.C.; Bharadwaj, A.G.; Waisman, D.M. Recent Advances in Molecular and Cellular Functions of S100A10. Biomolecules 2023, 13, 1450. [Google Scholar] [CrossRef] [PubMed] [PubMed Central]
  168. Gerke, V.; Gavins, F.N.E.; Geisow, M.; Grewal, T.; Jaiswal, J.K.; Nylandsted, J.; Rescher, U. Annexins-a family of proteins with distinctive tastes for cell signaling and membrane dynamics. Nat. Commun. 2024, 15, 1574. [Google Scholar] [CrossRef] [PubMed] [PubMed Central]
  169. Abd El-Aleem, S.A.; Dekker, L.V. Assessment of the cellular localisation of the annexin A2/S100A10 complex in human placenta. J. Mol. Histol. 2018, 49, 531–543. [Google Scholar] [CrossRef] [PubMed] [PubMed Central]
  170. Kaczan-Bourgois, D.; Salles, J.P.; Hullin, F.; Fauvel, J.; Moisand, A.; Duga-Neulat, I.; Berrebi, A.; Campistron, G.; Chap, H. Increased content of annexin II (p36) and p11 in human placenta brush-border membrane vesicles during syncytiotrophoblast maturation and differentiation. Placenta 1996, 17, 669–676. [Google Scholar] [CrossRef] [PubMed]
  171. Huber, R.; Schneider, M.; Mayr, I.; Römisch, J.; Paques, E.P. The calcium binding sites in human annexin V by crystal structure analysis at 2.0 A resolution. Implications for membrane binding and calcium channel activity. FEBS Lett. 1990, 275, 15–21. [Google Scholar] [CrossRef] [PubMed]
  172. Drücker, P.; Pejic, M.; Grill, D.; Galla, H.J.; Gerke, V. Cooperative binding of annexin A2 to cholesterol- and phosphatidylinositol-4,5-bisphosphate-containing bilayers. Biophys. J. 2014, 107, 2070–2081. [Google Scholar] [CrossRef] [PubMed] [PubMed Central]
  173. Gerke, V.; Creutz, C.E.; Moss, S.E. Annexins: Linking Ca2+ signalling to membrane dynamics. Nat. Rev. Mol. Cell Biol. 2005, 6, 449–461. [Google Scholar] [CrossRef] [PubMed]
  174. Liu, Y.; Myrvang, H.K.; Dekker, L.V. Annexin A2 complexes with S100 proteins: Structure, function and pharmacological manipulation. Br. J. Pharmacol. 2015, 172, 1664–1676. [Google Scholar] [CrossRef] [PubMed] [PubMed Central]
  175. Loffing, J.; Loffing-Cueni, D.; Valderrabano, V.; Kläusli, L.; Hebert, S.C.; Rossier, B.C.; Hoenderop, J.G.; Bindels, R.J.; Kaissling, B. Distribution of transcellular calcium and sodium transport pathways along mouse distal nephron. Am. J. Physiol. Renal Physiol. 2001, 281, F1021–F1027. [Google Scholar] [CrossRef] [PubMed]
  176. Putney, J.W., Jr. A model for receptor-regulated calcium entry. Cell Calcium. 1986, 7, 1–12. [Google Scholar] [CrossRef] [PubMed]
  177. Putney, J.W.; Steinckwich-Besançon, N.; Numaga-Tomita, T.; Davis, F.M.; Desai, P.N.; D’Agostin, D.M.; Wu, S.; Bird, G.S. The functions of store-operated calcium channels. Biochim. Biophys. Acta Mol. Cell Res. 2017, 1864, 900–906. [Google Scholar] [CrossRef] [PubMed] [PubMed Central]
  178. Clarson, L.H.; Roberts, V.H.; Hamark, B.; Elliott, A.C.; Powell, T. Store-operated Ca2+ entry in first trimester and term human placenta. J. Physiol. 2003, 550 Pt 2, 515–528. [Google Scholar] [CrossRef] [PubMed] [PubMed Central]
  179. Vig, M.; Peinelt, C.; Beck, A.; Koomoa, D.L.; Rabah, D.; Koblan-Huberson, M.; Kraft, S.; Turner, H.; Fleig, A.; Penner, R.; et al. CRACM1 is a plasma membrane protein essential for store-operated Ca2+ entry. Science 2006, 312, 1220–1223. [Google Scholar] [CrossRef] [PubMed] [PubMed Central]
  180. Zhang, S.L.; Yeromin, A.V.; Zhang, X.H.; Yu, Y.; Safrina, O.; Penna, A.; Roos, J.; Stauderman, K.A.; Cahalan, M.D. Genome-wide RNAi screen of Ca(2+) influx identifies genes that regulate Ca (2+) release-activated Ca (2+) channel activity. Proc. Natl. Acad. Sci. USA 2006, 103, 9357–9362. [Google Scholar] [CrossRef] [PubMed] [PubMed Central]
  181. Feske, S.; Müller, J.M.; Graf, D.; Kroczek, R.A.; Dräger, R.; Niemeyer, C.; Baeuerle, P.A.; Peter, H.H.; Schlesier, M. Severe combined immunodeficiency due to defective binding of the nuclear factor of activated T cells in T lymphocytes of two male siblings. Eur. J. Immunol. 1996, 26, 2119–2126. [Google Scholar] [CrossRef] [PubMed]
  182. Feske, S.; Gwack, Y.; Prakriya, M.; Srikanth, S.; Puppel, S.H.; Tanasa, B.; Hogan, P.G.; Lewis, R.S.; Daly, M.; Rao, A. A mutation in Orai1 causes immune deficiency by abrogating CRAC channel function. Nature 2006, 441, 179–185. [Google Scholar] [CrossRef] [PubMed]
  183. Picard, C.; McCarl, C.A.; Papolos, A.; Khalil, S.; Lüthy, K.; Hivroz, C.; LeDeist, F.; Rieux-Laucat, F.; Rechavi, G.; Rao, A.; et al. STIM1 mutation associated with a syndrome of immunodeficiency and autoimmunity. N. Engl. J. Med. 2009, 360, 1971–1980. [Google Scholar] [CrossRef] [PubMed] [PubMed Central]
  184. Rubaiy, H.N. ORAI Calcium Channels: Regulation, Function, Pharmacology, and Therapeutic Targets. Pharmaceuticals 2023, 16, 162. [Google Scholar] [CrossRef] [PubMed] [PubMed Central]
  185. Cheng, K.T.; Liu, X.; Ong, H.L.; Swaim, W.; Ambudkar, I.S. Local Ca2+ entry via Orai1 regulates plasma membrane recruitment of TRPC1 and controls cytosolic Ca2+ signals required for specific cell functions. PLoS Biol. 2011, 9, e1001025. [Google Scholar] [CrossRef] [PubMed] [PubMed Central]
  186. Ahmad, M.; Narayanasamy, S.; Ong, H.L.; Ambudkar, I. STIM Proteins and Regulation of SOCE in ER-PM Junctions. Biomolecules 2022, 12, 1152. [Google Scholar] [CrossRef] [PubMed] [PubMed Central]
  187. Ambudkar, I.S.; de Souza, L.B.; Ong, H.L. TRPC1, Orai1, and STIM1 in SOCE: Friends in tight spaces. Cell Calcium 2017, 63, 33–39. [Google Scholar] [CrossRef] [PubMed] [PubMed Central]
  188. Lopez, J.J.; Jardin, I.; Sanchez-Collado, J.; Salido, G.M.; Smani, T.; Rosado, J.A. TRPC Channels in the SOCE Scenario. Cells 2020, 9, 126. [Google Scholar] [CrossRef] [PubMed] [PubMed Central]
  189. Lewis, R.S. Store-Operated Calcium Channels: From Function to Structure and Back Again. Cold Spring Harb. Perspect. Biol. 2020, 12, a035055. [Google Scholar] [CrossRef] [PubMed] [PubMed Central]
  190. Prakriya, M.; Lewis, R.S. Regulation of CRAC channel activity by recruitment of silent channels to a high open-probability gating mode. J. Gen. Physiol. 2006, 128, 373–386. [Google Scholar] [CrossRef] [PubMed] [PubMed Central]
  191. Sallinger, M.; Grabmayr, H.; Humer, C.; Bonhenry, D.; Romanin, C.; Schindl, R.; Derler, I. Activation mechanisms and structural dynamics of STIM proteins. J. Physiol. 2024, 602, 1475–1507. [Google Scholar] [CrossRef] [PubMed]
  192. Cao, X.; Choi, S.; Maléth, J.J.; Park, S.; Ahuja, M.; Muallem, S. The ER/PM microdomain, PI(4,5)P₂ and the regulation of STIM1-Orai1 channel function. Cell Calcium 2015, 58, 342–348. [Google Scholar] [CrossRef] [PubMed] [PubMed Central]
  193. Hogan, P.G. The STIM1-ORAI1 microdomain. Cell Calcium 2015, 58, 357–367. [Google Scholar] [CrossRef] [PubMed] [PubMed Central]
  194. Maltan, L.; Andova, A.M.; Derler, I. The Role of Lipids in CRAC Channel Function. Biomolecules 2022, 12, 352. [Google Scholar] [CrossRef] [PubMed] [PubMed Central]
  195. Ong, H.L.; de Souza, L.B.; Zheng, C.; Cheng, K.T.; Liu, X.; Goldsmith, C.M.; Feske, S.; Ambudkar, I.S. STIM2 enhances receptor-stimulated Ca²⁺ signaling by promoting recruitment of STIM1 to the endoplasmic reticulum-plasma membrane junctions. Sci. Signal. 2015, 8, ra3. [Google Scholar] [CrossRef] [PubMed] [PubMed Central]
  196. Soboloff, J.; Rothberg, B.S.; Madesh, M.; Gill, D.L. STIM proteins: Dynamic calcium signal transducers. Nat. Rev. Mol. Cell Biol. 2012, 13, 549–565. [Google Scholar] [CrossRef] [PubMed] [PubMed Central]
  197. Stathopulos, P.B.; Ikura, M. Structure and function of endoplasmic reticulum STIM calcium sensors. Curr. Top. Membr. 2013, 71, 59–93. [Google Scholar] [CrossRef] [PubMed]
  198. Hirve, N.; Rajanikanth, V.; Hogan, P.G.; Gudlur, A. Coiled-Coil Formation Conveys a STIM1 Signal from ER Lumen to Cytoplasm. Cell Rep. 2018, 22, 72–83. [Google Scholar] [CrossRef] [PubMed] [PubMed Central]
  199. Calloway, N.; Holowka, D.; Baird, B. A basic sequence in STIM1 promotes Ca2+ influx by interacting with the C-terminal acidic coiled coil of Orai1. Biochemistry 2010, 49, 1067–1071. [Google Scholar] [CrossRef] [PubMed] [PubMed Central]
  200. Carafoli, E. Calcium pump of the plasma membrane. Physiol. Rev. 1991, 71, 129–153. [Google Scholar] [CrossRef] [PubMed]
  201. Huang, G.N.; Zeng, W.; Kim, J.Y.; Yuan, J.P.; Han, L.; Muallem, S.; Worley, P.F. STIM1 carboxyl-terminus activates native SOC, I(crac) and TRPC1 channels. Nat. Cell Biol. 2006, 8, 1003–1010. [Google Scholar] [CrossRef] [PubMed]
  202. Yuan, J.P.; Zeng, W.; Huang, G.N.; Worley, P.F.; Muallem, S. STIM1 heteromultimerizes TRPC channels to determine their function as store-operated channels. Nat. Cell Biol. 2007, 9, 636–645. [Google Scholar] [CrossRef] [PubMed] [PubMed Central]
  203. Wang, Q.; Fang, Y.; Li, Y.; Liu, H.; Zhu, M.; Hu, X.; Zhou, J.; Deng, A.; Shen, B.; Chen, H. Pregnancy-Specific Beta-1-Glycoprotein 1 Increases HTR-8/SVneo Cell Migration through the Orai1/Akt Signaling Pathway. Biomolecules 2024, 14, 293. [Google Scholar] [CrossRef] [PubMed] [PubMed Central]
  204. Qin, Y.; Meng, Q.; Wang, Q.; Wu, M.; Fang, Y.; Tu, C.; Hu, X.; Shen, B.; Chen, H.; Xu, X. Pregnancy-Specific Glycoprotein 9 Enhances Store-Operated Calcium Entry and Nitric Oxide Release in Human Umbilical Vein Endothelial Cells. Diagnostics 2023, 13, 1134. [Google Scholar] [CrossRef] [PubMed] [PubMed Central]
  205. Ha, C.T.; Wu, J.A.; Irmak, S.; Lisboa, F.A.; Dizon, A.M.; Warren, J.W.; Ergun, S.; Dveksler, G.S. Human pregnancy specific beta-1-glycoprotein 1 (PSG1) has a potential role in placental vascular morphogenesis. Biol. Reprod. 2010, 83, 27–35. [Google Scholar] [CrossRef] [PubMed] [PubMed Central]
  206. Tu, C.; Tao, F.; Qin, Y.; Wu, M.; Cheng, J.; Xie, M.; Shen, B.; Ren, J.; Xu, X.; Huang, D.; et al. Serum proteins differentially expressed in early- and late-onset preeclampsia assessed using iTRAQ proteomics and bioinformatics analyses. PeerJ 2020, 8, e9753. [Google Scholar] [CrossRef] [PubMed] [PubMed Central]
  207. Abou-Kheir, W.; Barrak, J.; Hadadeh, O.; Daoud, G. HTR-8/SVneo cell line contains a mixed population of cells. Placenta 2017, 50, 1–7. [Google Scholar] [CrossRef] [PubMed]
  208. Koo, B.H.; Hwang, H.M.; Yi, B.G.; Lim, H.K.; Jeon, B.H.; Hoe, K.L.; Kwon, Y.G.; Won, M.H.; Kim, Y.M.; Berkowitz, D.E.; et al. Arginase II Contributes to the Ca2+/CaMKII/eNOS Axis by Regulating Ca2+ Concentration Between the Cytosol and Mitochondria in a p32-Dependent Manner. J. Am. Heart Assoc. 2018, 7, e009579. [Google Scholar] [CrossRef] [PubMed] [PubMed Central]
  209. Brini, M.; Carafoli, E. The plasma membrane Ca2+ ATPase and the plasma membrane sodium calcium exchanger cooperate in the regulation of cell calcium. Cold Spring Harb. Perspect. Biol. 2011, 3, a004168. [Google Scholar] [CrossRef] [PubMed] [PubMed Central]
  210. Kline, D.; Kline, J.T. Thapsigargin activates a calcium influx pathway in the unfertilized mouse egg and suppresses repetitive calcium transients in the fertilized egg. J. Biol. Chem. 1992, 267, 17624–17630. [Google Scholar] [CrossRef] [PubMed]
  211. Lawrence, Y.M.; Cuthbertson, K.S.R. Thapsigargin induces cytoplasmic free Ca2+ oscillations in mouse oocytes. Cell Calcium 1995, 17, 154–164. [Google Scholar] [CrossRef]
  212. Wakai, T.; Zhang, N.; Vangheluwe, P.; Fissore, R.A. Regulation of endoplasmic reticulum Ca2+ oscillations in mammalian eggs. J. Cell Sci. 2013, 126, 5714–5724. [Google Scholar] [CrossRef]
  213. Brini, M.; Carafoli, E. Calcium pumps in health and disease. Physiol. Rev. 2009, 89, 1341–1378. [Google Scholar] [CrossRef] [PubMed]
  214. Moreau, R.; Daoud, G.; Masse, A.; Simoneau, L.; Lafond, J. Expression and role of calcium-ATPase pump and sodium-calcium exchanger in differentiated trophoblasts from human term placenta. Mol. Reprod. Dev. 2003, 65, 283–288. [Google Scholar] [CrossRef] [PubMed]
  215. Okunade, G.W.; Miller, M.L.; Pyne, G.J.; Sutliff, R.L.; O’Connor, K.T.; Neumann, J.C.; Andringa, A.; Miller, D.A.; Prasad, V.; Doetschman, T.; et al. Targeted ablation of plasma membrane Ca2+-ATPase (PMCA) 1 and 4 indicates a major housekeeping function for PMCA1 and a critical role in hyperactivated sperm motility and male fertility for PMCA4. J. Biol. Chem. 2004, 279, 33742–33750. [Google Scholar] [CrossRef] [PubMed]
  216. Strehler, E.E.; Caride, A.J.; Filoteo, A.G.; Xiong, Y.; Penniston, J.T.; Enyedi, A. Plasma membrane Ca2+ ATPases as dynamic regulators of cellular calcium handling. Ann. N. Y. Acad. Sci. 2007, 1099, 226–236. [Google Scholar] [CrossRef] [PubMed]
  217. Jarrett, H.W.; Penniston, J.T. Purification of the Ca2+-stimulated ATPase activator from human erythrocytes. Its membership in the class of Ca2+-binding modulator proteins. J. Biol. Chem. 1978, 253, 4676–4682. [Google Scholar] [CrossRef] [PubMed]
  218. DeMarco, S.J.; Chicka, M.C.; Strehler, E.E. Plasma membrane Ca2+ ATPase isoform 2b interacts preferentially with Na+/H+ exchanger regulatory factor 2 in apical plasma membranes. J. Biol. Chem. 2002, 277, 10506–10511. [Google Scholar] [CrossRef] [PubMed]
  219. Xue, J.; Zeng, W.; Han, Y.; John, S.; Ottolia, M.; Jiang, Y. Structural mechanisms of the human cardiac sodium-calcium exchanger NCX1. Nat. Commun. 2023, 14, 6181. [Google Scholar] [CrossRef] [PubMed] [PubMed Central]
  220. Belkacemi, L.; Bédard, I.; Simoneau, L.; Lafond, J. Calcium channels, transporters and exchangers in placenta: A review. Cell Calcium 2005, 37, 1–8. [Google Scholar] [CrossRef] [PubMed]
  221. Michel, L.Y.M.; Verkaart, S.; Koopman, W.J.H.; Willems, P.H.G.M.; Hoenderop, J.G.J.; Bindels, R.J.M. Function and regulation of the Na+-Ca2+ exchanger NCX3 splice variants in brain and skeletal muscle. J. Biol. Chem. 2014, 289, 11293–11303. [Google Scholar] [CrossRef] [PubMed] [PubMed Central]
  222. Liao, J.; Li, H.; Zeng, W.; Sauer, D.B.; Belmares, R.; Jiang, Y. Structural insight into the ion-exchange mechanism of the sodium/calcium exchanger. Science 2012, 335, 686–690. [Google Scholar] [CrossRef] [PubMed]
  223. Szerencsei, R.T.; Kinjo, T.G.; Schnetkamp, P.P. The topology of the C-terminal sections of the NCX1 Na(+) /Ca(2+) exchanger and the NCKX2 Na(+)/Ca(2+)-K(+) exchanger. Channels 2013, 7, 109–114. [Google Scholar] [CrossRef] [PubMed] [PubMed Central]
  224. Sharma, V.; O’Halloran, D.M. Recent structural and functional insights into the family of sodium calcium exchangers. Genesis 2014, 52, 93–109. [Google Scholar] [CrossRef] [PubMed]
  225. DiPolo, R.; Beaugé, L. Sodium/calcium exchanger: Influence of metabolic regulation on ion carrier interactions. Physiol. Rev. 2006, 86, 155–203. [Google Scholar] [CrossRef] [PubMed]
  226. Kofuji, P.; Hadley, R.W.; Kieval, R.S.; Lederer, W.J.; Schulze, D.H. Expression of the Na-Ca exchanger in diverse tissues: A study using the cloned human cardiac Na-Ca exchanger. Am. J. Physiol. 1992, 263 Pt 1, C1241–C1249. [Google Scholar] [CrossRef] [PubMed]
  227. Moreau, R.; Simoneau, L.; Lafond, J. Calcium fluxes in human trophoblast (BeWo) cells: Calcium channels, calcium-ATPase, and sodium-calcium exchanger expression. Mol. Reprod. Dev. 2003, 64, 189–198. [Google Scholar] [CrossRef] [PubMed]
  228. King, A.J.; Siegel, M.; He, Y.; Nie, B.; Wang, J.; Koo-McCoy, S.; Minassian, N.A.; Jafri, Q.; Pan, D.; Kohler, J.; et al. Inhibition of sodium/hydrogen exchanger 3 in the gastrointestinal tract by tenapanor reduces paracellular phosphate permeability. Sci. Transl. Med. 2018, 10, eaam6474. [Google Scholar] [CrossRef] [PubMed] [PubMed Central]
  229. Sibley, C.P.; Brownbill, P.; Glazier, J.D.; Greenwood, S.L. Knowledge needed about the exchange physiology of the placenta. Placenta. 2018, 64 (Suppl. 1), S9–S15. [Google Scholar] [CrossRef] [PubMed]
  230. Bikle, D.D. Vitamin D: Production, Metabolism and Mechanisms of Action. In Endotext [Internet]; Feingold, K.R., Anawalt, B., Blackman, M.R., Boyce, A., Chrousos, G., Corpas, E., de Herder, W.W., Dhatariya, K., Dungan, K., Hofland, J., et al., Eds.; MDText.com, Inc.: South Dartmouth, MA, USA, 20 January 2000. [Google Scholar] [PubMed]
  231. Bikle, D.D.; Gee, E. Free, and not total, 1,25-dihydroxyvitamin D regulates 25- hydroxyvitamin D metabolism by keratinocytes. Endocrinology 1989, 124, 649–654. [Google Scholar] [CrossRef] [PubMed]
  232. Hoenderop, J.G.J.; van der Kemp, A.W.C.M.; Urben, C.M.; Strugnell, S.A.; Bindels, R.J.M. Effects of vitamin D compounds on renal and intestinal Ca2+ transport proteins in 25-hydroxyvitamin D3-1α-hydroxylase knockout mice. Kidney Int. 2004, 66, 1082–1089. [Google Scholar] [CrossRef]
  233. Glazier, J.D.; Atkinson, D.E.; Thornburg, K.L.; Sharpe, P.T.; Edwards, D.; Boyd, R.D.; Sibley, C.P. Gestational changes in Ca2+ transport across rat placenta and mRNA for calbindin9K and Ca(2+)-ATPase. Am. J. Physiol. 1992, 263 Pt 2, R930–R935. [Google Scholar] [CrossRef] [PubMed]
  234. Glenney, J.R., Jr.; Glenney, P. Comparison of Ca++-regulated events in the intestinal brush border. J. Cell Biol. 1985, 100, 754–763. [Google Scholar] [CrossRef]
  235. Bikle, D.D.; Munson, S.; Chafouleas, J. Calmodulin may mediate 1,25-dihydroxyvitamin D-stimulated intestinal calcium transport. FEBS Lett. 1984, 174, 30–33. [Google Scholar] [CrossRef] [PubMed]
  236. Bikle, D.D.; Munson, S. The villus gradient of brush border membrane calmodulin and the calcium-independent calmodulin-binding protein parallels that of calcium-accumulating ability. Endocrinology 1986, 118, 727–732. [Google Scholar] [CrossRef] [PubMed]
  237. Kutuzova, G.D.; Deluca, H.F. Gene expression profiles in rat intestine identify pathways for 1,25-dihydroxyvitamin D (3) stimulated calcium absorption and clarify its immunomodulatory properties. Arch. Biochem. Biophys. 2004, 432, 152–166. [Google Scholar] [CrossRef] [PubMed] [PubMed Central]
  238. Davis, W.L.; Hagler, H.K.; Jones, R.G.; Farmer, G.R.; Cooper, O.J.; Martin, J.H.; Bridges, G.E.; Goodman, D.B. Cryofixation, ultracryomicrotomy, and X-ray microanalysis of enterocytes from chick duodenum: Vitamin-D-induced formation of an apical tubulovesicular system. Anat. Rec. 1991, 229, 227–239. [Google Scholar] [CrossRef] [PubMed]
  239. Kip, S.N.; Strehler, E.E. Vitamin D3 upregulates plasma membrane Ca2+-ATPase expression and potentiates apico-basal Ca2+ flux in MDCK cells. Am. J. Physiol. Renal Physiol. 2004, 286, F363–F369. [Google Scholar] [CrossRef] [PubMed]
  240. Ryan, Z.C.; Craig, T.A.; Filoteo, A.G.; Westendorf, J.J.; Cartwright, E.J.; Neyses, L.; Strehler, E.E.; Kumar, R. Deletion of the intestinal plasma membrane calcium pump, isoform 1, Atp2b1, in mice is associated with decreased bone mineral density and impaired responsiveness to 1, 25-dihydroxyvitamin D3. Biochem. Biophys. Res. Commun. 2015, 467, 152–156. [Google Scholar] [CrossRef] [PubMed] [PubMed Central]
  241. Wasserman, R.H.; Chandler, J.S.; Meyer, S.A.; Smith, C.A.; Brindak, M.E.; Fullmer, C.S.; Penniston, J.T.; Kumar, R. Intestinal calcium transport and calcium extrusion processes at the basolateral membrane. J. Nutr. 1992, 122 (Suppl. 3), 662–671. [Google Scholar] [CrossRef] [PubMed]
  242. Lee, G.S.; Lee, K.Y.; Choi, K.C.; Ryu, Y.H.; Paik, S.G.; Oh, G.T.; Jeung, E.B. Phenotype of a calbindin-D9k gene knockout is compensated for by the induction of other calcium transporter genes in a mouse model. J. Bone Miner. Res. 2007, 22, 1968–1978. [Google Scholar] [CrossRef] [PubMed]
  243. Zheng, W.; Xie, Y.; Li, G.; Kong, J.; Feng, J.Q.; Li, Y.C. Critical role of calbindin-D28k in calcium homeostasis revealed by mice lacking both vitamin D receptor and calbindin-D28k. J. Biol. Chem. 2004, 279, 52406–52413. [Google Scholar] [CrossRef] [PubMed]
  244. Munson, S.; Wang, Y.; Chang, W.; Bikle, D.D. Myosin 1a Regulates Osteoblast Differentiation Independent of Intestinal Calcium Transport. J. Endocr. Soc. 2019, 3, 1993–2011. [Google Scholar] [CrossRef] [PubMed] [PubMed Central]
  245. Strid, H.; Powell, T.L. ATP-dependent Ca2+ transport is up-regulated during third trimester in human syncytiotrophoblast basal membranes. Pediatr. Res. 2000, 48, 58–63. [Google Scholar] [CrossRef] [PubMed]
  246. Kosk-Kosicka, D.; Zylińska, L. Protein kinase C and calmodulin effects on the plasma membrane Ca2+-ATPase from excitable and nonexcitable cells. Mol. Cell Biochem. 1997, 173, 79–87. [Google Scholar] [CrossRef] [PubMed]
  247. Jeung, E.B.; Leung, P.C.; Krisinger, J. The human calbindin-D9k gene. Complete structure and implications on steroid hormone regulation. J. Mol. Biol. 1994, 235, 1231–1238. [Google Scholar] [CrossRef] [PubMed]
  248. Choi, K.C.; Leung, P.C.; Jeung, E.B. Biology and physiology of Calbindin-D9k in female reproductive tissues: Involvement of steroids and endocrine disruptors. Reprod. Biol. Endocrinol. 2005, 3, 66. [Google Scholar] [CrossRef] [PubMed] [PubMed Central]
  249. Belkacemi, L.; Gariépy, G.; Mounier, C.; Simoneau, L.; Lafond, J. Expression of calbindin-D28k (CaBP28k) in trophoblasts from human term placenta. Biol. Reprod. 2003, 68, 1943–1950. [Google Scholar] [CrossRef] [PubMed]
  250. Sélénou, C.; Brioude, F.; Giabicani, E.; Sobrier, M.L.; Netchine, I. IGF2: Development, Genetic and Epigenetic Abnormalities. Cells 2022, 11, 1886. [Google Scholar] [CrossRef] [PubMed] [PubMed Central]
  251. Dilworth, M.R.; Kusinski, L.C.; Cowley, E.; Ward, B.S.; Husain, S.M.; Constância, M.; Sibley, C.P.; Glazier, J.D. Placental-specific Igf2 knockout mice exhibit hypocalcemia and adaptive changes in placental calcium transport. Proc. Natl. Acad. Sci. USA 2010, 107, 3894–3899. [Google Scholar] [CrossRef] [PubMed] [PubMed Central]
  252. Baral, K.; Rotwein, P. The Insulin-like Growth Factor. 2 Gene in Mammals: Organizational Complexity within a Conserved Locus. PLoS ONE 2019, 14, e0219155. [Google Scholar] [CrossRef]
  253. Sferruzzi-Perri, A.N.; Vaughan, O.R.; Coan, P.M.; Suciu, M.C.; Darbyshire, R.; Constancia, M.; Burton, G.J.; Fowden, A.L. Placental-specific Igf2 deficiency alters developmental adaptations to undernutrition in mice. Endocrinology 2011, 152, 3202–3212. [Google Scholar] [CrossRef] [PubMed]
  254. Sferruzzi-Perri, A.N.; Lopez-Tello, J.; Salazar-Petres, E. Placental adaptations supporting fetal growth during normal and adverse gestational environments. Exp. Physiol. 2023, 108, 371–397. [Google Scholar] [CrossRef] [PubMed] [PubMed Central]
  255. Gohlke, B.C.; Fahnenstich, H.; Dame, C.; Albers, N. Longitudinal data for intrauterine levels of fetal IGF-I and IGF-II. Horm. Res. 2004, 61, 200–204. [Google Scholar] [CrossRef]
  256. Oxvig, C. The role of PAPP-A in the IGF system: Location, location, location. J. Cell Commun. Signal. 2015, 9, 177–187. [Google Scholar] [CrossRef] [PubMed] [PubMed Central]
  257. Barrios, V.; Chowen, J.A.; Martín-Rivada, Á.; Guerra-Cantera, S.; Pozo, J.; Yakar, S.; Rosenfeld, R.G.; Pérez-Jurado, L.A.; Suárez, J.; Argente, J. Pregnancy-Associated Plasma Protein (PAPP)-A2 in Physiology and Disease. Cells 2021, 10, 3576. [Google Scholar] [CrossRef]
  258. Henderson, S.T.; Brierley, G.V.; Surinya, K.H.; Priebe, I.K.; Catcheside, D.E.; Wallace, J.C.; Forbes, B.E.; Cosgrove, L.J. Delineation of the IGF-II C domain elements involved in binding and activation of the IR-A, IR-B and IGF-IR. Growth Horm. IGF Res. 2015, 25, 20–27. [Google Scholar] [CrossRef] [PubMed]
  259. Bergman, D.; Halje, M.; Nordin, M.; Engström, W. Insulin-like growth factor 2 in development and disease: A mini-review. Gerontology 2013, 59, 240–249. [Google Scholar] [CrossRef] [PubMed]
  260. Chao, W.; D’Amore, P.A. IGF2: Epigenetic regulation and role in development and disease. Cytokine Growth Factor. Rev. 2008, 19, 111–120. [Google Scholar] [CrossRef] [PubMed] [PubMed Central]
  261. Matthews, J.C.; Beveridge, M.J.; Dialynas, E.; Bartke, A.; Kilberg, M.S.; Novak, D.A. Placental anionic and cationic amino acid transporter expression in growth hormone overexpressing and null IGF-II or null IGF-I receptor mice. Placenta 1999, 20, 639–650. [Google Scholar] [CrossRef] [PubMed]
  262. Hayward, C.E.; McIntyre, K.R.; Sibley, C.P.; Greenwood, S.L.; Dilworth, M.R. Mechanisms Underpinning Adaptations in Placental Calcium Transport in Normal Mice and Those with Fetal Growth Restriction. Front. Endocrinol. 2018, 9, 671. [Google Scholar] [CrossRef] [PubMed] [PubMed Central]
  263. Abi Habib, W.; Brioude, F.; Azzi, S.; Salem, J.; Das Neves, C.; Personnier, C.; Chantot-Bastaraud, S.; Keren, B.; Le Bouc, Y.; Harbison, M.D.; et al. 11p15 ICR1 Partial Deletions Associated with IGF2/H19 DMR Hypomethylation and Silver-Russell Syndrome. Hum. Mutat. 2017, 38, 105–111. [Google Scholar] [CrossRef] [PubMed]
  264. Yamazawa, K.; Kagami, M.; Nagai, T.; Kondoh, T.; Onigata, K.; Maeyama, K.; Hasegawa, T.; Hasegawa, Y.; Yamazaki, T.; Mizuno, S.; et al. Molecular and clinical findings and their correlations in Silver-Russell syndrome: Implications for a positive role of IGF2 in growth determination and differential imprinting regulation of the IGF2-H19 domain in bodies and placentas. J. Mol. Med. 2008, 86, 1171–1181. [Google Scholar] [CrossRef] [PubMed]
  265. Gaillot-Durand, L.; Brioude, F.; Beneteau, C.; Le Breton, F.; Massardier, J.; Michon, L.; Devouassoux-Shisheboran, M.; Allias, F. Placental Pathology in Beckwith-Wiedemann Syndrome According to Genotype/Epigenotype Subgroups. Fetal Pediatr. Pathol. 2018, 37, 387–399. [Google Scholar] [CrossRef] [PubMed]
  266. Armes, J.E.; McGown, I.; Williams, M.; Broomfield, A.; Gough, K.; Lehane, F.; Lourie, R. The placenta in Beckwith-Wiedemann syndrome: Genotype-phenotype associations, excessive extravillous trophoblast and placental mesenchymal dysplasia. Pathology 2012, 44, 519–527. [Google Scholar] [CrossRef] [PubMed]
  267. Azzi, S.; Abi Habib, W.; Netchine, I. Beckwith-Wiedemann and Russell-Silver Syndromes: From new molecular insights to the comprehension of imprinting regulation. Curr. Opin. Endocrinol. Diabetes Obes. 2014, 21, 30–38. [Google Scholar] [CrossRef]
  268. Duval, C.; Dilworth, M.R.; Tunster, S.J.; Kimber, S.J.; Glazier, J.D. PTHrP is essential for normal morphogenetic and functional development of the murine placenta. Dev. Biol. 2017, 430, 325–336. [Google Scholar] [CrossRef] [PubMed]
  269. Philbrick, W.M.; Wysolmerski, J.J.; Galbraith, S.; Holt, E.; Orloff, J.J.; Yang, K.H.; Vasavada, R.C.; Weir, E.C.; Broadus, A.E.; Stewart, A.F. Defining the roles of parathyroid hormone-related protein in normal physiology. Physiol. Rev. 1996, 76, 127–173. [Google Scholar] [CrossRef] [PubMed]
  270. Wysolmerski, J.J.; Stewart, A.F. The physiology of parathyroid hormone-related protein: An emerging role as a developmental factor. Annu. Rev. Physiol. 1998, 60, 431–460. [Google Scholar] [CrossRef] [PubMed]
  271. Maeda, S.; Sutliff, R.L.; Qian, J.; Lorenz, J.N.; Wang, J.; Tang, H.; Nakayama, T.; Weber, C.; Witte, D.; Strauch, A.R.; et al. Targeted overexpression of parathyroid hormone related protein (PTHrP) to vascular smooth muscle in transgenic mice lowers blood pressure and alters vascular contractility. Endocrinology 1999, 140, 1815–1825. [Google Scholar] [CrossRef]
  272. Bruns, M.E.; Ferguson, J.E., 2nd; Bruns, D.E.; Burton, D.W.; Brandt, D.W.; Jüppner, H.; Segre, G.V.; Deftos, L.J. Expression of parathyroid hormone-related peptide and its receptor messenger ribonucleic acid in human amnion and chorion-decidua: Implications for secretion and function. Am. J. Obstet. Gynecol. 1995, 173 Pt 1, 739–746. [Google Scholar] [CrossRef] [PubMed]
  273. Wysolmerski, J.J. Parathyroid hormone-related protein: An update. J. Clin. Endocrinol. Metab. 2012, 97, 2947–2956. [Google Scholar] [CrossRef] [PubMed] [PubMed Central]
  274. Librizzi, M.; Naselli, F.; Abruscato, G.; Luparello, C.; Caradonna, F. Parathyroid Hormone Related Protein (PTHrP)-Associated Molecular Signatures in Tissue Differentiation and Non-Tumoral Diseases. Biology 2023, 12, 950. [Google Scholar] [CrossRef] [PubMed] [PubMed Central]
  275. Pioszak, A.A.; Parker, N.R.; Gardella, T.J.; Xu, H.E. Structural basis for parathyroid hormone-related protein binding to the parathyroid hormone receptor and design of conformation-selective peptides. J. Biol. Chem. 2009, 284, 28382–28391. [Google Scholar] [CrossRef] [PubMed] [PubMed Central]
  276. Care, A.D.; Abbas, S.K.; Pickard, D.W.; Barri, M.; Drinkhill, M.; Findlay, J.B.; White, I.R.; Caple, I.W. Stimulation of ovine placental transport of calcium and magnesium by mid-molecule fragments of human parathyroid hormone-related protein. Exp. Physiol. 1990, 75, 605–608. [Google Scholar] [CrossRef] [PubMed]
  277. Kovacs, C.S.; Lanske, B.; Hunzelman, J.L.; Guo, J.; Karaplis, A.C.; Kronenberg, H.M. Parathyroid hormone-related peptide (PTHrP) regulates fetal-placental calcium transport through a receptor distinct from the PTH/PTHrP receptor. Proc. Natl. Acad. Sci. USA 1996, 93, 15233–15238. [Google Scholar] [CrossRef] [PubMed] [PubMed Central]
  278. Jans, D.A.; Thomas, R.J.; Gillespie, M.T. Parathyroid hormone-related protein (PTHrP): A nucleocytoplasmic shuttling protein with distinct paracrine and intracrine roles. Vitam. Horm. 2003, 66, 345–384. [Google Scholar] [CrossRef] [PubMed]
  279. Vilardaga, J.P.; Clark, L.J.; White, A.D.; Sutkeviciute, I.; Lee, J.Y.; Bahar, I. Molecular Mechanisms of PTH/PTHrP Class B GPCR Signaling and Pharmacological Implications. Endocr. Rev. 2023, 44, 474–491. [Google Scholar] [CrossRef] [PubMed] [PubMed Central]
  280. Klenk, C.; Hommers, L.; Lohse, M.J. Proteolytic Cleavage of the Extracellular Domain Affects Signaling of Parathyroid Hormone 1 Receptor. Front. Endocrinol. 2022, 13, 839351. [Google Scholar] [CrossRef] [PubMed] [PubMed Central]
  281. Cary, B.P.; Zhang, X.; Cao, J.; Johnson, R.M.; Piper, S.J.; Gerrard, E.J.; Wootten, D.; Sexton, P.M. New Insights into the Structure and Function of Class B1 GPCRs. Endocr. Rev. 2023, 44, 492–517. [Google Scholar] [CrossRef] [PubMed] [PubMed Central]
  282. Wehbi, V.L.; Stevenson, H.P.; Feinstein, T.N.; Calero, G.; Romero, G.; Vilardaga, J.P. Noncanonical GPCR signaling arising from a PTH receptor-arrestin-Gβγ complex. Proc. Natl. Acad. Sci. USA 2013, 110, 1530–1535. [Google Scholar] [CrossRef] [PubMed] [PubMed Central]
  283. El-Hashash, A.H.; Warburton, D.; Kimber, S.J. Genes and signals regulating murine trophoblast cell development. Mech. Dev. 2010, 127, 1–20. [Google Scholar] [CrossRef] [PubMed] [PubMed Central]
  284. Guo, L.; Qi, S.T.; Miao, D.Q.; Liang, X.W.; Li, H.; Ou, X.H.; Huang, X.; Yang, C.R.; Ouyang, Y.C.; Hou, Y.; et al. The roles of parathyroid hormone-like hormone during mouse preimplantation embryonic development. PLoS ONE 2012, 7, e40528. [Google Scholar] [CrossRef] [PubMed] [PubMed Central]
  285. Macgill, K.; Moseley, J.M.; Martin, T.J.; Brennecke, S.P.; Rice, G.E.; Wlodek, M.E. Vascular effects of PTHrP (1-34) and PTH (1-34) in the human fetal-placental circulation. Placenta 1997, 18, 587–592. [Google Scholar] [CrossRef] [PubMed]
  286. Bond, H.; Dilworth, M.R.; Baker, B.; Cowley, E.; Requena Jimenez, A.; Boyd, R.D.; Husain, S.M.; Ward, B.S.; Sibley, C.P.; Glazier, J.D. Increased maternofetal calcium flux in parathyroid hormone-related protein-null mice. J. Physiol. 2008, 586, 2015–2025. [Google Scholar] [CrossRef] [PubMed] [PubMed Central]
  287. Kovacs, C.S.; Chafe, L.L.; Fudge, N.J.; Friel, J.K.; Manley, N.R. PTH regulates fetal blood calcium and skeletal mineralization independently of PTHrP. Endocrinology 2001, 142, 4983–4993. [Google Scholar] [CrossRef] [PubMed]
  288. Karaplis, A.C.; Luz, A.; Glowacki, J.; Bronson, R.T.; Tybulewicz, V.L.; Kronenberg, H.M.; Mulligan, R.C. Lethal skeletal dysplasia from targeted disruption of the parathyroid hormone-related peptide gene. Genes Dev. 1994, 8, 277–289. [Google Scholar] [CrossRef] [PubMed]
  289. Lanske, B.; Kronenberg, H.M. Parathyroid hormone-related peptide (PTHrP) and parathyroid hormone (PTH)/PTHrP receptor. Crit. Rev. Eukaryot. Gene Expr. 1998, 8, 297–320. [Google Scholar] [CrossRef] [PubMed]
  290. Sirico, A.; Dell’Aquila, M.; Tartaglione, L.; Moresi, S.; Farì, G.; Pitocco, D.; Arena, V.; Lanzone, A. PTH-rP and PTH-R1 Expression in Placentas from Pregnancies Complicated by Gestational Diabetes: New Insights into the Pathophysiology of Hyperglycemia in Pregnancy. Diagnostics 2021, 11, 1356. [Google Scholar] [CrossRef] [PubMed] [PubMed Central]
  291. Miao, D.; Su, H.; He, B.; Gao, J.; Xia, Q.; Zhu, M.; Gu, Z.; Goltzman, D.; Karaplis, A.C. Severe growth retardation and early lethality in mice lacking the nuclear localization sequence and C-terminus of PTH-related protein. Proc. Natl. Acad. Sci. USA 2008, 105, 20309–20314. [Google Scholar] [CrossRef] [PubMed] [PubMed Central]
  292. Anderson, J.A.; Grabowska, A.M.; Watson, S.A. PTHrP increases transcriptional activity of the integrin subunit alpha5. Br. J. Cancer 2007, 96, 1394–1403. [Google Scholar] [CrossRef] [PubMed] [PubMed Central]
  293. Bhatia, V.; Mula, R.V.; Falzon, M. Parathyroid hormone-related protein regulates integrin α6 and β4 levels via transcriptional and post-translational pathways. Exp. Cell Res. 2013, 319, 1419–1430. [Google Scholar] [CrossRef] [PubMed] [PubMed Central]
  294. Gorvin, C.M. Molecular and clinical insights from studies of calcium-sensing receptor mutations. J. Mol. Endocrinol. 2019, 63, R1–R16. [Google Scholar] [CrossRef] [PubMed]
  295. Conigrave, A.D. The Calcium-Sensing Receptor and the Parathyroid: Past, Present, Future. Front. Physiol. 2016, 7, 563. [Google Scholar] [CrossRef] [PubMed] [PubMed Central]
  296. Chang, W.; Tu, C.; Chen, T.H.; Bikle, D.; Shoback, D. The extracellular calcium-sensing receptor (CaSR) is a critical modulator of skeletal development. Sci. Signal. 2008, 1, ra1. [Google Scholar] [CrossRef]
  297. Geng, Y.; Mosyak, L.; Kurinov, I.; Zuo, H.; Sturchler, E.; Cheng, T.C.; Subramanyam, P.; Brown, A.P.; Brennan, S.C.; Mun, H.C.; et al. Structural mechanism of ligand activation in human calcium-sensing receptor. eLife 2016, 5, e13662. [Google Scholar] [CrossRef]
  298. Zhang, C.; Zhang, T.; Zou, J.; Miller, C.L.; Gorkhali, R.; Yang, J.Y.; Schilmiller, A.; Wang, S.; Huang, K.; Brown, E.M.; et al. 2016 Structural basis for regulation of human calcium-sensing receptor by magnesium ions and an unexpected tryptophan derivative co-agonist. Sci. Adv. 2016, 2, e1600241. [Google Scholar] [CrossRef] [PubMed]
  299. Kovacs, C.S.; Ho-Pao, C.L.; Hunzelman, J.L.; Lanske, B.; Fox, J.; Seidman, J.G.; Seidman, C.E.; Kronenberg, H.M. Regulation of murine fetal-placental calcium metabolism by the calcium-sensing receptor. J. Clin. Investig. 1998, 101, 2812–2820. [Google Scholar] [CrossRef] [PubMed] [PubMed Central]
  300. Kovacs, C.S. The Role of PTHrP in Regulating Mineral Metabolism During Pregnancy, Lactation, and Fetal/Neonatal Development. Clinic Rev. Bone Miner. Metab. 2014, 12, 142–164. [Google Scholar] [CrossRef]
  301. Sadacharan, D.; Mahadevan, S.; Rao, S.S.; Kumar, A.P.; Swathi, S.; Kumar, S.; Kannan, S. Neonatal Severe Primary Hyperparathyroidism: A Series of Four Cases and their Long-term Management in India. Indian. J. Endocrinol. Metab. 2020, 24, 196–201. [Google Scholar] [CrossRef] [PubMed] [PubMed Central]
  302. Gorvin, C.M.; Rogers, A.; Hastoy, B.; Tarasov, A.I.; Frost, M.; Sposini, S.; Inoue, A.; Whyte, M.P.; Rorsman, P.; Hanyaloglu, A.C.; et al. AP2σ Mutations Impair Calcium-Sensing Receptor Trafficking and Signaling, and Show an Endosomal Pathway to Spatially Direct G-Protein Selectivity. Cell Rep. 2018, 22, 1054–1066. [Google Scholar] [CrossRef] [PubMed] [PubMed Central]
  303. Ganor, Y.; Drillet-Dangeard, A.S.; Lopalco, L.; Tudor, D.; Tambussi, G.; Delongchamps, N.B.; Zerbib, M.; Bomsel, M. Calcitonin gene-related peptide inhibits Langerhans cell-mediated HIV-1 transmission. J. Exp. Med. 2013, 210, 2161–2170. [Google Scholar] [CrossRef] [PubMed] [PubMed Central]
  304. Schifter, S. Expression of the calcitonin gene family in medullary thyroid carcinoma. Peptides 1997, 18, 307–317. [Google Scholar] [CrossRef] [PubMed]
  305. Poyner, D.R.; Sexton, P.M.; Marshall, I.; Smith, D.M.; Quirion, R.; Born, W.; Muff, R.; Fischer, J.A.; Foord, S.M. International Union of Pharmacology. XXXII. The mammalian calcitonin gene-related peptides, adrenomedullin, amylin, and calcitonin receptors. Pharmacol. Rev. 2002, 54, 233–246. [Google Scholar] [CrossRef] [PubMed]
  306. Yallampalli, C.; Chauhan, M.; Endsley, J.; Sathishkumar, K. Calcitonin gene related family peptides: Importance in normal placental and fetal development. Adv. Exp. Med. Biol. 2014, 814, 229–240. [Google Scholar] [CrossRef] [PubMed]
  307. Hay, D.L.; Garelja, M.L.; Poyner, D.R.; Walker, C.S. Update on the pharmacology of calcitonin/CGRP family of peptides: IUPHAR Review 25. Br. J. Pharmacol. 2018, 175, 3–17. [Google Scholar] [CrossRef] [PubMed] [PubMed Central]
  308. Granstein, R.D.; Wagner, J.A.; Stohl, L.L.; Ding, W. Calcitonin gene-related peptide: Key regulator of cutaneous immunity. Acta Physiol. 2015, 213, 586–594. [Google Scholar] [CrossRef] [PubMed] [PubMed Central]
  309. Kovacs, C.S.; Chafe, L.L.; Woodland, M.L.; McDonald, K.R.; Fudge, N.J.; Wookey, P.J. Calcitropic gene expression suggests a role for the intraplacental yolk sac in maternal-fetal calcium exchange. Am. J. Physiol. Endocrinol. Metab. 2002, 282, E721–E732. [Google Scholar] [CrossRef] [PubMed]
  310. Kumar, S.; Zhu, L.J.; Polihronis, M.; Cameron, S.T.; Baird, D.T.; Schatz, F.; Dua, A.; Ying, Y.K.; Bagchi, M.K.; Bagchi, I.C. Progesterone induces calcitonin gene expression in human endometrium within the putative window of implantation. J. Clin. Endocrinol. Metab. 1998, 83, 4443–4450. [Google Scholar] [CrossRef] [PubMed]
  311. Zhu, L.J.; Bagchi, M.K.; Bagchi, I.C. Attenuation of calcitonin gene expression in pregnant rat uterus leads to a block in embryonic implantation. Endocrinology 1998, 139, 330–339. [Google Scholar] [CrossRef] [PubMed]
  312. Kuestner, R.E.; Elrod, R.D.; Grant, F.J.; Hagen, F.S.; Kuijper, J.L.; Matthewes, S.L.; O’Hara, P.J.; Sheppard, P.O.; Stroop, S.D.; Thompson, D.L.; et al. Cloning and characterization of an abundant subtype of the human calcitonin receptor. Mol. Pharmacol. 1994, 46, 246–255. [Google Scholar] [PubMed]
  313. Gao, F.; Guo, Y.; Chen, X.; Gu, Q.; Huang, S.; Chen, Q.; Xu, X.; Zeng, K.; Zhou, H.; Zou, Y.; et al. DNA Methylation Pattern of CALCA and CALCB in Extremely Premature Infants with Monochorionic Triplets after Single-Embryo Transfer. Oxid. Med. Cell Longev. 2021, 2021, 1438837. [Google Scholar] [CrossRef] [PubMed] [PubMed Central]
  314. Dong, Y.L.; Vegiraju, S.; Gangula, P.R.; Kondapaka, S.B.; Wimalawansa, S.J.; Yallampalli, C. Expression and regulation of calcitonin gene-related Peptide receptor in rat placentas. Biol. Reprod. 2002, 67, 1321–1326. [Google Scholar] [CrossRef] [PubMed]
  315. Brain, S.D.; Williams, T.J.; Tippins, J.R.; Morris, H.R.; MacIntyre, I. Calcitonin gene-related peptide is a potent vasodilator. Nature 1985, 313, 54–56. [Google Scholar] [CrossRef] [PubMed]
  316. Walker, C.S.; Conner, A.C.; Poyner, D.R.; Hay, D.L. Regulation of signal transduction by calcitonin gene-related peptide receptors. Trends Pharmacol. Sci. 2010, 31, 476–483. [Google Scholar] [CrossRef] [PubMed]
  317. McLatchie, L.M.; Fraser, N.J.; Main, M.J.; Wise, A.; Brown, J.; Thompson, N.; Solari, R.; Lee, M.G.; Foord, S.M. RAMPs regulate the transport and ligand specificity of the calcitonin-receptor-like receptor. Nature 1998, 393, 333–339. [Google Scholar] [CrossRef] [PubMed]
  318. McDonald, K.R.; Fudge, N.J.; Woodrow, J.P.; Friel, J.K.; Hoff, A.O.; Gagel, R.F.; Kovacs, C.S. Ablation of calcitonin/calcitonin gene-related peptide-alpha impairs fetal magnesium but not calcium homeostasis. Am. J. Physiol. Endocrinol. Metab. 2004, 287, E218–E226. [Google Scholar] [CrossRef] [PubMed]
  319. Ferioli, S.; Zierler, S.; Zaißerer, J.; Schredelseker, J.; Gudermann, T.; Chubanov, V. TRPM6 and TRPM7 differentially contribute to the relief of heteromeric TRPM6/7 channels from inhibition by cytosolic Mg2+ and Mg·ATP. Sci. Rep. 2017, 7, 8806. [Google Scholar] [CrossRef] [PubMed] [PubMed Central]
  320. Nadezhdin, K.D.; Correia, L.; Narangoda, C.; Patel, D.S.; Neuberger, A.; Gudermann, T.; Kurnikova, M.G.; Chubanov, V.; Sobolevsky, A.I. Structural mechanisms of TRPM7 activation and inhibition. Nat. Commun. 2023, 14, 2639. [Google Scholar] [CrossRef] [PubMed] [PubMed Central]
  321. Walder, R.Y.; Yang, B.; Stokes, J.B.; Kirby, P.A.; Cao, X.; Shi, P.; Searby, C.C.; Husted, R.F.; Sheffield, V.C. Mice defective in Trpm6 show embryonic mortality and neural tube defects. Hum. Mol. Genet. 2009, 18, 4367–4375. [Google Scholar] [CrossRef] [PubMed] [PubMed Central]
  322. Bennin, D.; Hartery, S.A.; Kirby, B.J.; Maekawa, A.S.; St-Arnaud, R.; Kovacs, C.S. Loss of 24-hydroxylated catabolism increases calcitriol and fibroblast growth factor 23 and alters calcium and phosphate metabolism in fetal mice. JBMR Plus 2024, 8, ziae012. [Google Scholar] [CrossRef] [PubMed] [PubMed Central]
  323. Delvin, E.E.; Glorieux, F.H.; Salle, B.L.; David, L.; Varenne, J.P. Control of vitamin D metabolism in preterm infants: Feto-maternal relationships. Arch. Dis. Child. 1982, 57, 754–757. [Google Scholar] [CrossRef] [PubMed] [PubMed Central]
  324. Dalgård, C.; Petersen, M.S.; Steuerwald, U.; Weihe, P.; Grandjean, P. Umbilical Cord Serum 25-Hydroxyvitamin D Concentrations and Relation to Birthweight, Head Circumference and Infant Length at Age 14 Days. Paediatr. Perinat. Epidemiol. 2016, 30, 238–245. [Google Scholar] [CrossRef] [PubMed] [PubMed Central]
  325. Walker, V.P.; Zhang, X.; Rastegar, I.; Liu, P.T.; Hollis, B.W.; Adams, J.S.; Modlin, R.L. Cord blood vitamin D status impacts innate immune responses. J. Clin. Endocrinol. Metab. 2011, 96, 1835–1843. [Google Scholar] [CrossRef] [PubMed] [PubMed Central]
  326. Bikle, D.D. Vitamin D metabolism, mechanism of action, and clinical applications. Chem. Biol. 2014, 21, 319–329. [Google Scholar] [CrossRef] [PubMed] [PubMed Central]
  327. Chun, R.F.; Shieh, A.; Gottlieb, C.; Yacoubian, V.; Wang, J.; Hewison, M.; Adams, J.S. Vitamin D Binding Protein and the Biological Activity of Vitamin, D. Front. Endocrinol. 2019, 10, 718. [Google Scholar] [CrossRef] [PubMed] [PubMed Central]
  328. Simner, C.L.; Ashley, B.; Cooper, C.; Harvey, N.C.; Lewis, R.M.; Cleal, J.K. Investigating a suitable model for the study of vitamin D mediated regulation of human placental gene expression. J. Steroid Biochem. Mol. Biol. 2020, 199, 105576. [Google Scholar] [CrossRef] [PubMed] [PubMed Central]
  329. Nykjaer, A.; Dragun, D.; Walther, D.; Vorum, H.; Jacobsen, C.; Herz, J.; Melsen, F.; Christensen, E.I.; Willnow, T.E. An endocytic pathway essential for renal uptake and activation of the steroid 25-(OH) vitamin D3. Cell 1999, 96, 507–515. [Google Scholar] [CrossRef] [PubMed]
  330. Sheikh-Hamad, D.; Holliday, M.; Li, Q. Megalin-Mediated Trafficking of Mitochondrial Intracrines: Relevance to Signaling and Metabolism. J. Cell Immunol. 2021, 3, 364–369. [Google Scholar] [PubMed] [PubMed Central]
  331. Ockleford, C.D.; Whyte, A. Differeniated regions of human placental cell surface associated with exchange of materials between maternal and foetal blood: Coated vesicles. J. Cell Sci. 1977, 25, 293–312. [Google Scholar] [CrossRef] [PubMed]
  332. King, B.F. Absorption of peroxidase-conjugated immunoglobulin G by human placenta: An in vitro study. Placenta 1982, 3, 395–406. [Google Scholar] [CrossRef] [PubMed]
  333. King, B.F. The organization of actin filaments in human placental villi. J. Ultrastruct. Res. 1983, 85, 320–328. [Google Scholar] [CrossRef] [PubMed]
  334. Storm, T.; Christensen, E.I.; Christensen, J.N.; Kjaergaard, T.; Uldbjerg, N.; Larsen, A.; Honoré, B.; Madsen, M. Megalin Is Predominantly Observed in Vesicular Structures in First and Third Trimester Cytotrophoblasts of the Human Placenta. J. Histochem. Cytochem. 2016, 64, 769–784. [Google Scholar] [CrossRef] [PubMed] [PubMed Central]
  335. Wu, S.; Chun, R.; Gacad, M.A.; Ren, S.; Chen, H.; Adams, J.S. Regulation of 1,25-dihydroxyvitamin d synthesis by intracellular vitamin d binding protein-1. Endocrinology 2002, 143, 4135. [Google Scholar] [CrossRef] [PubMed]
  336. Carlberg, C. Vitamin D and Its Target Genes. Nutrients 2022, 14, 1354. [Google Scholar] [CrossRef]
  337. Allen, B.L.; Taatjes, D.J. The Mediator complex: A central integrator of transcription. Nat. Rev. Mol. Cell Biol. 2015, 16, 155–166. [Google Scholar] [CrossRef] [PubMed] [PubMed Central]
  338. Levine, M.A. Diagnosis and Management of Vitamin D Dependent Rickets. Front. Pediatr. 2020, 8, 315. [Google Scholar] [CrossRef] [PubMed] [PubMed Central]
  339. Fortin, C.A.; Girard, L.; Bonenfant, C.; Leblanc, J.; Cruz-Marino, T.; Blackburn, M.E.; Desmeules, M.; Bouchard, L. Benefits of Newborn Screening for Vitamin D-Dependant Rickets Type 1A in a Founder Population. Front. Endocrinol. (Lausanne) 2022, 13, 887371. [Google Scholar] [CrossRef] [PubMed] [PubMed Central]
  340. Schlingmann, K.P.; Kaufmann, M.; Weber, S.; Irwin, A.; Goos, C.; John, U.; Misselwitz, J.; Klaus, G.; Kuwertz-Bröking, E.; Fehrenbach, H.; et al. Mutations in CYP24A1 and idiopathic infantile hypercalcemia. N. Engl. J. Med. 2011, 365, 410–421. [Google Scholar] [CrossRef] [PubMed]
  341. Varshney, S.; Adela, R.; Kachhawa, G.; Dada, R.; Kulshreshtha, V.; Kumari, R.; Agarwal, R.; Khadgawat, R. Disrupted placental vitamin D metabolism and calcium signaling in gestational diabetes and pre-eclampsia patients. Endocrine 2023, 80, 191–200. [Google Scholar] [CrossRef] [PubMed]
  342. Wang, Y.; Wang, T.; Huo, Y.; Liu, L.; Liu, S.; Yin, X.; Wang, R.; Gao, X. Placenta expression of vitamin D and related genes in pregnant women with gestational diabetes mellitus. J. Steroid Biochem. Mol. Biol. 2020, 204, 105754. [Google Scholar] [CrossRef] [PubMed]
  343. Lee, J.H.; Davaatseren, M.; Lee, S. Rare PTH Gene Mutations Causing Parathyroid Disorders: A Review. Endocrinol. Metab. 2020, 35, 64–70. [Google Scholar] [CrossRef] [PubMed] [PubMed Central]
  344. Parkinson, D.B.; Thakker, R.V. A donor splice site mutation in the parathyroid hormone gene is associated with autosomal recessive hypoparathyroidism. Nat. Genet. 1992, 1, 149–152. [Google Scholar] [CrossRef] [PubMed]
  345. Bringhurst, F.R.; Demay, M.B.; Kronenberg, H.M. Hormones and Disorders of Mineral Metabolism. In Williams Textbook of Endocrinology, 14th ed.; Melmed, S., Koenig, R., Rosen, C.J., Auchus, R.J., Goldfine, A.B., Eds.; Chapter 29; Elsevier: Amsterdam, The Netherlands, 2019; pp. 1196–1255. ISBN 9780323555968. [Google Scholar]
  346. Bastepe, M.; Raas-Rothschild, A.; Silver, J.; Weissman, I.; Wientroub, S.; Jüppner, H.; Gillis, D. A form of Jansen’s metaphyseal chondrodysplasia with limited metabolic and skeletal abnormalities is caused by a novel activating parathyroid hormone (PTH)/PTH-related peptide receptor mutation. J. Clin. Endocrinol. Metab. 2004, 89, 3595–3600. [Google Scholar] [CrossRef] [PubMed]
  347. Schipani, E.; Kruse, K.; Jüppner, H. A constitutively active mutant PTH-PTHrP receptor in Jansen-type metaphyseal chondrodysplasia. Science 1995, 268, 98–100. [Google Scholar] [CrossRef] [PubMed]
  348. Savoldi, G.; Izzi, C.; Signorelli, M.; Bondioni, M.P.; Romani, C.; Lanzi, G.; Moratto, D.; Verdoni, L.; Pinotti, M.; Prefumo, F.; et al. Prenatal presentation and postnatal evolution of a patient with Jansen metaphyseal dysplasia with a novel missense mutation in PTH1R. Am. J. Med. Genet. A 2013, 161, 2614–2619. [Google Scholar] [CrossRef] [PubMed]
  349. Turan, S.; Bastepe, M. The GNAS complex locus and human diseases associated with loss-of-function mutations or epimutations within this imprinted gene. Horm. Res. Paediatr. 2013, 80, 229–241. [Google Scholar] [CrossRef] [PubMed] [PubMed Central]
  350. Lemos, M.C.; Thakker, R.V. GNAS mutations in Pseudohypoparathyroidism type 1a and related disorders. Hum. Mutat. 2015, 36, 11–19. [Google Scholar] [CrossRef] [PubMed] [PubMed Central]
  351. Chase, L.R.; Melson, G.L.; Aurbach, G.D. Pseudohypoparathyroidism: Defective excretion of 3’,5’-AMP in response to parathyroid hormone. J. Clin. Investig. 1969, 48, 1832–1844. [Google Scholar] [CrossRef] [PubMed] [PubMed Central]
  352. Saraff, V.; Nadar, R.; Shaw, N. Neonatal Bone Disorders. Front. Pediatr. 2021, 9, 602552. [Google Scholar] [CrossRef] [PubMed] [PubMed Central]
  353. Wakeling, E.L.; Brioude, F.; Lokulo-Sodipe, O.; O’Connell, S.M.; Salem, J.; Bliek, J.; Canton, A.P.; Chrzanowska, K.H.; Davies, J.H.; Dias, R.P.; et al. Diagnosis and management of Silver-Russell syndrome: First international consensus statement. Nat. Rev. Endocrinol. 2017, 13, 105–124. [Google Scholar] [CrossRef] [PubMed]
  354. Juriaans, A.F.; Kerkhof, G.F.; Mahabier, E.F.; Sas, T.C.J.; Zwaveling-Soonawala, N.; Touwslager, R.N.H.; Rotteveel, J.; Hokken-Koelega, A.C.S. Temple Syndrome: Clinical Findings, Body Composition and Cognition in 15 Patients. J. Clin. Med. 2022, 11, 6289. [Google Scholar] [CrossRef] [PubMed] [PubMed Central]
  355. Borjas Mendoza, P.A.; Daley, S.F.; Mendez, M.D. Beckwith-Wiedemann Syndrome. [Updated 2024 Jan 7]. In StatPearls [Internet]; StatPearls Publishing: Treasure Island, FL, USA, 2024. Available online: https://www.ncbi.nlm.nih.gov/books/NBK558993/ (accessed on 29 December 2024).
  356. Papadopoulou, A.; Bountouvi, E.; Karachaliou, F.E. The Molecular Basis of Calcium and Phosphorus Inherited Metabolic Disorders. Genes 2021, 12, 734. [Google Scholar] [CrossRef] [PubMed] [PubMed Central]
  357. Lacruz, R.S.; Feske, S. Diseases caused by mutations in ORAI1 and STIM1. Ann. N. Y. Acad. Sci. 2015, 1356, 45–79. [Google Scholar] [CrossRef] [PubMed] [PubMed Central]
  358. Justiz Vaillant, A.A.; Mohseni, M. Severe Combined Immunodeficiency. [Updated 2023 Aug 8]. In StatPearls [Internet]; StatPearls Publishing: Treasure Island, FL, USA, 2024. Available online: https://www.ncbi.nlm.nih.gov/books/NBK539762/ (accessed on 29 December 2024).
  359. Harris, E.; Burki, U.; Marini-Bettolo, C.; Neri, M.; Scotton, C.; Hudson, J.; Bertoli, M.; Evangelista, T.; Vroling, B.; Polvikoski, T.; et al. Complex phenotypes associated with STIM1 mutations in both coiled coil and EF-hand domains. Neuromuscul. Disord. 2017, 27, 861–872. [Google Scholar] [CrossRef] [PubMed]
  360. Borsani, O.; Piga, D.; Costa, S.; Govoni, A.; Magri, F.; Artoni, A.; Cinnante, C.M.; Fagiolari, G.; Ciscato, P.; Moggio, M.; et al. Stormorken Syndrome Caused by a p.R304W STIM1 Mutation: The First Italian Patient and a Review of the Literature. Front. Neurol. 2018, 9, 859. [Google Scholar] [CrossRef] [PubMed] [PubMed Central]
  361. Gamberucci, A.; Nanni, C.; Pierantozzi, E.; Serano, M.; Protasi, F.; Rossi, D.; Sorrentino, V. TAM-associated CASQ1 mutants diminish intracellular Ca2+ content and interfere with regulation of SOCE. J. Muscle Res. Cell Motil. 2024, 45, 275–284. [Google Scholar] [CrossRef] [PubMed]
  362. Riehle, M.; Büscher, A.K.; Gohlke, B.O.; Kaßmann, M.; Kolatsi-Joannou, M.; Bräsen, J.H.; Nagel, M.; Becker, J.U.; Winyard, P.; Hoyer, P.F.; et al. TRPC6 G757D Loss-of-Function Mutation Associates with FSGS. J. Am. Soc. Nephrol. 2016, 27, 2771–2783. [Google Scholar] [CrossRef] [PubMed] [PubMed Central]
  363. Barker, D.J. The fetal and infant origins of adult disease. BMJ 1990, 301, 1111. [Google Scholar] [CrossRef] [PubMed] [PubMed Central]
  364. Anas, M.; Diniz, W.J.S.; Menezes, A.C.B.; Reynolds, L.P.; Caton, J.S.; Dahlen, C.R.; Ward, A.K. Maternal Mineral Nutrition Regulates Fetal Genomic Programming in Cattle: A Review. Metabolites 2023, 13, 593. [Google Scholar] [CrossRef] [PubMed] [PubMed Central]
  365. Roseboom, T.J. Epidemiological evidence for the developmental origins of health and disease: Effects of prenatal undernutrition in humans. J. Endocrinol. 2019, 242, T135–T144. [Google Scholar] [CrossRef] [PubMed]
  366. Cooper, C.; Fall, C.; Egger, P.; Hobbs, R.; Eastell, R.; Barker, D. Growth in infancy and bone mass in later life. Ann. Rheum. Dis. 1997, 56, 17–21. [Google Scholar] [CrossRef]
  367. Bateson, P.; Barker, D.; Clutton-Brock, T.; Deb, D.; D’Udine, B.; Foley, R.A.; Gluckman, P.; Godfrey, K.; Kirkwood, T.; Lahr, M.M.; et al. Developmental plasticity and human health. Nature 2004, 430, 419–421. [Google Scholar] [CrossRef] [PubMed]
  368. Javaid, M.K.; Crozier, S.R.; Harvey, N.C.; Gale, C.R.; Dennison, E.M.; Boucher, B.J.; Arden, N.K.; Godfrey, K.M.; Cooper, C.; Princess Anne Hospital Study Group. Maternal vitamin D status during pregnancy and childhood bone mass at age 9 years: A longitudinal study. Lancet 2006, 367, 36–43. [Google Scholar] [CrossRef] [PubMed]
  369. Lawlor, D.A.; Wills, A.K.; Fraser, A.; Sayers, A.; Fraser, W.D.; Tobias, J.H. Association of maternal vitamin D status during pregnancy with bone-mineral content in offspring: A prospective cohort study. Lancet 2013, 381, 2176–2183. [Google Scholar] [CrossRef] [PubMed] [PubMed Central]
  370. Ganpule, A.; Yajnik, C.S.; Fall, C.H.; Rao, S.; Fisher, D.J.; Kanade, A.; Cooper, C.; Naik, S.; Joshi, N.; Lubree, H.; et al. Bone mass in Indian children—relationships to maternal nutritional status and diet during pregnancy: The Pune Maternal Nutrition Study. J. Clin. Endocrinol. Metab. 2006, 91, 2994–3001. [Google Scholar] [CrossRef] [PubMed]
  371. Heppe, D.H.; Medina-Gomez, C.; Hofman, A.; Franco, O.H.; Rivadeneira, F.; Jaddoe, V.W. Maternal first-trimester diet and childhood bone mass: The Generation R Study. Am. J. Clin. Nutr. 2013, 98, 224–232. [Google Scholar] [CrossRef] [PubMed]
  372. Vaiserman, A.; Lushchak, O. Prenatal famine exposure and adult health outcomes: An epigenetic link. Environ. Epigenet. 2021, 7, dvab013. [Google Scholar] [CrossRef] [PubMed] [PubMed Central]
  373. Li, C.; Lumey, L.H. Exposure to the Chinese famine of 1959-61 in early life and long-term health conditions: A systematic review and meta-analysis. Int. J. Epidemiol. 2017, 46, 1157–1170. [Google Scholar] [CrossRef] [PubMed]
  374. Ravelli, A.C.; van der Meulen, J.H.; Michels, R.P.; Osmond, C.; Barker, D.J.; Hales, C.N.; Bleker, O.P. Glucose tolerance in adults after prenatal exposure to famine. Lancet 1998, 351, 173–177. [Google Scholar] [CrossRef] [PubMed]
  375. Waterland, R.A.; Kellermayer, R.; Laritsky, E.; Rayco-Solon, P.; Harris, R.A.; Travisano, M.; Zhang, W.; Torskaya, M.S.; Zhang, J.; Shen, L.; et al. Season of conception in rural Gambia affects DNA methylation at putative human metastable epialleles. PLoS Genet. 2010, 6, e1001252. [Google Scholar] [CrossRef] [PubMed] [PubMed Central]
  376. Heijmans, B.T.; Tobi, E.W.; Stein, A.D.; Putter, H.; Blauw, G.J.; Susser, E.S.; Slagboom, P.E.; Lumey, L.H. Persistent epigenetic differences associated with prenatal exposure to famine in humans. Proc. Natl. Acad. Sci. USA 2008, 105, 17046–17049. [Google Scholar] [CrossRef] [PubMed] [PubMed Central]
  377. Tobi, E.W.; Goeman, J.J.; Monajemi, R.; Gu, H.; Putter, H.; Zhang, Y.; Slieker, R.C.; Stok, A.P.; Thijssen, P.E.; Müller, F.; et al. DNA methylation signatures link prenatal famine exposure to growth and metabolism. Nat. Commun. 2014, 5, 5592. [Google Scholar] [CrossRef] [PubMed] [PubMed Central]
  378. Dominguez-Salas, P.; Moore, S.E.; Baker, M.S.; Bergen, A.W.; Cox, S.E.; Dyer, R.A.; Fulford, A.J.; Guan, Y.; Laritsky, E.; Silver, M.J.; et al. Maternal nutrition at conception modulates DNA methylation of human metastable epialleles. Nat. Commun. 2014, 5, 3746. [Google Scholar] [CrossRef] [PubMed] [PubMed Central]
  379. Harvey, N.C.; Sheppard, A.; Godfrey, K.M.; McLean, C.; Garratt, E.; Ntani, G.; Davies, L.; Murray, R.; Inskip, H.M.; Gluckman, P.D.; et al. Childhood bone mineral content is associated with methylation status of the RXRA promoter at birth. J. Bone Miner. Res. 2014, 29, 600–607. [Google Scholar] [CrossRef] [PubMed] [PubMed Central]
  380. Curtis, E.M.; Murray, R.; Titcombe, P.; Cook, E.; Clarke-Harris, R.; Costello, P.; Garratt, E.; Holbrook, J.D.; Barton, S.; Inskip, H.; et al. Perinatal DNA Methylation at CDKN2A Is Associated with Offspring Bone Mass: Findings from the Southampton Women’s Survey. J. Bone Miner. Res. 2017, 32, 2030–2040. [Google Scholar] [CrossRef] [PubMed] [PubMed Central]
  381. Godfrey, K.M.; Sheppard, A.; Gluckman, P.D.; Lillycrop, K.A.; Burdge, G.C.; McLean, C.; Rodford, J.; Slater-Jefferies, J.L.; Garratt, E.; Crozier, S.R.; et al. Epigenetic gene promoter methylation at birth is associated with child’s later adiposity. Diabetes 2011, 60, 1528–1534. [Google Scholar] [CrossRef] [PubMed] [PubMed Central]
  382. Moon, R.J.; Curtis, E.M.; Woolford, S.J.; Ashai, S.; Cooper, C.; Harvey, N.C. The importance of maternal pregnancy vitamin D for offspring bone health: Learnings from the MAVIDOS trial. Ther. Adv. Musculoskelet. Dis. 2021, 13, 1759720X211006979. [Google Scholar] [CrossRef] [PubMed] [PubMed Central]
  383. Moon, R.J.; D’Angelo, S.; Curtis, E.M.; Ward, K.A.; Crozier, S.R.; Schoenmakers, I.; Javaid, M.K.; Bishop, N.J.; Godfrey, K.M.; Cooper, C.; et al. Pregnancy vitamin D supplementation and offspring bone mineral density in childhood follow-up of a randomized controlled trial. Am. J. Clin. Nutr. 2024, 120, 1134–1142. [Google Scholar] [CrossRef] [PubMed]
  384. Brustad, N.; Garland, J.; Thorsen, J.; Sevelsted, A.; Krakauer, M.; Vinding, R.K.; Stokholm, J.; Bønnelykke, K.; Bisgaard, H.; Chawes, B.L. Effect of High-Dose vs Standard-Dose Vitamin D Supplementation in Pregnancy on Bone Mineralization in Offspring Until Age 6 Years: A Prespecified Secondary Analysis of a Double-Blinded, Randomized Clinical Trial. JAMA Pediatr. 2020, 174, 419–427. [Google Scholar] [CrossRef] [PubMed] [PubMed Central]
  385. Palacios, C.; Kostiuk, L.L.; Cuthbert, A.; Weeks, J. Vitamin D supplementation for women during pregnancy. Cochrane Database Syst. Rev. 2024, 7, CD008873. [Google Scholar] [CrossRef] [PubMed] [PubMed Central]
  386. Kerr, B.; Leiva, A.; Farías, M.; Contreras-Duarte, S.; Toledo, F.; Stolzenbach, F.; Silva, L.; Sobrevia, L. Foetoplacental epigenetic changes associated with maternal metabolic dysfunction. Placenta 2018, 69, 146–152. [Google Scholar] [CrossRef] [PubMed]
  387. Dahlen, C.R.; Borowicz, P.P.; Ward, A.K.; Caton, J.S.; Czernik, M.; Palazzese, L.; Loi, P.; Reynolds, L.P. Programming of Embryonic Development. Int. J. Mol. Sci. 2021, 22, 11668. [Google Scholar] [CrossRef] [PubMed] [PubMed Central]
  388. Fitz-James, M.H.; Cavalli, G. Molecular mechanisms of transgenerational epigenetic inheritance. Nat. Rev. Genet. 2022, 23, 325–341. [Google Scholar] [CrossRef] [PubMed]
  389. Le Blévec, E.; Muroňová, J.; Ray, P.F.; Arnoult, C. Paternal epigenetics: Mammalian sperm provide much more than DNA at fertilization. Mol. Cell Endocrinol. 2020, 518, 110964. [Google Scholar] [CrossRef] [PubMed]
  390. Zafar, M.I.; Lu, S.; Li, H. Sperm-oocyte interplay: An overview of spermatozoon’s role in oocyte activation and current perspectives in diagnosis and fertility treatment. Cell Biosci. 2021, 11, 4. [Google Scholar] [CrossRef] [PubMed] [PubMed Central]
  391. Ito, S.; Shen, L.; Dai, Q.; Wu, S.C.; Collins, L.B.; Swenberg, J.A.; He, C.; Zhang, Y. Tet proteins can convert 5-methylcytosine to 5-formylcytosine and 5-carboxylcytosine. Science 2011, 333, 1300–1303. [Google Scholar] [CrossRef] [PubMed] [PubMed Central]
  392. Liu, W.; Wu, G.; Xiong, F.; Chen, Y. Advances in the DNA methylation hydroxylase TET1. Biomark. Res. 2021, 9, 76. [Google Scholar] [CrossRef] [PubMed] [PubMed Central]
  393. Ross, S.E.; Bogdanovic, O. TET enzymes, DNA demethylation and pluripotency. Biochem. Soc. Trans. 2019, 47, 875–885. [Google Scholar] [CrossRef] [PubMed]
  394. Yang, J.; Bashkenova, N.; Zang, R.; Huang, X.; Wang, J. The roles of TET family proteins in development and stem cells. Development 2020, 147, dev183129. [Google Scholar] [CrossRef] [PubMed] [PubMed Central]
  395. Dawlaty, M.M.; Breiling, A.; Le, T.; Barrasa, M.I.; Raddatz, G.; Gao, Q.; Powell, B.E.; Cheng, A.W.; Faull, K.F.; Lyko, F.; et al. Loss of Tet enzymes compromises proper differentiation of embryonic stem cells. Dev. Cell. 2014, 29, 102–111. [Google Scholar] [CrossRef] [PubMed] [PubMed Central]
  396. Zhang, X.; Zhang, Y.; Wang, C.; Wang, X. TET (Ten-eleven translocation) family proteins: Structure, biological functions and applications. Signal Transduct. Target. Ther. 2023, 8, 297. [Google Scholar] [CrossRef] [PubMed] [PubMed Central]
  397. Wang, X.X.; Xiao, F.H.; Li, Q.G.; Liu, J.; He, Y.H.; Kong, Q.P. Large-scale DNA methylation expression analysis across 12 solid cancers reveals hypermethylation in the calcium-signaling pathway. Oncotarget 2017, 8, 11868–11876. [Google Scholar] [CrossRef] [PubMed] [PubMed Central]
  398. Cantone, I.; Fisher, A.G. Epigenetic programming and reprogramming during development. Nat. Struct. Mol. Biol. 2013, 20, 282–289. [Google Scholar] [CrossRef] [PubMed]
  399. Sendžikaitė, G.; Kelsey, G. The role and mechanisms of DNA methylation in the oocyte. Essays Biochem. 2019, 63, 691–705. [Google Scholar] [CrossRef] [PubMed] [PubMed Central]
  400. Burton, A.; Brochard, V.; Galan, C.; Ruiz-Morales, E.R.; Rovira, Q.; Rodriguez-Terrones, D.; Kruse, K.; Le Gras, S.; Udayakumar, V.S.; Chin, H.G.; et al. Heterochromatin establishment during early mammalian development is regulated by pericentromeric RNA and characterized by non-repressive H3K9me3. Nat. Cell Biol. 2020, 22, 767–778. [Google Scholar] [CrossRef] [PubMed] [PubMed Central]
  401. Stein, P.; Savy, V.; Williams, A.M.; Williams, C.J. Modulators of calcium signalling at fertilization. Open Biol. 2020, 10, 200118. [Google Scholar] [CrossRef] [PubMed] [PubMed Central]
  402. Ozil, J.P.; Banrezes, B.; Tóth, S.; Pan, H.; Schultz, R.M. Ca2+ oscillatory pattern in fertilized mouse eggs affects gene expression and development to term. Dev. Biol. 2006, 300, 534–544. [Google Scholar] [CrossRef] [PubMed]
  403. Lu, T.; Zhang, Y.; Su, Y.; Zhou, D.; Xu, Q. Role of store-operated Ca2+ entry in cardiovascular disease. Cell Commun. Signal. 2022, 20, 33. [Google Scholar] [CrossRef] [PubMed] [PubMed Central]
  404. Masson, B.; Montani, D.; Humbert, M.; Capuano, V.; Antigny, F. Role of Store-Operated Ca2+ Entry in the Pulmonary Vascular Remodeling Occurring in Pulmonary Arterial Hypertension. Biomolecules 2021, 11, 1781. [Google Scholar] [CrossRef] [PubMed] [PubMed Central]
  405. Kono, T.; Tong, X.; Taleb, S.; Bone, R.N.; Iida, H.; Lee, C.C.; Sohn, P.; Gilon, P.; Roe, M.W.; Evans-Molina, C. Impaired Store-Operated Calcium Entry and STIM1 Loss Lead to Reduced Insulin Secretion and Increased Endoplasmic Reticulum Stress in the Diabetic β-Cell. Diabetes 2018, 67, 2293–2304. [Google Scholar] [CrossRef] [PubMed] [PubMed Central]
  406. Tao, Y.; Yazdizadeh Shotorbani, P.; Inman, D.; Das-Earl, P.; Ma, R. Store-operated Ca2+ entry inhibition ameliorates high glucose and ANG II-induced podocyte apoptosis and mitochondrial damage. Am. J. Physiol. Renal Physiol. 2023, 324, F494–F504. [Google Scholar] [CrossRef] [PubMed] [PubMed Central]
  407. Frost, J.M.; Amante, S.M.; Okae, H.; Jones, E.M.; Ashley, B.; Lewis, R.M.; Cleal, J.K.; Caley, M.P.; Arima, T.; Maffucci, T.; et al. Regulation of human trophoblast gene expression by endogenous retroviruses. Nat. Struct. Mol. Biol. 2023, 30, 527–538. [Google Scholar] [CrossRef] [PubMed] [PubMed Central]
  408. Oomen, M.E.; Torres-Padilla, M.E. Jump-starting life: Balancing transposable element co-option and genome integrity in the developing mammalian embryo. EMBO Rep. 2024, 25, 1721–1733. [Google Scholar] [CrossRef] [PubMed] [PubMed Central]
  409. Yu, M.; Hu, X.; Pan, Z.; Du, C.; Jiang, J.; Zheng, W.; Cai, H.; Wang, Y.; Deng, W.; Wang, H.; et al. Endogenous retrovirus-derived enhancers confer the transcriptional regulation of human trophoblast syncytialization. Nucleic Acids Res. 2023, 51, 4745–4759. [Google Scholar] [CrossRef] [PubMed] [PubMed Central]
  410. Huang, L.; Tu, Z.; Wei, L.; Sun, W.; Wang, Y.; Bi, S.; He, F.; Du, L.; Chen, J.; Kzhyshkowska, J.; et al. Generating Functional Multicellular Organoids from Human Placenta Villi. Adv. Sci. 2023, 10, e2301565. [Google Scholar] [CrossRef] [PubMed] [PubMed Central]
  411. Okae, H.; Toh, H.; Sato, T.; Hiura, H.; Takahashi, S.; Shirane, K.; Kabayama, Y.; Suyama, M.; Sasaki, H.; Arima, T. Derivation of Human Trophoblast Stem Cells. Cell Stem Cell 2018, 22, 50–63.e6. [Google Scholar] [CrossRef] [PubMed]
  412. Barnes, M.V.C.; Pantazi, P.; Holder, B. Circulating extracellular vesicles in healthy and pathological pregnancies: A scoping review of methodology, rigour and results. J. Extracell. Vesicles. 2023, 12, e12377. [Google Scholar] [CrossRef] [PubMed] [PubMed Central]
  413. Aplin, J.D.; Jones, C.J.P. Cell dynamics in human villous trophoblast. Hum. Reprod. Update 2021, 27, 904–922. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Development and implantation of the blastocyst. Compiled from Burton GJ and Jauniaux E 2021 [6], Mitchell B and Sharma R 2005 [39], Harrison RG 1963 [40], and Sadler TW 2023 [41]; ~, approximately.
Figure 1. Development and implantation of the blastocyst. Compiled from Burton GJ and Jauniaux E 2021 [6], Mitchell B and Sharma R 2005 [39], Harrison RG 1963 [40], and Sadler TW 2023 [41]; ~, approximately.
Ijms 26 00383 g001
Figure 2. Section of a human placenta showing three cotyledons separated by septa (S). Within each of these is a mass of foetal villi branching from a primary villous stem anchored to the decidua basalis, which forms the maternal placenta. The villi are covered by trophoblasts: an inner layer of cytotrophoblasts and an outer layer of syncytiotrophoblasts. These erode the walls of small uterine spiral arteries, and the blood empties into the intervillous spaces. The highly branched terminal villi (red arrows) float freely in a lake of maternal blood. Nutrients and minerals pass from mother to foetus, but there is no continuity between foetal and maternal circulations. AM, amnion; U.A., umbilical arteries; U.V., umbilical vein; UT.A., uterine artery; V., uterine vein. Source: Harrison RG: A Textbook of Human Embryology 2nd Ed. Blackwell Scientific Publications Ltd. Oxford 1963 [40]. Figure 35 reproduced with permission from Wiley-Blackwell, John Wiley and Sons.
Figure 2. Section of a human placenta showing three cotyledons separated by septa (S). Within each of these is a mass of foetal villi branching from a primary villous stem anchored to the decidua basalis, which forms the maternal placenta. The villi are covered by trophoblasts: an inner layer of cytotrophoblasts and an outer layer of syncytiotrophoblasts. These erode the walls of small uterine spiral arteries, and the blood empties into the intervillous spaces. The highly branched terminal villi (red arrows) float freely in a lake of maternal blood. Nutrients and minerals pass from mother to foetus, but there is no continuity between foetal and maternal circulations. AM, amnion; U.A., umbilical arteries; U.V., umbilical vein; UT.A., uterine artery; V., uterine vein. Source: Harrison RG: A Textbook of Human Embryology 2nd Ed. Blackwell Scientific Publications Ltd. Oxford 1963 [40]. Figure 35 reproduced with permission from Wiley-Blackwell, John Wiley and Sons.
Ijms 26 00383 g002
Figure 3. Proposed scheme in which ezrin bridges the scaffolding protein NHERF1 and the actin skeleton, so stabilising proteins in the microvillar membrane and connecting them with cytosolic signalling complexes. (i) In a dormant form, the C-terminal tail of ezrin binds to the N-terminal and closes the molecule. (ii) Phosphorylation of residues located between the N- and C-terminals blocks the association and opens the ezrin molecule. (iii) The freed C-terminal binds to actin; the N-terminal binds to the scaffolding protein NHERF1 associated with the plasma membrane. (iv) NHERF1 connects membrane-associated proteins with transiently assembled cytosolic signalling complexes. NaPi-2a, PTH1R, and NHE3 are shown as examples. NaPi-2a, sodium–phosphate cotransporter 2a; NHE3, sodium–hydrogen exchanger 3; NHERF1, Na+/H+ exchanger regulatory factor 1; PTH1R, parathyroid hormone 1 receptor.
Figure 3. Proposed scheme in which ezrin bridges the scaffolding protein NHERF1 and the actin skeleton, so stabilising proteins in the microvillar membrane and connecting them with cytosolic signalling complexes. (i) In a dormant form, the C-terminal tail of ezrin binds to the N-terminal and closes the molecule. (ii) Phosphorylation of residues located between the N- and C-terminals blocks the association and opens the ezrin molecule. (iii) The freed C-terminal binds to actin; the N-terminal binds to the scaffolding protein NHERF1 associated with the plasma membrane. (iv) NHERF1 connects membrane-associated proteins with transiently assembled cytosolic signalling complexes. NaPi-2a, PTH1R, and NHE3 are shown as examples. NaPi-2a, sodium–phosphate cotransporter 2a; NHE3, sodium–hydrogen exchanger 3; NHERF1, Na+/H+ exchanger regulatory factor 1; PTH1R, parathyroid hormone 1 receptor.
Ijms 26 00383 g003
Figure 4. Schematic of Na+/H+ exchanger regulatory factor 1 (NHERF1) based on the models of Zhang et al. (2019) [62] and Bhattycharia et al. (2019) [63]. GYGF, core PDZ-binding motif; PPI, protein-phosphatase-1-binding site; S290, key phosphorylation site.
Figure 4. Schematic of Na+/H+ exchanger regulatory factor 1 (NHERF1) based on the models of Zhang et al. (2019) [62] and Bhattycharia et al. (2019) [63]. GYGF, core PDZ-binding motif; PPI, protein-phosphatase-1-binding site; S290, key phosphorylation site.
Ijms 26 00383 g004
Figure 5. Phospholipase C releases inositol trisphosphate from phosphatidylinositol 4,5 bisphosphate (PIP2). Arachidonic acid is released from PIP2 by phospholipase A2.
Figure 5. Phospholipase C releases inositol trisphosphate from phosphatidylinositol 4,5 bisphosphate (PIP2). Arachidonic acid is released from PIP2 by phospholipase A2.
Ijms 26 00383 g005
Figure 6. Structure of TRPV6, as deduced from structural studies (PDB codes 5IWK, 5IWP, and 62EF) [132,133,134]. (A) Side and (B) top views of the four TRPV6 monomers, displayed as ribbons with α-helices and β-sheets as cylinders and arrows, respectively. Three monomers are coloured uniformly (purple, blue, and dark pink); one monomer is differentially coloured to highlight crucial regions: N-terminal helix (yellow), ankyrin repeats (cyan), intracellular N-terminal region (dark green), membrane region (green), and intra-cellular C-terminal region (orange). The Ca2+ ion in the pore is displayed as a yellow ball. (C) Schematic overview of secondary structural elements and other structural features. Colour coding as in panels (A,B). Note that the C-terminal helices 1–3 (CtH1–3; light orange) are not observed in the 5IWK structure displayed in panels (A,B,D). (D) Close-up of the TRPV6 pore-forming region. The ion selectivity filter forming the side chains of D542 and surrounding residues are shown in ball-and-stick representation, colour-coded on a by-atom-type basis. The carboxylic groups lining the pore are clearly visible. The ion selectivity filter is the narrowest part of the channel. This figure was published in Advances in Clinical Chemistry, vol. 113, Walker V, Vuister GW, Biochemistry and pathophysiology of the transient potential receptor vanilloid 6 (TRPV6) calcium channel, pp 43–100, 2023 [115]. Copyright Elsevier. Reproduced with permission.
Figure 6. Structure of TRPV6, as deduced from structural studies (PDB codes 5IWK, 5IWP, and 62EF) [132,133,134]. (A) Side and (B) top views of the four TRPV6 monomers, displayed as ribbons with α-helices and β-sheets as cylinders and arrows, respectively. Three monomers are coloured uniformly (purple, blue, and dark pink); one monomer is differentially coloured to highlight crucial regions: N-terminal helix (yellow), ankyrin repeats (cyan), intracellular N-terminal region (dark green), membrane region (green), and intra-cellular C-terminal region (orange). The Ca2+ ion in the pore is displayed as a yellow ball. (C) Schematic overview of secondary structural elements and other structural features. Colour coding as in panels (A,B). Note that the C-terminal helices 1–3 (CtH1–3; light orange) are not observed in the 5IWK structure displayed in panels (A,B,D). (D) Close-up of the TRPV6 pore-forming region. The ion selectivity filter forming the side chains of D542 and surrounding residues are shown in ball-and-stick representation, colour-coded on a by-atom-type basis. The carboxylic groups lining the pore are clearly visible. The ion selectivity filter is the narrowest part of the channel. This figure was published in Advances in Clinical Chemistry, vol. 113, Walker V, Vuister GW, Biochemistry and pathophysiology of the transient potential receptor vanilloid 6 (TRPV6) calcium channel, pp 43–100, 2023 [115]. Copyright Elsevier. Reproduced with permission.
Ijms 26 00383 g006
Figure 7. The domains of STIM1 and their functions [61,185,187]: SP, signal peptide; EF1 and EF2, EF-hand domains; SAM, sterile alpha motif; TM, transmembrane domain; CC1, CC2, and CC3, coiled-coil domains; CAD, CRAC activation domain; SOAR, STIM-Orai activating domain; ID, inactivation domain; STIM1L, a peptide insert; P/S, proline/serine-rich peptide; EB, EB1-binding domain; PBD, polybasic domain; U, peptide sequences.
Figure 7. The domains of STIM1 and their functions [61,185,187]: SP, signal peptide; EF1 and EF2, EF-hand domains; SAM, sterile alpha motif; TM, transmembrane domain; CC1, CC2, and CC3, coiled-coil domains; CAD, CRAC activation domain; SOAR, STIM-Orai activating domain; ID, inactivation domain; STIM1L, a peptide insert; P/S, proline/serine-rich peptide; EB, EB1-binding domain; PBD, polybasic domain; U, peptide sequences.
Ijms 26 00383 g007
Figure 8. Binding of activated STIM1 to Orai1 to open the Orai1 channel for Ca2+ influx. SAM, sterile alpha motif; TM, transmembrane domain; CC1, CC2, and CC3, coiled-coil domains; PBD, polybasic domain; PIP2, phosphatidylinositol 4,5 bisphosphate.
Figure 8. Binding of activated STIM1 to Orai1 to open the Orai1 channel for Ca2+ influx. SAM, sterile alpha motif; TM, transmembrane domain; CC1, CC2, and CC3, coiled-coil domains; PBD, polybasic domain; PIP2, phosphatidylinositol 4,5 bisphosphate.
Ijms 26 00383 g008
Figure 9. Store-operated calcium entry (SOCE). (A) (1) Agonist stimulation of a membrane receptor (2) activates phospholipase C (PLC), releasing inositol trisphosphate (IP3), and (3) IP3 activates the IP3 receptor in the ER membrane, leading to release of Ca2+ into the cytosol to activate signalling cascades. Some Ca2+ is taken into mitochondria, and (4) surplus Ca2+ is extruded from the cytosol by plasma membrane Ca2+-ATPase 1 (PMCA1) and/or the NCX Na+/Ca2+ exchanger. (B) (5) Ca2+ depletion in the endoplasmic reticulum (ER) activates the Ca2+ sensor STIM1, and (6) activated STIM1 binds to Orai1 in the plasma membrane, opening the channel for Ca2+ entry. (7) Activated STIM1 can also associate with transient receptor C (TRPC) proteins, opening their channels for Ca2+ entry. (8) Imported Ca2+ is transported into the ER by sarcoendoplasmic reticulum Ca2+ ATPase (SERCA) to replenish the ER stores.
Figure 9. Store-operated calcium entry (SOCE). (A) (1) Agonist stimulation of a membrane receptor (2) activates phospholipase C (PLC), releasing inositol trisphosphate (IP3), and (3) IP3 activates the IP3 receptor in the ER membrane, leading to release of Ca2+ into the cytosol to activate signalling cascades. Some Ca2+ is taken into mitochondria, and (4) surplus Ca2+ is extruded from the cytosol by plasma membrane Ca2+-ATPase 1 (PMCA1) and/or the NCX Na+/Ca2+ exchanger. (B) (5) Ca2+ depletion in the endoplasmic reticulum (ER) activates the Ca2+ sensor STIM1, and (6) activated STIM1 binds to Orai1 in the plasma membrane, opening the channel for Ca2+ entry. (7) Activated STIM1 can also associate with transient receptor C (TRPC) proteins, opening their channels for Ca2+ entry. (8) Imported Ca2+ is transported into the ER by sarcoendoplasmic reticulum Ca2+ ATPase (SERCA) to replenish the ER stores.
Ijms 26 00383 g009
Table 1. Events in the formation of the mature human placenta [6,39,40,41].
Table 1. Events in the formation of the mature human placenta [6,39,40,41].
Time pf. ꝉꝉ Figure NumberEventsSource of Ca2+
01AFertilisation. Sperm releases phospholipase zeta (PLCζ), leading to Ca2+ release and cell division.Endoplasmic reticulum (ER) stores in oocyte
Approx. 36 h Zygote cleaved to two cells.ꝉꝉꝉ? ER stores
Approx. 40 h1BFour-cell stage.ꝉꝉꝉ? ER stores
3–4 d1CMorula, a sphere of 12–16 cells enclosed within the ovarian zona pellucida, reaches junction of fallopian tube with uterus.ꝉꝉꝉ? ER stores
Approx.
4.5 d
1DCavity in morula (blastocoel) fills with fluid, forming a blastocyst. The outer cells form a spherical wall of trophoblasts and inner cells accumulate at one pole as an inner cell mass, which forms the embryo.Uterine secretions diffusing into blastocoel
Approx. 5 d Start of blastocyst implantation in uterine endometrium induces decidualisation of endometrial glands.Uterine secretions via blastocoel
Approx.
7–8 d
1EBlastocyst partially implanted. Inner cell mass now has two layers: epiblast and hypoblast. The amniotic cavity develops in the epiblast, and hypoblast cells extend around the blastocoel, forming the primary yolk sac. The trophoblast wall produces the following: i) a cellular syncytium, which surrounds the blastocyst, invades the endometrium, and contains lacunae and ii) a lining of extra-embryonic mesoderm, which later fuses with the syncytium to form the foetal chorion.Decidual fluid-? via yolk sac
Approx. 9 d1FBlastocyst is fully implanted within the endometrium, syncytial lacunae extend into the endometrial stroma, and columns of syncytial villi start to grow towards the basal layer of uterine epithelium (anchoring/stem villi).Decidual fluid-? via yolk sac
Approx.
12 d
Syncytial clefts fuse with endometrial glands containing decidual secretion. Decidual fluid (uterine secretion) histotrophic nutrition
By 14 d Syncytial trophoblasts breach uterine vessels, and blood escapes into the syncytial lacunae. Syncytial columns now anchor the foetal chorion to the basal layer of the uterine epithelium (anchoring/stem villi). Finger-like processes of cytotrophoblasts grow into the syncytium, forming primary villi that surround the blastocyst. Villi then degenerate, except those adjacent to the uterus, which form the chorion frondosum and ultimately the discoid placenta. Decidual fluid, histotrophic nutrition, and some maternal blood
15 d1GExtra-embryonic mesoderm invades the primary villi, converting them to secondary villi, now classed as chorionic villi, and forms a connecting stalk to the embryo, which will become the umbilical cord. The chorionic villi are suspended in spaces of maternal blood.Decidual fluid, histotrophic nutrition, and some maternal blood
15–20 d Foetal capillaries develop in the secondary villi, which then become tertiary villi.Histotrophic nutrition and some maternal blood
10–12 weeks Maternal circulation to placenta established.Maternal blood and some histotrophic nutrition
By 20 weeksFigure 2Mature placenta formed with free-floating terminal villi—the dominant absorption structures.Maternal blood
From 20 weeks Exponential growth of placenta to term.Maternal blood
pf., postfertilisation. ꝉꝉ Figure numbers 1A to IG are the numbers for these stages as depicted in Figure 1. ꝉꝉꝉ? Indicates that this is the likely source but conjectural. Approx., approximately.
Table 2. Human mutations relevant to placental calcium transport.
Table 2. Human mutations relevant to placental calcium transport.
Disorder
[References]
GeneProteinPlasma Ca2+PTH Bone DeformityBirth Weight
Isolated PTH deficiency
[343,344]
PTRHPTHNoNormal?
Jansen’s metaphyseal chondrodysplasia
[345,346,347,348]
PTH1RPTH/PTHrP receptorYes
Pseudohypoparathyroidism 1A
[345,349,350,351]
GNAS from motherGsα YesNormal?
Pseudo-pseudohypoparathyroidism
[345,349,350,351]
GNAS from fatherGsαNormalNormalYesNormal?
I-cell disease
[352]
GNPTABN-acetylglucosamine phosphotransferase ↓ or normalYesLow
Severe neonatal hyperparathyroidism
[294,301,352]
CaSR
inactivating
Calcium sensor receptor↑↑Not suppressedYesUsually diagnosed postnatally
Silver–Russell syndrome
[267,353]
IGF2IGF2 deficiencyNormalNormalYesLow
Temple’s syndrome
[250,354]
IGF2IGF2 deficiency Normal?Normal?NoLow
Beckwith–Wiedemann syndrome
[267,355]
IGF2 loss of imprinting
IGF2 excess Normal?Normal?HemihypertrophyIncreased
Transient neonatal hyperparathyroidism
[155,156,157,158,159]
TRPV6Transient potential receptor vanilloid 6
(TRPV6)
↓ or normalYesNormal or low
Normal? assumed normal because there is no reported information. Add to legend: ↑↑ large increase, ↑ increase, ↓ decrease.
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Walker, V. The Molecular Biology of Placental Transport of Calcium to the Human Foetus. Int. J. Mol. Sci. 2025, 26, 383. https://doi.org/10.3390/ijms26010383

AMA Style

Walker V. The Molecular Biology of Placental Transport of Calcium to the Human Foetus. International Journal of Molecular Sciences. 2025; 26(1):383. https://doi.org/10.3390/ijms26010383

Chicago/Turabian Style

Walker, Valerie. 2025. "The Molecular Biology of Placental Transport of Calcium to the Human Foetus" International Journal of Molecular Sciences 26, no. 1: 383. https://doi.org/10.3390/ijms26010383

APA Style

Walker, V. (2025). The Molecular Biology of Placental Transport of Calcium to the Human Foetus. International Journal of Molecular Sciences, 26(1), 383. https://doi.org/10.3390/ijms26010383

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop