Next Article in Journal
Molecular Targets of Pinocembrin Underlying Its Regenerative Activities in Human Keratinocytes
Next Article in Special Issue
The Potential Therapeutic Role of Metformin in Diabetic and Non-Diabetic Bone Impairment
Previous Article in Journal
Continuous Subcutaneous Insulin Infusion (CSII) Combined with Oral Glucose-Lowering Drugs in Type 2 Diabetes: A Systematic Review and Network Meta-Analysis of Randomized, Controlled Trials
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Role of Calcimimetics in Treating Bone and Mineral Disorders Related to Chronic Kidney Disease

1
Division of Nephrology, Department of Medicine, Cardinal-Tien Hospital, School of Medicine, Fu-Jen Catholic University, New Taipei City 24205, Taiwan
2
Division of Nephrology, Department of Internal Medicine, Shuang Ho Hospital, School of Medicine, College of Medicine, Taipei Medical University, New Taipei City 11031, Taiwan
3
TMU Research Centre of Urology and Kidney, Taipei Medical University, New Taipei City 11031, Taiwan
4
Graduate Institute of Clinical Medicine, College of Medicine, Taipei Medical University, New Taipei City 11031, Taiwan
5
Department of Medical Research, Shuang Ho Hospital, Taipei Medical University, New Taipei City 11031, Taiwan
6
Department of Biology and Anatomy, National Defense Medical Center, Taipei 11490, Taiwan
7
Section of Nephrology, Department of Medicine, Antai Medical Care Corporation, Anti Tian-Sheng Memorial Hospital, Pingtung 92842, Taiwan
8
Division of Nephrology, Department of Medicine, Taipei Tzu Chi Hospital, Buddhist Tzu Chi Medical Foundation, School of Medicine, Buddhist Tzu Chi University, Hualien 97004, Taiwan
9
Division of Nephrology, Department of Medicine, Fu-Jen Catholic University Hospital, School of Medicine, Fu-Jen Catholic University, New Taipei City 24205, Taiwan
*
Author to whom correspondence should be addressed.
Pharmaceuticals 2022, 15(8), 952; https://doi.org/10.3390/ph15080952
Submission received: 4 July 2022 / Revised: 25 July 2022 / Accepted: 27 July 2022 / Published: 31 July 2022

Abstract

:
Renal osteodystrophy is common in patients with chronic kidney disease and end-stage renal disease and leads to the risks of fracture and extraosseous vascular calcification. Secondary hyperparathyroidism (SHPT) is characterized by a compensatory increase in parathyroid hormone (PTH) secretion in response to decreased renal phosphate excretion, resulting in potentiating bone resorption and decreased bone quantity and quality. Calcium-sensing receptors (CaSRs) are group C G-proteins and negatively regulate the parathyroid glands through (1) increasing CaSR insertion within the plasma membrane, (2) increasing 1,25-dihydroxy vitamin D3 within the kidney and parathyroid glands, (3) inhibiting fibroblast growth factor 23 (FGF23) in osteocytes, and (4) attenuating intestinal calcium absorption through Transient Receptor Potential Vanilloid subfamily member 6 (TRPV6). Calcimimetics (CaMs) decrease PTH concentrations without elevating the serum calcium levels or extraosseous calcification through direct interaction with cell membrane CaSRs. CaMs reduce osteoclast activity by reducing stress-induced oxidative autophagy and improving Wnt-10b release, which promotes the growth of osteoblasts and subsequent mineralization. CaMs also directly promote osteoblast proliferation and survival. Consequently, bone quality may improve due to decreased bone resorption and improved bone formation. CaMs modulate cardiovascular fibrosis, calcification, and renal fibrosis through different mechanisms. Therefore, CaMs assist in treating SHPT. This narrative review focuses on the role of CaMs in renal osteodystrophy, including their mechanisms and clinical efficacy.

1. Introduction

Chronic kidney disease (CKD) is characterized by a chronic decrease in glomerular filtration and chronic structural damage to the kidney [1]. The decrease in glomerular filtration engenders multiple comorbidities, as well as the loss of nephrons, resulting in complications, including fluid overload, electrolyte imbalance, hypertension, and hormonal dysregulation; such dysregulation leads to complications such as insufficient production of erythropoietin, vitamin D, or phosphaturic hormones [2,3]. Such comorbidities may also result in multiple organ dysfunction syndromes. Renal osteodystrophy—which is characterized by impaired bone remodeling caused by the interplay between vitamin D deficiency, elevated parathyroid hormone (PTH) level, and uremic toxin accumulation due to a decline in glomerular filtration—is common in patients with CKD and end-stage renal disease (ESRD), and its associated vascular calcification and fracture may negatively influence the quality of life or even lead to mortality [4]. Calcimimetics (CaMs), which bind to the calcium-sensing receptors (CaSRs) of the parathyroid gland (PTG), have been used to treat primary or secondary hyperparathyroidism (SHPT). Furthermore, the pleiotropic effects induced by CaMs may engender improvements in other systems apart from the skeletal system [5]. This article presents a review of the pathogenesis of renal osteodystrophy and the pleiotropic effects of CaMs in renal osteodystrophy treatment.

2. Pathogenesis of CKD–Mineral and Bone Disorder: PTH–Vitamin D Interaction Dysregulation as Glomerular Filtration Declines

Phosphate is strongly regulated by phosphaturic hormones, including fibroblast growth factor 23 (FGF-23) and PTH [6,7]. Physiologically, with the assistance of a skeletal tissue buffer, the kidney excretes approximately 800 mg of phosphate ingested daily. The sodium–phosphate cotransporter within the proximal renal tubule transports the excreted phosphate from the glomerulus through an interaction with phosphaturic PTH–FGF-23/klotho complex signaling [8]. As the glomerular filtration rate (GFR) declines, a compensatory increase in phosphaturic hormone production occurs, which regulates the function of the remaining nephrons in maintaining phosphate excretion. To regulate the production of PTH, the CaSRs on the parathyroid glands modulate the synthesis and release of PTH, the concentration of which is inversely correlated with the plasma calcium concentration [9]. Moreover, the concentration of active vitamin D—that is, 1,25-dihydroxy vitamin D (1,25(OH)2D)—derived from the kidney decreases in tubule interstitial cells. Persistent parathyroid gland hyperplasia directly activates the expression of 1-α hydroxylase within the renal tubules to maintain the production of active vitamin D and downstream intestinal absorption in the intestine. However, FGF-23 reduces vitamin D production by directly inhibiting kidney 1-α hydroxylase. The resulting vitamin D deficiency reduces intestinal calcium absorption and CaSRs mediated PTH inhibition.
Physiologically, bone remodeling is essential for bone health. Osteoblasts are the main cells responsible for bone formation, because they interact with each other to form units of bone called osteons [10]. Mesenchymal and skeletal stem cells differentiate into osteoprogenitor cells, pre-osteoblasts, and then osteoblasts. The osteoblasts are embedded into the mineralized bone matrix and, finally, evolve into osteocytes [11]. In bone, osteoclasts remove mineralized bone (bone resorption) to facilitate the formation of a bone matrix by osteoblasts (i.e., bone formation) [12]. Osteocyte-lining cells recruit osteoclasts through an interaction between the receptor activator of nuclear κB (RANK) and the RANK ligand (RANKL). Multiple signals are involved in bone resorption activation, including PTH signals. PTH activates osteoblasts to release the RANKL in mature osteoblasts and suppress its decoy receptor osteoprotegerin (OPG) production in early osteoblasts, which further activates the osteoclasts [13] and induces osteoblastic differentiation and maturation through the downregulation of osteocyte-derived sclerostin and upregulation of Wnt/β–catenin signaling [14]. The interaction of the osteoblasts and osteoclasts governs bone turnover, which is reflected in the osteoid volume per bone volume and the osteoid maturation time [15,16]. In the various stages of CKD, the bone remodeling process may be disrupted by several complications. However, in most cases of renal osteodystrophy with a high bone turnover, the bone is roughly in balance because of the high bone formation accompanied by high resorption rates. For renal osteodystrophy with a low bone turnover rate, such as adynamic bone disease, low bone formation by indolent osteoblasts is accompanied by low bone resorption rates.

2.1. Early-Stage CKD–Mineral and Bone Disorder: Low-Bone-Turnover Disease

In patients with early CKD (stages 1–2), the serum concentration of Wnt signaling inhibitors, such as DKK1 (Dickkopf-1), SOST (sclerostin), and sFRP (secreted frizzled-related proteins), increases because of the release from osteocyte or extraosseous-calcifying tissue and the decrease in renal excretion [17]. The Wnt-signaling inhibitor acts on the osteoblast (OB), resulting in decreased OB viability. In addition, the retention of protein-bound uremic toxins such as indoxyl sulfate (IS)/p-cresol sulfate (PCS) further attenuated the OB and OC viability and function. Uremic osteoporosis was used to describe the effects of protein-bound uremic toxins on bone quality loss with normal bone mass [18].
Coen et al. conducted a cross-sectional study to compare the bone turnover rates between different stages of CKD based on GFR estimates [19]. They observed that the fraction of woven osteoid decreased as the GFR increased. Barreto et al. also demonstrated the difference in bone histomorphometry between CKD stages 2 and 3 and CKD stages 4 and 5. Specifically, they observed that decreases in the osteoid and osteoblast surface areas reflected decreased bone formation and resistance to bone remodeling [20]. Barreto et al. reported that the indoxyl sulfate concentration was the main factor contributing to low bone turnover independent of vitamin D deficiency or PTH concentration [20]. As a protein-bound uremic toxin, indoxyl sulfate can induce PTH resistance in osteoblasts by generating oxidative stress [21]. Indoxyl sulfate also inhibits osteoclast differentiation after cellular entry through organic anion transporters [22]. Furthermore, indoxyl sulfate inhibits osteoclastogenesis by regulating the nuclear factor of activated T-cell cytoplasmic 1 through the aryl hydrocarbon receptor [23]. In addition to uremic toxin accumulation, insulin resistance and deficiency could lead to a decrease in osteoblast proliferation and, thus, degrade bone formation [24,25].

2.2. Advanced CKD–Mineral and Bone Disorder: High-Bone-Turnover Disease Due to SHPT

The development of high-turnover bone disease occurs only later in CKD, when the serum PTH levels can overcome the peripheral resistance to PTH and other inhibitory factors in bone formation. When the renal functions progress (>stage 3) to calcium, phosphate, vitamin D, and PTH dysregulation, patients may present with high or low PTH levels. High PTH levels drive the indolent OB into high viability and function but poor quality in behavior, resulting in both bone quality and quantity loss [18]. In advanced CKD, secondary hyperparathyroidism (SHPT) accelerates bone remodeling and thereby induces osteitis fibrosa, which is characterized by the loss of mineralization and the formation of woven osteoids [26]. Osteitis fibrosa is mainly due to SHPT, and mixed uremic osteodystrophy is characterized by a lack of mineralization most often attributed to vitamin D deficiency. Hypocalcemia mediated by vitamin D deficiency can intensify the sensitivity of the parathyroid glands to vitamin D by reducing CaSR expression. When uremia occurs, the circadian rhythm of the parathyroid glands is dysregulated, and signaling pathways that govern cell proliferation, such as the mammalian target of rapamycin complex 1, are activated by ribosomal protein S6 or the TGF-α/epithelial growth factor receptor [27] and induce parathyroid nodular hyperplasia [28,29,30]. The persistent secretion of PTH is due to peripheral PTH resistance and PTG hyperplasia, which will enhance bone resorption. Increased bone destruction induces the release of calcium and phosphate from bone tissue into the extraosseous tissue. Consequently, calcium apatites may be deposited in the vascular system and promote vascular calcification.

2.3. Advanced CKD–Mineral and Bone Disorder: Low Bone Turnover Disease Due to Medical or Surgical Parathyroidectomy or Aluminum Intoxication

In the medical or surgical treatment of SHPT with the removal of the stimulator of PTH, the bone cells may return to their innate low bone cell viability status, the low bone turnover disorders. We are using renal osteodystrophy for CKD patients’ bone diseases [18].
The low-turnover bone diseases include osteomalacia, aluminum-induced bone disease, and adynamic bone disease (ABD). ABD becomes common in dialysis centers with the increasing use of calcium and potent vitamin D analogs to suppress PTH. PTH resistance is unique to CKD and occurs because of PTH receptor downregulation and osteoblast dysfunction [31] by persistently elevated PTH, excessively low 1,25-dihydroxy vitamin D levels [32], and accumulated protein-bound uremic toxins. In low-turnover bone disease, disturbed bone remodeling occurs with decreased bone resorption and bone formation. Osteoblast apoptosis is a major mechanism that suppresses bone formation [33] [34]. The rate of bone microdamage repair is reduced in low-turnover CKD patients [35] and causes clinical fractures.

3. Role of CaSRs in SHPT

CaSRs are present in different systems, including osseous tissues, cardiomyocytes, parathyroid glands, immune cells, and vascular smooth muscle cells. A CaSR is a group of class C G-protein-coupled receptors that govern calcium homeostasis [36]. PTH release is regulated through minute changes in serum calcium concentrations (1.1–1.3 mM) [37]. The expression of CaSRs is directly regulated by CaSR gene promoters through 1,25(OH)2D, glial cells missing-2 (GCM2), or various extracellular changes in calcium-mediated proinflammatory cytokines [38,39,40,41]. Vitamin D deficiency in CKD induces hypocalcemia as a result of the decreased intestinal absorption of calcium [42]. Prolonged hypocalcemia will decrease the inhibitory effects of CaSRs on PTG and thereby induces parathyroid hyperplasia [43]. Furthermore, hypocalcemia modulates PTH secretion through posttranslational modification [44]. In CKD, PTH nodular hyperplasia may be intensified by the dysregulation of the molecular circadian rhythm of Gcm2 and other proteins, such as cyclin D1 [28]. The associated downregulation of CaSRs reflects further parathyroid gland proliferation [45]. In different calciotropic tissues, such as the parathyroid glands or intestine, CaSR homodimers also regulate cell proliferation, differentiation, or apoptosis (Figure 1).

3.1. Action of CaMs

3.1.1. The Pharmacodynamics of CaMs

CaMs are ligands that potentiate the activity of CaSRs. These ligands can be categorized as agonist (type I) or allosteric (type II) modulators [5]. The binding affinity of intracellular calcium with CaSRs is low (EC50: 1.25 mM); however, CaMs, with their allosteric binding, have a high capacity to rapidly activate CaSRs (EC(50) = 34 nM). Thus, CaMs can reduce PTH secretion by approximately 30-fold compared with calcitonin [46]. The available CaMs include cinacalcet and etelcalcetide [47,48], which bind to the intramembranous and extracellular domains of CaSRs, respectively [49].
CaSRs comprise three domains: a conserved extracellular Venus flytrap (VFT) domain, a cysteine-rich domain, and a heptahelical transmembrane (7TM) domain [50]. Physiologically, calcium can serve as an orthosteric ligand and reduce the secretion of PTH by interacting with the CaSR VFT domain. However, hypercalcemia due to the excessive intestinal absorption of calcium or excessive vitamin D receptor analog use is common in CKD and ESRD [51]. Persistent hypercalcemia induces extraosseous calcification and vascular calcification, which are associated with relatively high mortality in patients with CKD and ESRD [52]. Therefore, CaMs may be less likely to induce hypercalcemia when interacting with CaSRs, and they may serve as positive allosteric modulators in the 7TM domain, which potentiates sensitivity to calcium in the VFT domain [53,54].

3.1.2. The Factors Influencing the Potency of CaMs

Several factors influence the potency of action of CaMs. Hyperphosphatemia, which is common in patients with CKD and ESRD, can inhibit CaSR activity. Phosphate binds to the extracellular domain of CaSRs and, thus, keeps the CaSRs inactive [55]. Furthermore, acidosis influences the effects of CaMs on CaSRs. A study conducted by our research group revealed that the correction of acidosis reduced the PTH concentrations in patients undergoing dialysis [56]. Campion et al. demonstrated that acidosis directly inhibited CaSR activity in a histidine-independent manner [57]. Therefore, the phosphate control, along with the correction of metabolic acidosis, is crucial for improving the potency of CaMs in treatments for patients with CKD and ESRD.

3.1.3. The Actions of CaMs beyond the Parathyroid Gland

In addition to direct CaSR activation, CaMs have other effects, which are listed as follows:

Increasing CaSR Insertion within the Plasma Membrane

Agonist insertion signaling (ADIS) contributes to the characteristic of a high degree of cooperation and absence of functional desensitization that characterizes the obvious CaSR signaling. ADIS occurs through agonist-driven release from pre-plasma membrane compartments and insertion at the plasma membrane, thus increasing and maintaining signaling. This agonist-mediated activation of CaSRs localized in the plasma membrane increases the rate CaSR insertion into the plasma membrane without altering the constitutive endocytosis rate, thereby considerably increasing the maximum signaling response. Prolonged CaSR signaling requires a large pool of intracellular ADIS-mobilizable CaSRs; this pool is sustained by signaling-mediated increases in biosynthesis [58].

Increasing Dihydroxy Vitamin D3 (1,25(OH)2D3) within the Kidney and Parathyroid Glands

CaMs can increase the activity of 1α-hydroxylase enzymes in the kidney and parathyroid cells in CKD. As a result, CaMs may increase the production of 1,25(OH)2D3 in the renal and parathyroid glands if the appropriate native vitamin D was present. By increasing the intracellular concentrations of 1,25(OH)2D3, CaMs can efficiently inhibit intra-gland PTH synthesis and secretion. Furthermore, CaMs can improve CaSR production and engender PTG proliferation inhibition [58].

Inhibiting FGF23 of Osteocytes

CaMs inhibit FGF23 production in bone cells, which may also improve klotho expression in parathyroid cells. Furthermore, the decrease in FGF23 can offset decreases in fibroblast growth factor receptor (FGFR) expression. Therefore, as FGF23 is reduced, the increase in FGFR and klotho expression can effectively inhibit PTH synthesis and secretion through a FGF-signaling cascade.

Attenuating Intestinal Calcium Absorption through TRPV6

CaSRs can be expressed in the intestinal epithelium, and CaSR activation attenuates TRPV6-dependent intestinal calcium absorption. However, chronic CaSR activation reduces the expression of genes involved in Ca2+ absorption. CaSR activation in the basolateral membrane of the intestine directly attenuates local Ca2+ absorption through TRPV6 to prevent hypercalcemia, which may explain how CaMs induce hypocalcemia [59].
Figure 1. Effects of calcimimetics (CaMs) on parathyroid hormone (PTH) synthesis and secretion in chronic kidney disease (CKD). CaM compounds prevent increases in serum PTH concentrations and parathyroid hyperplasia in CKD and severe secondary hyperparathyroidism (SHPT). These changes occur despite decreases in serum 1,25-dihydroxy vitamin D3 (1,25(OH)2D3) and increases in serum phosphate, suggesting that calcium receptors play a prominent role in regulating parathyroid cell proliferation [60].
Figure 1. Effects of calcimimetics (CaMs) on parathyroid hormone (PTH) synthesis and secretion in chronic kidney disease (CKD). CaM compounds prevent increases in serum PTH concentrations and parathyroid hyperplasia in CKD and severe secondary hyperparathyroidism (SHPT). These changes occur despite decreases in serum 1,25-dihydroxy vitamin D3 (1,25(OH)2D3) and increases in serum phosphate, suggesting that calcium receptors play a prominent role in regulating parathyroid cell proliferation [60].
Pharmaceuticals 15 00952 g001

3.2. Effects of CaSRs on Renal Osteodystrophy

As CKD progresses, CaSR expression decreases within different tissues [40,61,62]. In advanced CKD, SHPT can activate PTH signal transduction on osteoblasts and eventually activate osteoclast-related bone resorption. This excessive bone resorption decreases the bone mass through the hyperactivity of osteoclasts. If PTH secretion is reduced, bone resorption can be alleviated. According to human studies, CaMs increased the serum bone formation markers, such as bone-specific alkaline phosphatase, in patients with SHPT or tertiary hyperparathyroidism but reduced the serum bone resorption markers [63,64,65]. In an in vivo study, CaMs maintained bone turnover while reducing PTH secretion [66]. CaMs reduce the release of calcium and phosphorous from the bones by inhibiting osteoclast bone resorption and stimulating osteoblast bone formation. In addition to reducing PTH secretion, CaMs might directly influence bone remodeling in osteoblasts and osteoclasts.

3.2.1. Effect of CaMs on Osteoclasts

Osteoclasts are derived from monocyte/macrophages, and the differentiation process depends on RANK/RANKL signaling [67]. Although several reports have demonstrated that PTH might directly influence osteoclastogenesis, PTH is mainly considered [68] to influence osteoclastogenesis indirectly through RANK/RANKL activity in osteoblasts [69]. Excessive PTH concentrations increase osteoclast differentiation and, thus, aggravate bone resorption [70]. However, a study reported the preservation of osteoclast surfaces after CaM administration [66]. Moreover, the Homer1 complex interacts with CaSRs to enhance the mTORC2 and Akt pathways to maintain osteoblast proliferation and, further, osteocyte proliferation [71]. Zheng et al. demonstrated CaMs to exert a therapeutic effect on osteoclasts through Wnt-10b release, which will promote osteoblastogenesis [72]. CaSR inhibition influences the differentiation of osteoblasts by reducing the expression of essential proteins such as osteocalcin and alkaline phosphatase [73]. Furthermore, excessive PTH secretion may increase endoplasmic reticular stress within osteoclasts and thus activate autophagy [74]. CaMs may induce apoptosis and reduce excessive osteoclastogenesis, resulting from excessive oxidative stress due to SHPT [74]. When a high bone turnover occurs, CaMs induce bone formation and, thus, maintain the bone mass (Figure 2).

3.2.2. Effects of CaMs on Osteoblasts

Wnt pathways are involved in the control of gene expression, cell behavior, cell adhesion, and cell polarity [75]. In these pathways, Wnt signaling inhibits β-catenin degradation, which can regulate the transcription of numerous genes [76]. Wnt signaling is activated through the ligation of Wnt proteins to their respective dimeric cell surface receptors that are composed of seven transmembrane Frizzled proteins and LRP5/6. Upon ligation to their receptors, the cytoplasmic protein disheveled (Dvl) is recruited, phosphorylated, and activated. Dvl activation induces the dissociation of GSK-3β from Axin and leads to GSK-3β inhibition [77]. β-catenin phosphorylation and degradation are then inhibited because of the inactivation of the “destruction complex”. Subsequently, stabilized β-catenin is translocated into the nucleus, leading to changes in the expression levels of various target genes [76].
Rodriguez et al. demonstrated that CaMs could increase the viability of osteoblasts. In nephrectomized mice with continuous PTH infusion, CaMs increased both the bone volume and osteoblast surface area. Moreover, increased nuclear Erk1/2 phosphorylation was observed in the osteoblasts, in addition to the increased expression of bone formation markers such as Runt-related transcription factor 2 (Runx2) and osteocalcin [66]. This finding demonstrates that CaMs influence bone formation through osteoblasts in SHPT (Figure 3).

3.2.3. Effects of CaMs on Bone Quality

Apart from the bone volume loss in SHPT, an unchanged bone volume may influence the bone quality. The bone structure is maintained by the deposition of calcium, inorganic phosphate, and type 1 collagen [78]. The crosslink between inorganic minerals and collagen contributes to bone strength. In CKD, both excessive oxidative stress and vitamin D deficiency influence bone strength by deteriorating collagen deposition. Advanced glycosylated end products may engender excessive oxidative stress in patients with CKD [79], and the deposition of nonenzymatic pentosidine in bone may influence the rate of bone formation and mineral apposition in patients undergoing dialysis [80]. In addition to inhibiting osteoclastogenesis, CaMs may affect bone strength. In adenine-induced CKD, the hardness and elastic modulus of bone were restored after cinacalcet supplementation [74]. Although the direct link between CaSRs and the expression of proteins associated with bone strength, such as type 1 collagen, is unclear, CaMs reduce oxidative stress in osteoclasts; therefore, they might be helpful in bone strength restoration.

3.2.4. Role in CaMs in Conjunction with Vitamin D for Bone Remodeling

In low-bone-turnover disease, vitamin D deficiency directly reduces CaSR expression by downregulating these receptors. In the early stages of CKD, the therapeutic effects of CaMs may be less prominent than those of calcitriol in maintaining bone mass [81]. Finch et al. demonstrated that, in uremic rats used to investigate CKD Stage 3 or 4 in humans, the osteoid surface areas decreased with bone formation. This may be explained by the relatively low bone turnover in the early stages of CKD due to uremic toxin accumulation [20]. In other uremic animal models, CaMs played a therapeutic role in reversing high-bone turnover, along with, or independent of, PTH secretion control [66,82]. These in vivo studies have demonstrated that cinacalcet increased distal metaphysis more than calcitriol did in uremic rats [82]. Cinacalcet can induce the release of Wnt-10b from osteoclasts, which is associated with bone formation within osteoblasts [72]. Furthermore, cinacalcet could improve vitamin D activity during osteoclast differentiation. Allard et al. demonstrated that a therapeutic dosage of 1,25(OH)2D directly inhibited osteoclastogenesis without modulating bone resorption activity [83]. In CKD, the inhibitory effects of vitamin D were reported to decrease with the severity of CKD, and the resistance of osteoclasts to vitamin D could be reversed by cinacalcet [84]. According to these findings, CaMs, along with vitamin D, may reverse high-bone turnover by maintaining osteoblasts and inhibiting osteoclastogenesis in advanced CKD.

3.3. Effects of CaSRs on Vascular Calcification

Vascular calcification is common in advanced CKD because of renal osteodystrophy. High-bone turnover disease induces excessive bone resorption and thus increases calcium and phosphate deposition in extraosseous tissues.
CaSRs are distributed within the endothelium, smooth muscle cells, and mesenchymal stem cells. Extracellular calcium can directly induce smooth muscle cell proliferation through mitogen-activated protein kinase (MAPK) kinase 1 [85]. A study reported that, in the uremic milieu, CaSR expression in the aorta decreased compared with that in a healthy control group [86]. Excessive calcium concentrations can downregulate the expression of CaSRs and the mineralization of vascular smooth muscle cells [87]. Decreased CaSR expression was reported to be associated with aortic calcification, with CaSRs serving as a proinflammatory cytokine. Mary et al. demonstrated that the expression of CaSRs in monocytes decreased in the uremic milieu, and the reduction of CaSR expression within the monocytes intensified vascular calcification [62]. In an animal model of spontaneous hypertension, CaSR expression was negatively associated with the plasma concentration of angiotensin II but positively associated with renin levels [88]. CaSRs may be distributed on vascular smooth muscle cells [85].
According to in vivo studies, CaMs alleviated vascular calcification and reduced PTH secretion in uremic rats [82]. Henley et al. compared the effectiveness of CaMs and calcitriol in alleviating vascular calcification. Calcitriol increased calcium and phosphate burdens by increasing calcium and phosphate reabsorption. As mentioned, excessive calcium intensifies vascular smooth muscle cell calcification by downregulating CaSRs [87]. Furthermore, the matrix Gla protein governing calciprotein particles is released under the influence of CaMs on vascular smooth muscle cells [89,90]. Additionally, nitric oxide might be directly released under CaM exposure, which could induce vasodilation and thus reduce the arterial pressure [91]. From the aspects above, the modulation of CaMs might play a role in alleviating vascular calcification.

3.4. CaSRs and Left Ventricular Remodeling

CaSRs may be distributed within cardiomyocytes. CaSRs within cardiomyocytes induce the release of calcium from the intracellular endo/sarcoplasmic reticulum into the cytoplasm and thus engender cardiomyocyte apoptosis through mitochondrial damage [92]. During congestive heart failure, CaSR expression increases, along with mitochondrial membrane potential and ultrastructural change. Increased calcium/calmodulin-dependent protein kinase II and calcineurin expression heightens cardiomyocyte apoptosis and thus engenders further cardiac remodeling [93]. CaSRs are essential in modulating the renin–angiotensin–aldosterone system, and they govern juxtaglomerular renin release [94]. CaMs alleviate aortic calcification directly in the vascular smooth muscle cells by inhibiting MEK1 [85]. In contrast to extracellular calcium, CaMs may not influence CaSR expression [95]. CaSR activation could induce myocardial apoptosis by modulating the mitochondrial dynamics. Hong et al. demonstrated that the myocardial expression of CaSRs increased in an animal model of spontaneous hypertension [96]. In the animal model, mitochondrial fission increased in the cardiomyocytes, and mitochondrial swelling, disorganized cristae, and a loss of normal striation activated downstream apoptosis. Generally, CaSR inhibitor administration can alleviate cardiac hypertrophy through multiple mechanisms. CaMs may play a protective role in alleviating changes in the mitochondrial membranous potential and the release of calcium/calmodulin II [93]. The direct modulation of CaSRs on the renin–angiotensin–aldosterone system (RAAS) also alters cardiac remodeling [97]. Additionally, CaM administration can reduce unwanted remodeling by ameliorating autophagy [98]. Therefore, CaMs should be a therapeutic choice under conditions involving vascular calcification.

3.5. Role of CaSRs in Renal Tubules and Vasculature

CaSRs are distributed within different tissues, including those of the heart, pancreas, and renal tubules. In proximal tubular cells, CaSRs are located on the luminal side. Filtered calcium on the luminal side directly activates CaSRs, thus leading to sodium/proton exchanger isoform 3 activation within the luminal side. Consequently, protons generated within the renal tubules can be extruded into the lumen and engender a net reabsorption of bicarbonate [99]. In the thick ascending tubules, CaSRs are located on the basolateral side. CaSR activation modulates the intracellular calcineurin–nuclear factor of activated T-cell signaling, which reduces paracellular calcium reabsorption by downregulating claudin expression [100]. Rothe et al. reported that the CaSR polymorphism Arg990Gly was associated with a more favorable response to CaSRs in transfected HEK cells [101]. In addition, CaSRs within the juxtaglomerular apparatus can modulate the release of renin. The interaction between CaSRs and RAAS may directly modulate the Na+/Ca2+ exchanger [102]. Ortiz-Capisano et al. revealed that CaM activation increased the intracellular calcium release through phospholipase C/inositol triphosphate3 and RyR activation; hence, the release of vasoactive renin could be inhibited [94]. The chronic activation of CaSRs could alter RAAS and lead to hyperreninemia [103]. The interaction between CaSRs and RAAS may be essential for treating cardiac remodeling in patients with CKD–mineral and bone disorder.

4. Clinical Efficacy of CaMs

4.1. CaMs in Renal Osteodystrophy/Fracture

In patients with primary hyperparathyroidism, CaMs provide a therapeutic option. Peacock et al. demonstrated that cinacalcet administration could lower the PTH concentrations in patients with primary hyperparathyroidism. In addition, bone resorption and formation markers both increased after treatment, without any changes occurring in the bone mineral density. Therefore, cinacalcet can successfully reduce PTH concentrations and improve the serum calcium and phosphate control [104], and hypercalcemia can be corrected under such treatment [105]. Moreover, the OPTIMA study, which compared the effects of cinacalcet with those of vitamin D, demonstrated that cinacalcet provided a more favorable PTH control than vitamin D [106]. Therefore, the combination of vitamin D with CaMs can be a rational strategy for treating or controlling SHPT. The ACHIEVE study revealed that CaMs, along with low-dose vitamin D, provided better control of SHPT and decreased PTH levels compared with the other treatment approaches [107]. The DUET trial demonstrated that CaMs, along with vitamin D, exhibited superior PTH control than vitamin D alone or CaMs with calcium supplements [108].
Scholars have demonstrated the clinical efficacy of CaMs concerning bone mineral density in patients with SHPT. For example, Shigematsu et al. reported that, during a 1-year cinacalcet treatment period for patients with hemodialysis, the bone resorption markers decreased with the transient elevation of bone-specific alkaline phosphatase [109]. Tsuruta et al. performed a clinical trial and reported that, in patients with HD with SHPT, a 1-year administration of cinacalcet engendered bone mineral density improvement, which was negatively correlated with variations in the PTH concentrations [110]. Specifically, the observed bone mineral density increased and serum alkaline phosphatase concentrations deteriorated. Hung et al. indicated that cinacalcet (25 mg/day) supplementation for 6 months increased bone mineral density in patients with HD. Among the patients who demonstrated a response to treatment, bone formation markers such as procollagen type I N-terminal pro-peptide (PINP) increased, but bone resorption markers such as Tartrate-resistant acid phosphatase isoform 5 regressed. Furthermore, plasma Wnt-10b increased in the cinacalcet-responsive group [65]. Ruderman et al. demonstrated the negative effects of cinacalcet withdrawal on the plasma parameters. Based on the aforementioned results, in SHPT with clear bone resorption, CaMs play a therapeutic role in maintaining the bone mass.

4.2. CaMs in Cardiovascular Mortality

The role of CaMs in cardiovascular disease was first elucidated in the EVLOVE study. In this study, the 21.2-month administration of cinacalcet did not lower the overall cardiovascular mortality in patients with HD with moderate-to-severe SHPT [111]. In a secondary investigation, the number of cardiovascular events mediated by nonatherosclerotic disease and the number of events of sudden death or congestive heart failure decreased in patients who received cinacalcet [48]. Furthermore, CaSR-associated polymorphisms influence the effectiveness of PTH reduction. For example, the rs9470 polymorphism was associated with more favorable reductions of PTH concentrations in patients who received cinacalcet treatment [112]. Another study revealed that a CaSR polymorphism was associated with dyslipidemia in patients with HD; however, the clinical relevance of this finding is unknown [113]. Accordingly, CaMs may play a therapeutic role in alleviating cardiovascular disease associated with non-dyslipidemia-related risk factors.
Ruderman et al. observed that primary calciprotein particles in plasma, which are negatively associated with vascular calcification and PTH concentrations increased 1 year after the cessation of cinacalcet treatment in patients with ESRD [114]. Although the clinical relevance of this change in calciprotein particles after cinacalcet withdrawal remains unclear, the production of primary calciprotein particles may alleviate the calcification burden after the cessation of cinacalcet treatment.

4.3. CaMs Compared with Parathyroidectomy

In SHPT, the diffusion of hyperplasia in the parathyroid glands transforms polyclonal hyperplasia into monoclonal hyperplasia. Thus, single nodular hyperplasia is common among patients with advanced CKD or those receiving dialysis [115]. Subtotal or total parathyroidectomy is a curative strategy for treating uncontrolled SHPT or SHPT with severe calciphylaxis. Only a few clinical trials have compared the efficacy of CaMs and parathyroidectomy. Keutgen et al. conducted a comparative study, in which the endpoint was bone mineral density improvement. The surgical group demonstrated greater PTH improvement than the CaM group; this PTH improvement was associated with a greater bone mineral density improvement regardless of the choice of treatment [116]. From the study by Cruzado et al., parathyroidectomy provided a better control of hypercalcemia than cinacalcet [117]. Almadén et al. reported a decreased expression of CaSRs in hyperplastic nodules and thus recommended surgical management in tertiary hyperplasia [118], which may result in lower responses to cinacalcet treatment. Nevertheless, surgical intervention may induce a relatively high incidence of hypocalcemia. To achieve sustained PTH secretion control and pleiotropic effects through CaSRs in SHPT, the application of CaMs is an essential therapeutic strategy.

4.4. Current Guidelines on Treating CKD–MBD with Calcimimetics

Based on the KDIGO guidelines on the treatment of CKD–MBD, secondary hyperparathyroidism in ESRD subjects could be treated by the CaMs, calcitriol, or vitamin D analogs or a combination of CaMs with calcitriol or vitamin D analogs [1]. Both vitamin D analogs and CaMs are acceptable options for treating SHPT. For patients with both hypercalcemia and hyperphosphatemia, CaMs is a reasonable choice in lowering the PTH level in comparison with the vitamin D analog. As hypocalcemia develops, the CaMs should be discontinued, and the vitamin D or vitamin D analog could be the agent lowering the PTH [119]. Recent clinical trials about the new generation of CaMs have provided the PTH-lowering effect. However, the effect of fracture prevention and the cardiovascular outcome in other CaMs should be noticed in the future.

5. Limitation of the CaMs in CKD–MBD

Based on the evidence from clinical trials such as the EVOLVE study, CaMs provided a sustained PTH-lowering effect and therefore might contribute to the improvement in cardiovascular mortalities in previous sections. However, there are still several limitations to use in CaMs. First, the serum calcium concentration should be monitored continuously to avoid hypocalcemia. Second, the cost of CaMs might pose an economic burden for the patients. Third, the trials for a head-to-head comparison between CaMs and parathyroidectomy are still lacking. Further evidence is needed to compare the efficacy between CaMs and parathyroidectomy.

6. Conclusions

CKD–MBD is a severe complication in patients with CKD and ESRD. Decreased renal phosphate excretion and SHPT increase bone resorption and not bone formation, thus intensifying the vascular calcification and left ventricular remodeling. Decreases in CaSR expression within the parathyroid glands and vasculature potentiate the hazards of SHPT because of the corresponding vitamin D deficiency and dysregulation of the proteins for cell cycling. To alleviate these effects, CaMs can be combined with vitamin D to treat CKD–MBD.

Author Contributions

Y.-C.H. and C.-M.Z. drafted the manuscript. H.-W.C. provided the figure design and correctness of the figures. C.-L.L. and W.-C.L. completed the data collection and the references. K.-C.L. conducted the whole structure of the manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

Abbreviations

ADISagonist-driven insertional signaling
CaMsCalcimimetics
CaSRsCalcium-sensing receptors
CKDChronic kidney disease
CKD–MBDChronic kidney disease–mineral and bone disorder
ER stressEndoplasmic reticulum stress
ERKExtracellular signal-regulated protein kinases
ESRDend stage renal disease
FGF23fibroblast growth factor 23
GCM2glial cells missing-2
GFRglomerular filtration rate
JNKc-Jun N-terminal kinase
MAPKMitogen-activated protein kinase
P1NPprocollagen type I N-terminal pro-peptide (PINP)
PTGParathyroid gland
PTHparathyroid hormone
RAASrenin–angiotensin–aldosterone system
Runx2Runt-related transcription factor 2
SHPTSecondary hyperparathyroidism
SRF-REserum response factor-response element
TGFαTransforming growth factor-α
TRPV6Transient Receptor Potential Vanilloid subfamily member 6
VFTVenus flytrap
WntWingless-related integration site

References

  1. Kidney Disease: Improving Global Outcomes (KDIGO) CKD-MBD Update Work Group. KDIGO 2017 Clinical Practice Guideline Update for the Diagnosis, Evaluation, Prevention, and Treatment of Chronic Kidney Disease-Mineral and Bone Disorder (CKD-MBD). Kidney Int. Suppl. 2011 2017, 7, 1–59. [Google Scholar] [CrossRef] [Green Version]
  2. Patel, N.M.; Gutiérrez, O.M.; Andress, D.L.; Coyne, D.W.; Levin, A.; Wolf, M. Vitamin D deficiency and anemia in early chronic kidney disease. Kidney Int. 2010, 77, 715–720. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. Cannata-Andía, J.B.; Carrera, F. The Pathophysiology of Secondary Hyperparathyroidism and the Consequences of Uncontrolled Mineral Metabolism in Chronic Kidney Disease: The Role of COSMOS. NDT Plus 2008, 1, i2–i6. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. Hou, Y.C.; Lu, C.L.; Lu, K.C. Mineral bone disorders in chronic kidney disease. Nephrology 2018, 23 (Suppl. S4), 88–94. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  5. Nemeth, E.F. Allosteric modulators of the extracellular calcium receptor. Drug Discov. Today. Technol. 2013, 10, e277–e284. [Google Scholar] [CrossRef] [PubMed]
  6. Liu, S.; Tang, W.; Zhou, J.; Stubbs, J.R.; Luo, Q.; Pi, M.; Quarles, L.D. Fibroblast growth factor 23 is a counter-regulatory phosphaturic hormone for vitamin D. J. Am. Soc. Nephrol. JASN 2006, 17, 1305–1315. [Google Scholar] [CrossRef] [Green Version]
  7. Onsgard-Meyer, M.J.; Berndt, T.J.; Khraibi, A.A.; Knox, F.G. Phosphaturic effect of parathyroid hormone in the spontaneously hypertensive rat. Am. J. Physiol. 1994, 267, R78–R83. [Google Scholar] [CrossRef]
  8. Razzaque, M.S. The FGF23-Klotho axis: Endocrine regulation of phosphate homeostasis. Nat. Rev. Endocrinol. 2009, 5, 611–619. [Google Scholar] [CrossRef] [Green Version]
  9. Hannan, F.M.; Kallay, E.; Chang, W.; Brandi, M.L.; Thakker, R.V. The calcium-sensing receptor in physiology and in calcitropic and noncalcitropic diseases. Nat. Rev. Endocrinol. 2018, 15, 33–51. [Google Scholar] [CrossRef]
  10. Long, F. Building strong bones: Molecular regulation of the osteoblast lineage. Nat. Rev. Mol. Cell Biol. 2011, 13, 27–38. [Google Scholar] [CrossRef] [PubMed]
  11. Compton, J.T.; Lee, F.Y. A review of osteocyte function and the emerging importance of sclerostin. J. Bone Jt. Surg. Am. Vol. 2014, 96, 1659–1668. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  12. Hadjidakis, D.J.; Androulakis, I.I. Bone remodeling. Ann. N. Y. Acad. Sci. 2006, 1092, 385–396. [Google Scholar] [CrossRef]
  13. Huang, J.C.; Sakata, T.; Pfleger, L.L.; Bencsik, M.; Halloran, B.P.; Bikle, D.D.; Nissenson, R.A. PTH differentially regulates expression of RANKL and OPG. J. Bone Miner. Res. Off. J. Am. Soc. Bone Miner. Res. 2004, 19, 235–244. [Google Scholar] [CrossRef]
  14. Silva, B.C.; Bilezikian, J.P. Parathyroid hormone: Anabolic and catabolic actions on the skeleton. Curr. Opin. Pharmacol. 2015, 22, 41–50. [Google Scholar] [CrossRef] [Green Version]
  15. Rauch, F. CHAPTER 15—Bone Histomorphometry. In Pediatric Bone; Glorieux, F.H., Pettifor, J.M., JÜPpner, H., Eds.; Academic Press: San Diego, CA, USA, 2003; pp. 359–374.I–VII. [Google Scholar] [CrossRef]
  16. Ott, S.M. Histomorphometric measurements of bone turnover, mineralization, and volume. Clin. J. Am. Soc. Nephrol. 2008, 3 (Suppl. S3), S151–S156. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  17. Evenepoel, P.; D’Haese, P.; Brandenburg, V. Sclerostin and DKK1: New players in renal bone and vascular disease. Kidney Int. 2015, 88, 235–240. [Google Scholar] [CrossRef] [PubMed]
  18. Zheng, C.M.; Zheng, J.Q.; Wu, C.C.; Lu, C.L.; Shyu, J.F.; Yung-Ho, H.; Wu, M.Y.; Chiu, I.J.; Wang, Y.H.; Lin, Y.F.; et al. Bone loss in chronic kidney disease: Quantity or quality? Bone 2016, 87, 57–70. [Google Scholar] [CrossRef] [PubMed]
  19. Coen, G.; Mazzaferro, S.; Ballanti, P.; Sardella, D.; Chicca, S.; Manni, M.; Bonucci, E.; Taggi, F. Renal bone disease in 76 patients with varying degrees of predialysis chronic renal failure: A cross-sectional study. Nephrol. Dial. Transplant. 1996, 11, 813–819. [Google Scholar] [CrossRef] [PubMed]
  20. Barreto, F.C.; Barreto, D.V.; Canziani, M.E.; Tomiyama, C.; Higa, A.; Mozar, A.; Glorieux, G.; Vanholder, R.; Massy, Z.; de Carvalho, A.B. Association between indoxyl sulfate and bone histomorphometry in pre-dialysis chronic kidney disease patients. J. Bras. Nefrol. 2014, 36, 289–296. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  21. Nii-Kono, T.; Iwasaki, Y.; Uchida, M.; Fujieda, A.; Hosokawa, A.; Motojima, M.; Yamato, H.; Kurokawa, K.; Fukagawa, M. Indoxyl sulfate induces skeletal resistance to parathyroid hormone in cultured osteoblastic cells. Kidney Int. 2007, 71, 738–743. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  22. Mozar, A.; Louvet, L.; Godin, C.; Mentaverri, R.; Brazier, M.; Kamel, S.; Massy, Z.A. Indoxyl sulphate inhibits osteoclast differentiation and function. Nephrol. Dial. Transplant. 2012, 27, 2176–2181. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  23. Liu, W.-C.; Shyu, J.-F.; Lim, P.S.; Fang, T.-C.; Lu, C.-L.; Zheng, C.-M.; Hou, Y.-C.; Wu, C.-C.; Lin, Y.-F.; Lu, K.-C. Concentration and Duration of Indoxyl Sulfate Exposure Affects Osteoclastogenesis by Regulating NFATc1 via Aryl Hydrocarbon Receptor. Int. J. Mol. Sci. 2020, 21, 3486. [Google Scholar] [CrossRef] [PubMed]
  24. Niu, T.; Rosen, C.J. The insulin-like growth factor-I gene and osteoporosis: A critical appraisal. Gene 2005, 361, 38–56. [Google Scholar] [CrossRef] [PubMed]
  25. Kawai, M.; Sousa, K.M.; MacDougald, O.A.; Rosen, C.J. The many facets of PPARgamma: Novel insights for the skeleton. Am. J. Physiol. Endocrinol. Metab. 2010, 299, E3–E9. [Google Scholar] [CrossRef] [PubMed]
  26. Drüeke, T.B.; Massy, Z.A. Changing bone patterns with progression of chronic kidney disease. Kidney Int. 2016, 89, 289–302. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  27. Volovelsky, O.; Cohen, G.; Kenig, A.; Wasserman, G.; Dreazen, A.; Meyuhas, O.; Silver, J.; Naveh-Many, T. Phosphorylation of Ribosomal Protein S6 Mediates Mammalian Target of Rapamycin Complex 1-Induced Parathyroid Cell Proliferation in Secondary Hyperparathyroidism. J. Am. Soc. Nephrol. JASN 2016, 27, 1091–1101. [Google Scholar] [CrossRef] [Green Version]
  28. Egstrand, S.; Nordholm, A.; Morevati, M.; Mace, M.L.; Hassan, A.; Naveh-Many, T.; Rukov, J.L.; Gravesen, E.; Olgaard, K.; Lewin, E. A molecular circadian clock operates in the parathyroid gland and is disturbed in chronic kidney disease associated bone and mineral disorder. Kidney Int. 2020, 98, 1461–1475. [Google Scholar] [CrossRef]
  29. Arcidiacono, M.V.; Sato, T.; Alvarez-Hernandez, D.; Yang, J.; Tokumoto, M.; Gonzalez-Suarez, I.; Lu, Y.; Tominaga, Y.; Cannata-Andia, J.; Slatopolsky, E.; et al. EGFR activation increases parathyroid hyperplasia and calcitriol resistance in kidney disease. J. Am. Soc. Nephrol. JASN 2008, 19, 310–320. [Google Scholar] [CrossRef] [Green Version]
  30. Dusso, A.S.; Pavlopoulos, T.; Naumovich, L.; Lu, Y.; Finch, J.; Brown, A.J.; Morrissey, J.; Slatopolsky, E. p21(WAF1) and transforming growth factor-alpha mediate dietary phosphate regulation of parathyroid cell growth. Kidney Int. 2001, 59, 855–865. [Google Scholar] [CrossRef] [Green Version]
  31. Iwasaki-Ishizuka, Y.; Yamato, H.; Nii-Kono, T.; Kurokawa, K.; Fukagawa, M. Downregulation of parathyroid hormone receptor gene expression and osteoblastic dysfunction associated with skeletal resistance to parathyroid hormone in a rat model of renal failure with low turnover bone. Nephrol. Dial. Transplant. 2005, 20, 1904–1911. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  32. Malluche, H.H.; Monier-Faugere, M.C.; Koszewski, N.J. Use and indication of vitamin D and vitamin D analogues in patients with renal bone disease. Nephrol. Dial. Transplant. 2002, 17 (Suppl. S10), 6–9. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  33. Andress, D.L. Adynamic bone in patients with chronic kidney disease. Kidney Int. 2008, 73, 1345–1354. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  34. Haarhaus, M.; Evenepoel, P. Differentiating the causes of adynamic bone in advanced chronic kidney disease informs osteoporosis treatment. Kidney Int. 2021, 100, 546–558. [Google Scholar] [CrossRef] [PubMed]
  35. Malluche, H.H.; Porter, D.S.; Pienkowski, D. Evaluating bone quality in patients with chronic kidney disease. Nat. Rev. Nephrol. 2013, 9, 671–680. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  36. Ward, D.T.; Riccardi, D. New concepts in calcium-sensing receptor pharmacology and signalling. Br. J. Pharmacol. 2012, 165, 35–48. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  37. Brown, E.M.; MacLeod, R.J. Extracellular calcium sensing and extracellular calcium signaling. Physiol. Rev. 2001, 81, 239–297. [Google Scholar] [CrossRef] [PubMed]
  38. Canaff, L.; Hendy, G.N. Calcium-sensing receptor gene transcription is up-regulated by the proinflammatory cytokine, interleukin-1beta. Role of the NF-kappaB PATHWAY and kappaB elements. J. Biol. Chem. 2005, 280, 14177–14188. [Google Scholar] [CrossRef] [Green Version]
  39. Canaff, L.; Zhou, X.; Mosesova, I.; Cole, D.E.; Hendy, G.N. Glial cells missing-2 (GCM2) transactivates the calcium-sensing receptor gene: Effect of a dominant-negative GCM2 mutant associated with autosomal dominant hypoparathyroidism. Hum. Mutat. 2009, 30, 85–92. [Google Scholar] [CrossRef] [PubMed]
  40. Hendy, G.N.; Canaff, L. Calcium-Sensing Receptor Gene: Regulation of Expression. Front. Physiol. 2016, 7, 394. [Google Scholar] [CrossRef] [Green Version]
  41. Brown, A.J.; Zhong, M.; Finch, J.; Ritter, C.; McCracken, R.; Morrissey, J.; Slatopolsky, E. Rat calcium-sensing receptor is regulated by vitamin D but not by calcium. Am. J. Physiol. 1996, 270, F454–F460. [Google Scholar] [CrossRef] [PubMed]
  42. Christakos, S.; Dhawan, P.; Porta, A.; Mady, L.J.; Seth, T. Vitamin D and intestinal calcium absorption. Mol. Cell. Endocrinol. 2011, 347, 25–29. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  43. Silver, J. Pathogenesis of parathyroid dysfunction in end-stage renal disease. Adv. Ren. Replace. Ther. 2002, 9, 159–167. [Google Scholar] [CrossRef] [PubMed]
  44. Naveh-Many, T.; Rahamimov, R.; Livni, N.; Silver, J. Parathyroid cell proliferation in normal and chronic renal failure rats. The effects of calcium, phosphate, and vitamin D. J. Clin. Investig. 1995, 96, 1786–1793. [Google Scholar] [CrossRef] [PubMed]
  45. Lee, H.J.; Seo, U.H.; Kim, W.Y.; Woo, S.U.; Lee, J.B. Calcium-sensing receptor and apoptosis in parathyroid hyperplasia of patients with secondary hyperparathyroidism. J. Int. Med. Res. 2013, 41, 97–105. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  46. Nemeth, E.F.; Heaton, W.H.; Miller, M.; Fox, J.; Balandrin, M.F.; Van Wagenen, B.C.; Colloton, M.; Karbon, W.; Scherrer, J.; Shatzen, E.; et al. Pharmacodynamics of the type II calcimimetic compound cinacalcet HCl. J. Pharmacol. Exp. Ther. 2004, 308, 627–635. [Google Scholar] [CrossRef] [Green Version]
  47. Block, G.A.; Bushinsky, D.A.; Cunningham, J.; Drueke, T.B.; Ketteler, M.; Kewalramani, R.; Martin, K.J.; Mix, T.C.; Moe, S.M.; Patel, U.D.; et al. Effect of Etelcalcetide vs Placebo on Serum Parathyroid Hormone in Patients Receiving Hemodialysis With Secondary Hyperparathyroidism: Two Randomized Clinical Trials. JAMA 2017, 317, 146–155. [Google Scholar] [CrossRef] [PubMed]
  48. Wheeler, D.C.; London, G.M.; Parfrey, P.S.; Block, G.A.; Correa-Rotter, R.; Dehmel, B.; Drüeke, T.B.; Floege, J.; Kubo, Y.; Mahaffey, K.W.; et al. Effects of cinacalcet on atherosclerotic and nonatherosclerotic cardiovascular events in patients receiving hemodialysis: The EValuation Of Cinacalcet HCl Therapy to Lower CardioVascular Events (EVOLVE) trial. J. Am. Heart Assoc. 2014, 3, e001363. [Google Scholar] [CrossRef] [Green Version]
  49. Bellasi, A.; Cozzolino, M.; Malberti, F.; Cancarini, G.; Esposito, C.; Guastoni, C.M.; Ondei, P.; Pontoriero, G.; Teatini, U.; Vezzoli, G.; et al. New scenarios in secondary hyperparathyroidism: Etelcalcetide. Position paper of working group on CKD-MBD of the Italian Society of Nephrology. J. Nephrol. 2020, 33, 211–221. [Google Scholar] [CrossRef] [Green Version]
  50. Ellaithy, A.; Gonzalez-Maeso, J.; Logothetis, D.A.; Levitz, J. Structural and Biophysical Mechanisms of Class C G Protein-Coupled Receptor Function. Trends Biochem. Sci. 2020, 45, 1049–1064. [Google Scholar] [CrossRef]
  51. Negrea, L. Active Vitamin D in Chronic Kidney Disease: Getting Right Back Where We Started from? Kidney Dis. 2019, 5, 59–68. [Google Scholar] [CrossRef] [PubMed]
  52. Floege, J.; Ketteler, M. Vascular calcification in patients with end-stage renal disease. Nephrol. Dial. Transplant. 2004, 19, v59–v66. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  53. Friedl, C.; Zitt, E. Role of etelcalcetide in the management of secondary hyperparathyroidism in hemodialysis patients: A review on current data and place in therapy. Drug Des. Dev. Ther. 2018, 12, 1589–1598. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  54. Steddon, S.J.; Cunningham, J. Calcimimetics and calcilytics--fooling the calcium receptor. Lancet 2005, 365, 2237–2239. [Google Scholar] [CrossRef]
  55. Centeno, P.P.; Herberger, A.; Mun, H.-C.; Tu, C.; Nemeth, E.F.; Chang, W.; Conigrave, A.D.; Ward, D.T. Phosphate acts directly on the calcium-sensing receptor to stimulate parathyroid hormone secretion. Nat. Commun. 2019, 10, 4693. [Google Scholar] [CrossRef] [PubMed]
  56. Lu, K.C.; Shieh, S.D.; Li, B.L.; Chu, P.; Jan, S.Y.; Lin, Y.F. Rapid correction of metabolic acidosis in chronic renal failure: Effect on parathyroid hormone activity. Nephron 1994, 67, 419–424. [Google Scholar] [CrossRef] [PubMed]
  57. Campion, K.L.; McCormick, W.D.; Warwicker, J.; Khayat, M.E.B.; Atkinson-Dell, R.; Steward, M.C.; Delbridge, L.W.; Mun, H.-C.; Conigrave, A.D.; Ward, D.T. Pathophysiologic Changes in Extracellular pH Modulate Parathyroid Calcium-Sensing Receptor Activity and Secretion <em>via</em> a Histidine-Independent Mechanism. J. Am. Soc. Nephrol. 2015, 26, 2163. [Google Scholar] [CrossRef] [Green Version]
  58. Breitwieser, G.E. Minireview: The intimate link between calcium sensing receptor trafficking and signaling: Implications for disorders of calcium homeostasis. Mol. Endocrinol. 2012, 26, 1482–1495. [Google Scholar] [CrossRef] [Green Version]
  59. Lee, J.J.; Liu, X.; O’Neill, D.; Beggs, M.R.; Weissgerber, P.; Flockerzi, V.; Chen, X.Z.; Dimke, H.; Alexander, R.T. Activation of the calcium sensing receptor attenuates TRPV6-dependent intestinal calcium absorption. JCI Insight 2019, 5, e128013. [Google Scholar] [CrossRef] [Green Version]
  60. Wada, M.; Nagano, N.; Furuya, Y.; Chin, J.; Nemeth, E.F.; Fox, J. Calcimimetic NPS R-568 prevents parathyroid hyperplasia in rats with severe secondary hyperparathyroidism. Kidney Int. 2000, 57, 50–58. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  61. Uchiyama, T.; Ohkido, I.; Nakashima, A.; Saito, Y.; Okabe, M.; Yokoo, T. Severe chronic kidney disease environment reduced calcium-sensing receptor expression in parathyroid glands of adenine-induced rats even without high phosphorus diet. BMC Nephrol. 2020, 21, 219. [Google Scholar] [CrossRef]
  62. Mary, A.; Objois, T.; Brazier, M.; Bennis, Y.; Boudot, C.; Lenglet, G.; Paccou, J.; Bugnicourt, J.-M.; Choukroun, G.; Drueke, T.B.; et al. Decreased monocyte calcium sensing receptor expression in patients with chronic kidney disease is associated with impaired monocyte ability to reduce vascular calcification. Kidney Int. 2021, 99, 1382–1391. [Google Scholar] [CrossRef] [PubMed]
  63. Schwarz, A.; Merkel, S.; Leitolf, H.; Haller, H. The effect of cinacalcet on bone remodeling and renal function in transplant patients with persistent hyperparathyroidism. Transplantation 2011, 91, 560–565. [Google Scholar] [CrossRef] [PubMed]
  64. Yano, S.; Suzuki, K.; Sumi, M.; Tokumoto, A.; Shigeno, K.; Himeno, Y.; Sugimoto, T. Bone metabolism after cinacalcet administration in patients with secondary hyperparathyroidism. J. Bone Miner. Metab. 2010, 28, 49–54. [Google Scholar] [CrossRef]
  65. Hung, K.C.; Chang, J.F. Therapeutic Effect of Calcimimetics on Osteoclast-Osteoblast Crosslink in Chronic Kidney Disease and Mineral Bone Disease. Int. J. Mol. Sci. 2020, 21, 8712. [Google Scholar] [CrossRef] [PubMed]
  66. Díaz-Tocados, J.M.; Rodríguez-Ortiz, M.E.; Almadén, Y.; Pineda, C.; Martínez-Moreno, J.M.; Herencia, C.; Vergara, N.; Pendón-Ruiz de Mier, M.V.; Santamaría, R.; Rodelo-Haad, C.; et al. Calcimimetics maintain bone turnover in uremic rats despite the concomitant decrease in parathyroid hormone concentration. Kidney Int. 2019, 95, 1064–1078. [Google Scholar] [CrossRef] [PubMed]
  67. Yasuda, H.; Shima, N.; Nakagawa, N.; Yamaguchi, K.; Kinosaki, M.; Mochizuki, S.; Tomoyasu, A.; Yano, K.; Goto, M.; Murakami, A.; et al. Osteoclast differentiation factor is a ligand for osteoprotegerin/osteoclastogenesis-inhibitory factor and is identical to TRANCE/RANKL. Proc. Natl. Acad. Sci. USA 1998, 95, 3597–3602. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  68. Liu, S.; Zhu, W.; Li, S.; Ma, J.; Zhang, H.; Li, Z.; Zhang, L.; Zhang, B.; Li, Z.; Liang, X.; et al. Bovine parathyroid hormone enhances osteoclast bone resorption by modulating V-ATPase through PTH1R. Int. J. Mol. Med. 2016, 37, 284–292. [Google Scholar] [CrossRef] [Green Version]
  69. Del Fattore, A.; Teti, A.; Rucci, N. Osteoclast receptors and signaling. Arch. Biochem. Biophys. 2008, 473, 147–160. [Google Scholar] [CrossRef]
  70. Zhao, W.; Byrne, M.H.; Boyce, B.F.; Krane, S.M. Bone resorption induced by parathyroid hormone is strikingly diminished in collagenase-resistant mutant mice. J. Clin. Investig. 1999, 103, 517–524. [Google Scholar] [CrossRef] [Green Version]
  71. Rybchyn, M.S.; Brennan-Speranza, T.C.; Mor, D.; Cheng, Z.; Chang, W.; Conigrave, A.D.; Mason, R.S. The mTORC2 Regulator Homer1 Modulates Protein Levels and Sub-Cellular Localization of the CaSR in Osteoblast-Lineage Cells. Int. J. Mol. Sci. 2021, 22, 6509. [Google Scholar] [CrossRef]
  72. Zheng, C.-M.; Hsu, Y.-H.; Wu, C.-C.; Lu, C.-L.; Liu, W.-C.; Zheng, J.-Q.; Lin, Y.-F.; Chiu, H.-W.; Chang, T.-J.; Shyu, J.-F.; et al. Osteoclast-Released Wnt-10b Underlies Cinacalcet Related Bone Improvement in Chronic Kidney Disease. Int. J. Mol. Sci. 2019, 20, 2800. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  73. Yamauchi, M.; Yamaguchi, T.; Kaji, H.; Sugimoto, T.; Chihara, K. Involvement of calcium-sensing receptor in osteoblastic differentiation of mouse MC3T3-E1 cells. Am. J. Physiol. Endocrinol. Metab. 2005, 288, E608–E616. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  74. Chiu, H.-W.; Hou, Y.-C.; Lu, C.-L.; Lu, K.-C.; Liu, W.-C.; Shyu, J.-F.; Chang, J.-F.; Zheng, C.-M. Cinacalcet improves bone parameters through regulation of osteoclast endoplasmic reticulum stress, autophagy, and apoptotic pathways in chronic kidney disease-mineral and bone disorder. J. Bone Miner. Res. 2021, 37, 215–225. [Google Scholar] [CrossRef] [PubMed]
  75. Komiya, Y.; Habas, R. Wnt signal transduction pathways. Organogenesis 2008, 4, 68–75. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  76. Pai, S.G.; Carneiro, B.A.; Mota, J.M.; Costa, R.; Leite, C.A.; Barroso-Sousa, R.; Kaplan, J.B.; Chae, Y.K.; Giles, F.J. Wnt/beta-catenin pathway: Modulating anticancer immune response. J. Hematol. Oncol. 2017, 10, 101. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  77. Wu, D.; Pan, W. GSK3: A multifaceted kinase in Wnt signaling. Trends Biochem. Sci. 2010, 35, 161–168. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  78. Banse, X.; Sims, T.J.; Bailey, A.J. Mechanical properties of adult vertebral cancellous bone: Correlation with collagen intermolecular cross-links. J. Bone Miner. Res. Off. J. Am. Soc. Bone Miner. Res. 2002, 17, 1621–1628. [Google Scholar] [CrossRef] [PubMed]
  79. Mallipattu, S.K.; Uribarri, J. Advanced glycation end product accumulation: A new enemy to target in chronic kidney disease? Curr. Opin. Nephrol. Hypertens. 2014, 23, 547–554. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  80. Mitome, J.; Yamamoto, H.; Saito, M.; Yokoyama, K.; Marumo, K.; Hosoya, T. Nonenzymatic Cross-Linking Pentosidine Increase in Bone Collagen and Are Associated with Disorders of Bone Mineralization in Dialysis Patients. Calcif. Tissue Int. 2011, 88, 521. [Google Scholar] [CrossRef] [PubMed]
  81. Finch, J.L.; Tokumoto, M.; Nakamura, H.; Yao, W.; Shahnazari, M.; Lane, N.; Slatopolsky, E. Effect of paricalcitol and cinacalcet on serum phosphate, FGF-23, and bone in rats with chronic kidney disease. Am. J. Physiol. Ren. Physiol. 2010, 298, F1315–F1322. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  82. Henley, C.; Davis, J.; Miller, G.; Shatzen, E.; Cattley, R.; Li, X.; Martin, D.; Yao, W.; Lane, N.; Shalhoub, V. The calcimimetic AMG 641 abrogates parathyroid hyperplasia, bone and vascular calcification abnormalities in uremic rats. Eur. J. Pharmacol. 2009, 616, 306–313. [Google Scholar] [CrossRef] [PubMed]
  83. Allard, L.; Demoncheaux, N.; Machuca-Gayet, I.; Georgess, D.; Coury-Lucas, F.; Jurdic, P.; Bacchetta, J. Biphasic Effects of Vitamin D and FGF23 on Human Osteoclast Biology. Calcif. Tissue Int. 2015, 97, 69–79. [Google Scholar] [CrossRef] [PubMed]
  84. Bernardor, J.; Flammier, S.; Ranchin, B.; Gaillard, S.; Platel, D.; Peyruchaud, O.; Machuca-Gayet, I.; Bacchetta, J. Inhibition of Osteoclast Differentiation by 1.25-D and the Calcimimetic KP2326 Reveals 1.25-D Resistance in Advanced CKD. J. Bone Miner. Res. Off. J. Am. Soc. Bone Miner. Res. 2020, 35, 2265–2274. [Google Scholar] [CrossRef] [PubMed]
  85. Smajilovic, S.; Hansen, J.L.; Christoffersen, T.E.; Lewin, E.; Sheikh, S.P.; Terwilliger, E.F.; Brown, E.M.; Haunso, S.; Tfelt-Hansen, J. Extracellular calcium sensing in rat aortic vascular smooth muscle cells. Biochem. Biophys. Res. Commun. 2006, 348, 1215–1223. [Google Scholar] [CrossRef] [PubMed]
  86. Molostvov, G.; James, S.; Fletcher, S.; Bennett, J.; Lehnert, H.; Bland, R.; Zehnder, D. Extracellular calcium-sensing receptor is functionally expressed in human artery. Am. J. Physiol.-Ren. Physiol. 2007, 293, F946–F955. [Google Scholar] [CrossRef] [PubMed]
  87. Alam, M.U.; Kirton, J.P.; Wilkinson, F.L.; Towers, E.; Sinha, S.; Rouhi, M.; Vizard, T.N.; Sage, A.P.; Martin, D.; Ward, D.T.; et al. Calcification is associated with loss of functional calcium-sensing receptor in vascular smooth muscle cells. Cardiovasc. Res. 2009, 81, 260–268. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  88. Qu, Y.Y.; Hui, J.; Wang, L.M.; Tang, N.; Zhong, H.; Liu, Y.M.; Li, Z.; Feng, Q.; He, F. Reduced Expression of the Extracellular Calcium-Sensing Receptor (CaSR) Is Associated with Activation of the Renin-Angiotensin System (RAS) to Promote Vascular Remodeling in the Pathogenesis of Essential Hypertension. PLoS ONE 2016, 11, e0157456. [Google Scholar] [CrossRef] [PubMed]
  89. Zhang, N.; Chow, S.K.H.; Leung, K.S.; Lee, H.H.; Cheung, W.H. An animal model of co-existing sarcopenia and osteoporotic fracture in senescence accelerated mouse prone 8 (SAMP8). Exp. Gerontol. 2017, 97, 1–8. [Google Scholar] [CrossRef]
  90. Lu, M.; Leng, B.; He, X.; Zhang, Z.; Wang, H.; Tang, F. Calcium Sensing Receptor-Related Pathway Contributes to Cardiac Injury and the Mechanism of Astragaloside IV on Cardioprotection. Front. Pharmacol. 2018, 9, 1163. [Google Scholar] [CrossRef]
  91. Ortiz-Capisano, M.C.; Reddy, M.; Mendez, M.; Garvin, J.L.; Beierwaltes, W.H. Juxtaglomerular cell CaSR stimulation decreases renin release via activation of the PLC/IP3 pathway and the ryanodine receptor. Am. J. Physiol. Ren. Physiol. 2012, 304, F248–F256. [Google Scholar] [CrossRef] [Green Version]
  92. Mendoza, F.J.; Martinez-Moreno, J.; Almaden, Y.; Rodriguez-Ortiz, M.E.; Lopez, I.; Estepa, J.C.; Henley, C.; Rodriguez, M.; Aguilera-Tejero, E. Effect of calcium and the calcimimetic AMG 641 on matrix-Gla protein in vascular smooth muscle cells. Calcif. Tissue Int. 2011, 88, 169–178. [Google Scholar] [CrossRef] [PubMed]
  93. Hou, Y.-C.; Lu, C.-L.; Zheng, C.-M.; Chen, R.-M.; Lin, Y.-F.; Liu, W.-C.; Yen, T.-H.; Chen, R.; Lu, K.-C. Emerging Role of Vitamins D and K in Modulating Uremic Vascular Calcification: The Aspect of Passive Calcification. Nutrients 2019, 11, 152. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  94. Ziegelstein, R.C.; Xiong, Y.; He, C.; Hu, Q. Expression of a functional extracellular calcium-sensing receptor in human aortic endothelial cells. Biochem. Biophys. Res. Commun. 2006, 342, 153–163. [Google Scholar] [CrossRef] [PubMed]
  95. Hénaut, L.; Boudot, C.; Massy, Z.A.; Lopez-Fernandez, I.; Dupont, S.; Mary, A.; Drüeke, T.B.; Kamel, S.; Brazier, M.; Mentaverri, R. Calcimimetics increase CaSR expression and reduce mineralization in vascular smooth muscle cells: Mechanisms of action. Cardiovasc. Res. 2014, 101, 256–265. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  96. Hong, S.; Zhang, X.; Zhang, X.; Liu, W.; Fu, Y.; Liu, Y.; Shi, Z.; Chi, J.; Zhao, M.; Yin, X. Role of the calcium sensing receptor in cardiomyocyte apoptosis via mitochondrial dynamics in compensatory hypertrophied myocardium of spontaneously hypertensive rat. Biochem. Biophys. Res. Commun. 2017, 487, 728–733. [Google Scholar] [CrossRef] [PubMed]
  97. Zhang, T.; Tang, N.; Xi, D.; Zhao, Y.; Liu, Y.; Wang, L.; Tang, Y.; Zhang, X.; Zhong, H.; He, F. Calcimimetic R568 improved cardiac remodeling by classic and novel renin-angiotensin system in spontaneously hypertensive rats. Exp. Biol. Med. 2019, 244, 789–801. [Google Scholar] [CrossRef] [PubMed]
  98. Liu, L.; Wang, C.; Lin, Y.; Xi, Y.; Li, H.; Shi, S.; Li, H.; Zhang, W.; Zhao, Y.; Tian, Y.; et al. Suppression of calcium-sensing receptor ameliorates cardiac hypertrophy through inhibition of autophagy. Mol. Med. Rep. 2016, 14, 111–120. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  99. Capasso, G.; Geibel, P.J.; Damiano, S.; Jaeger, P.; Richards, W.G.; Geibel, J.P. The calcium sensing receptor modulates fluid reabsorption and acid secretion in the proximal tubule. Kidney Int. 2013, 84, 277–284. [Google Scholar] [CrossRef] [Green Version]
  100. Loupy, A.; Ramakrishnan, S.K.; Wootla, B.; Chambrey, R.; de la Faille, R.; Bourgeois, S.; Bruneval, P.; Mandet, C.; Christensen, E.I.; Faure, H.; et al. PTH-independent regulation of blood calcium concentration by the calcium-sensing receptor. J. Clin. Investig. 2012, 122, 3355–3367. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  101. Rothe, H.; Shapiro, W.; Sun, W.Y.; Matalon, A. CaSR polymorphism Arg990Gly and response to calcimimetic agents in end-stage kidney disease patients with secondary hyperparathyroidism and in cell culture. Pers. Med. 2008, 5, 109–116. [Google Scholar] [CrossRef] [PubMed]
  102. Okechukwu, C.; Awumey, E.M. Calcium-sensing receptor (CaSR) expression and Renin/Angiotensin System (RAS) in salt-sensitive rats. FASEB J. 2019, 33, 577.6. [Google Scholar] [CrossRef]
  103. Watanabe, S.; Fukumoto, S.; Chang, H.; Takeuchi, Y.; Hasegawa, Y.; Okazaki, R.; Chikatsu, N.; Fujita, T. Association between activating mutations of calcium-sensing receptor and Bartter’s syndrome. Lancet 2002, 360, 692–694. [Google Scholar] [CrossRef]
  104. Fukagawa, M.; Yumita, S.; Akizawa, T.; Uchida, E.; Tsukamoto, Y.; Iwasaki, M.; Koshikawa, S. Cinacalcet (KRN1493) effectively decreases the serum intact PTH level with favorable control of the serum phosphorus and calcium levels in Japanese dialysis patients. Nephrol. Dial. Transplant. 2008, 23, 328–335. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  105. Peacock, M.; Bilezikian, J.P.; Klassen, P.S.; Guo, M.D.; Turner, S.A.; Shoback, D. Cinacalcet Hydrochloride Maintains Long-Term Normocalcemia in Patients with Primary Hyperparathyroidism. J. Clin. Endocrinol. Metab. 2005, 90, 135–141. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  106. Messa, P.; Macário, F.; Yaqoob, M.; Bouman, K.; Braun, J.; von Albertini, B.; Brink, H.; Maduell, F.; Graf, H.; Frazão, J.M.; et al. The OPTIMA study: Assessing a new cinacalcet (Sensipar/Mimpara) treatment algorithm for secondary hyperparathyroidism. Clin. J. Am. Soc. Nephrol. 2008, 3, 36–45. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  107. Fishbane, S.; Shapiro, W.B.; Corry, D.B.; Vicks, S.L.; Roppolo, M.; Rappaport, K.; Ling, X.; Goodman, W.G.; Turner, S.; Charytan, C. Cinacalcet HCl and concurrent low-dose vitamin D improves treatment of secondary hyperparathyroidism in dialysis patients compared with vitamin D alone: The ACHIEVE study results. Clin. J. Am. Soc. Nephrol. 2008, 3, 1718–1725. [Google Scholar] [CrossRef] [Green Version]
  108. Itano, Y.; Kato, S.; Tsuboi, M.; Kasuga, H.; Tsuruta, Y.; Sato, F.; Hishida, M.; Ishimoto, T.; Kosugi, T.; Ando, M.; et al. A Prospective, Randomized Clinical Trial of Etelcalcetide in Patients Receiving Hemodialysis With Secondary Hyperparathyroidism (the DUET Trial). Kidney Int. Rep. 2020, 5, 2168–2177. [Google Scholar] [CrossRef] [PubMed]
  109. Shigematsu, T.; Akizawa, T.; Uchida, E.; Tsukamoto, Y.; Iwasaki, M.; Koshikawa, S. Long-Term Cinacalcet HCl Treatment Improved Bone Metabolism in Japanese Hemodialysis Patients with Secondary Hyperparathyroidism. Am. J. Nephrol. 2009, 29, 230–236. [Google Scholar] [CrossRef] [Green Version]
  110. Tsuruta, Y.; Okano, K.; Kikuchi, K.; Tsuruta, Y.; Akiba, T.; Nitta, K. Effects of cinacalcet on bone mineral density and bone markers in hemodialysis patients with secondary hyperparathyroidism. Clin. Exp. Nephrol. 2013, 17, 120–126. [Google Scholar] [CrossRef]
  111. Chertow, G.M.; Block, G.A.; Correa-Rotter, R.; Drüeke, T.B.; Floege, J.; Goodman, W.G.; Herzog, C.A.; Kubo, Y.; London, G.M.; Mahaffey, K.W.; et al. Effect of Cinacalcet on Cardiovascular Disease in Patients Undergoing Dialysis. N. Engl. J. Med. 2012, 367, 2482–2494. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  112. Moe, S.M.; Wetherill, L.; Decker, B.S.; Lai, D.; Abdalla, S.; Long, J.; Vatta, M.; Foroud, T.M.; Chertow, G.M. Calcium-Sensing Receptor Genotype and Response to Cinacalcet in Patients Undergoing Hemodialysis. Clin. J. Am. Soc. Nephrol. CJASN 2017, 12, 1128–1138. [Google Scholar] [CrossRef]
  113. Grzegorzewska, A.E.; Frycz, B.A.; Świderska, M.; Niepolski, L.; Mostowska, A.; Jagodziński, P.P. Calcium-sensing receptor gene (CASR) polymorphisms and CASR transcript level concerning dyslipidemia in hemodialysis patients: A cross-sectional study. BMC Nephrol. 2019, 20, 436. [Google Scholar] [CrossRef] [PubMed]
  114. Ruderman, I.; Smith, E.R.; Toussaint, N.D.; Hewitson, T.D.; Holt, S.G. Longitudinal changes in bone and mineral metabolism after cessation of cinacalcet in dialysis patients with secondary hyperparathyroidism. BMC Nephrol. 2018, 19, 113. [Google Scholar] [CrossRef] [Green Version]
  115. Lau, W.L.; Obi, Y.; Kalantar-Zadeh, K. Parathyroidectomy in the Management of Secondary Hyperparathyroidism. Clin. J. Am. Soc. Nephrol. 2018, 13, 952–961. [Google Scholar] [CrossRef]
  116. Keutgen, X.M.; Buitrago, D.; Filicori, F.; Kundel, A.; Elemento, O.; Fahey, T.J., 3rd; Zarnegar, R. Calcimimetics versus parathyroidectomy for treatment of primary hyperparathyroidism: Retrospective chart analysis of a prospective database. Ann. Surg. 2012, 255, 981–985. [Google Scholar] [CrossRef]
  117. Cruzado, J.M.; Moreno, P.; Torregrosa, J.V.; Taco, O.; Mast, R.; Gómez-Vaquero, C.; Polo, C.; Revuelta, I.; Francos, J.; Torras, J.; et al. A Randomized Study Comparing Parathyroidectomy with Cinacalcet for Treating Hypercalcemia in Kidney Allograft Recipients with Hyperparathyroidism. J. Am. Soc. Nephrol. 2016, 27, 2487–2494. [Google Scholar] [CrossRef]
  118. Cañadillas, S.; Canalejo, A.; Santamaría, R.; Rodríguez, M.E.; Estepa, J.C.; Martín-Malo, A.; Bravo, J.; Ramos, B.; Aguilera-Tejero, E.; Rodríguez, M.; et al. Calcium-Sensing Receptor Expression and Parathyroid Hormone Secretion in Hyperplastic Parathyroid Glands from Humans. J. Am. Soc. Nephrol. 2005, 16, 2190. [Google Scholar] [CrossRef] [Green Version]
  119. Fukagawa, M.; Yokoyama, K.; Koiwa, F.; Taniguchi, M.; Shoji, T.; Kazama, J.J.; Komaba, H.; Ando, R.; Kakuta, T.; Fujii, H.; et al. Clinical practice guideline for the management of chronic kidney disease-mineral and bone disorder. Ther. Apher. Dial. 2013, 17, 247–288. [Google Scholar] [CrossRef]
Figure 2. Molecular mechanism of CaMs on osteoclasts in CKD. Osteoblast and osteoclast proliferation, differentiation, and apoptosis are influenced by local extracellular calcium concentrations. The activation of the RANK/RANKL signal from osteoblasts modulates osteoclastogenesis. The signal induces the elevation of endoplasmic reticulum stress (ER stress), which activates the cAMP response element-binding protein (CREB) phosphorylation. The downstream transcription of the nuclear factor of activated T cells, cytoplasmic (NFATc1) activates osteoclastogenesis. On the other hand, ER stress also activates the autophagy process to maintain osteoclast viability. RANK/RANKL also inhibits osteoclast apoptosis. CaMs induce the release of Wnt-10b from osteoclasts and inhibit the autophagy vial, activating the Akt/mTOR pathway. In this manner, CaMs increase bone formation, along with the apoptosis of the osteoclast.
Figure 2. Molecular mechanism of CaMs on osteoclasts in CKD. Osteoblast and osteoclast proliferation, differentiation, and apoptosis are influenced by local extracellular calcium concentrations. The activation of the RANK/RANKL signal from osteoblasts modulates osteoclastogenesis. The signal induces the elevation of endoplasmic reticulum stress (ER stress), which activates the cAMP response element-binding protein (CREB) phosphorylation. The downstream transcription of the nuclear factor of activated T cells, cytoplasmic (NFATc1) activates osteoclastogenesis. On the other hand, ER stress also activates the autophagy process to maintain osteoclast viability. RANK/RANKL also inhibits osteoclast apoptosis. CaMs induce the release of Wnt-10b from osteoclasts and inhibit the autophagy vial, activating the Akt/mTOR pathway. In this manner, CaMs increase bone formation, along with the apoptosis of the osteoclast.
Pharmaceuticals 15 00952 g002
Figure 3. The molecular mechanism underlying the effects of CaMs on osteoblasts in CKD. The Wnt-10b released from osteoclasts by CaMs activates β-catenin by inhibiting GSK-3β-mediated degradation. On the other hand, nuclear Erk1/2 phosphorylation activated by CaMs activates the transcription, along with c-Jun N-terminal kinase (JNK)/PI3K, which increases osteoblast proliferation and maintains survival. The activation of Erk1/2 with serum response factor-response element (SRF-RE)-luc and insulin-like growth hormone factor is responsible for osteoblast proliferation/survival and differentiation and its maturation.
Figure 3. The molecular mechanism underlying the effects of CaMs on osteoblasts in CKD. The Wnt-10b released from osteoclasts by CaMs activates β-catenin by inhibiting GSK-3β-mediated degradation. On the other hand, nuclear Erk1/2 phosphorylation activated by CaMs activates the transcription, along with c-Jun N-terminal kinase (JNK)/PI3K, which increases osteoblast proliferation and maintains survival. The activation of Erk1/2 with serum response factor-response element (SRF-RE)-luc and insulin-like growth hormone factor is responsible for osteoblast proliferation/survival and differentiation and its maturation.
Pharmaceuticals 15 00952 g003
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Hou, Y.-C.; Zheng, C.-M.; Chiu, H.-W.; Liu, W.-C.; Lu, K.-C.; Lu, C.-L. Role of Calcimimetics in Treating Bone and Mineral Disorders Related to Chronic Kidney Disease. Pharmaceuticals 2022, 15, 952. https://doi.org/10.3390/ph15080952

AMA Style

Hou Y-C, Zheng C-M, Chiu H-W, Liu W-C, Lu K-C, Lu C-L. Role of Calcimimetics in Treating Bone and Mineral Disorders Related to Chronic Kidney Disease. Pharmaceuticals. 2022; 15(8):952. https://doi.org/10.3390/ph15080952

Chicago/Turabian Style

Hou, Yi-Chou, Cai-Mei Zheng, Hui-Wen Chiu, Wen-Chih Liu, Kuo-Cheng Lu, and Chien-Lin Lu. 2022. "Role of Calcimimetics in Treating Bone and Mineral Disorders Related to Chronic Kidney Disease" Pharmaceuticals 15, no. 8: 952. https://doi.org/10.3390/ph15080952

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