Sie sind auf Seite 1von 10

Special Issue: Metabolism Through the Lens of GPCRs

Review
Activating Calcium-Sensing
Receptor Mutations:
Prospects for Future
Treatment with Calcilytics
Bernhard Mayr,1,* Markus Glaudo,1 and Christof Schöfl1

Activating mutations of the G protein-coupled receptor, calcium-sensing recep-


Trends
tor (CaSR), cause autosomal dominant hypocalcemia and Bartter syndrome
The CaSR determines the set-point of
type 5. These mutations lower the set-point for extracellular calcium sensing, serum calcium.
thereby causing decreased parathyroid hormone secretion and disturbed renal
Activating CaSR mutations cause the
calcium handling with hypercalciuria. Available therapies increase serum cal- hypocalcemic disorders autosomal
cium levels but raise the risk of complications in affected patients. Symptom dominant hypocalcemia (ADH) and
relief and the prevention of adverse outcome is currently very difficult to Bartter syndrome type 5 (BS5) which
are difficult to treat.
achieve. Calcilytics act as CaSR antagonists that attenuate its activity, thereby
correcting the molecular defect of activating CaSR proteins in vitro and elevat- Calcilytics are allosteric modulators of
ing serum calcium in mice and humans in vivo, and have emerged as the most the CaSR and can correct the malfunc-
tion of most activating CaSR proteins
promising therapeutics for the treatment of these rare and difficult to treat in vitro.
diseases.
Calcilytics reduce the disease pheno-
type in animal studies and humans
Introduction in vivo.
Stable serum calcium levels are maintained by a homeostatic feedback loop. The CaSR
measures the main regulatory target, calcium, and determines its set-point [1,2]. Mutations Calcilytics appear to be the most pro-
mising future therapy for ADH or BS5.
that alter the function of CaSR shift this set-point and lead to hypo- and hypercalcemic
disorders. For the hypercalcemic disorders there are several established therapeutic alternatives
that in most patients yield clinically satisfactory outcomes. By contrast the hypocalcemic
disorders autosomal dominant hypocalcemia (ADH) and Bartter syndrome type 5 (BS5) caused
by activating CaSR mutations are clinically difficult to treat [1]. Most patients remain symptomatic
and develop serious sequelae despite intensive therapy with all available strategies. None of the
current therapeutic strategies corrects the underlying pathophysiology.

The function of CaSR can be modified by allosteric modulators. A positive allosteric CaSR
modulator, the calcimimetic cinacalcet, is already used in clinical practice [1,3]. Negative
allosteric modulators called calcilytics directly inhibit CaSR and have been developed to 1
Division of Endocrinology and
stimulate endogenous parathyroid hormone (PTH) secretion as an alternative to PTH injections Diabetes, Department of Medicine I,
Universitätsklinikum Erlangen,
to promote bone formation in osteoporosis. These drugs may be helpful in reducing excessive Friedrich-Alexander University
CaSR activity in ADH and BS5 patients [1,3]. Erlangen-Nuremberg, Germany

We review here the latest developments in the field of calcilytics as therapeutic options for
*Correspondence:
patients harboring mutations that activate CaSR function, in the light of the current molecular, bernhard.mayr@uk-erlangen.de
physiological, and clinical knowledge about CaSR. We focus on the effects of calcilytics on (B. Mayr).

Trends in Endocrinology & Metabolism, September 2016, Vol. 27, No. 9 http://dx.doi.org/10.1016/j.tem.2016.05.005 643
© 2016 Elsevier Ltd. All rights reserved.
calcium homeostasis in ADH and BS5 patients, while possible effects on renal loss of sodium
and other salts that are only found in BS5 patients are beyond the scope of this review.

Calcium Homeostasis and the CaSR


CaSR Structure and Function
CaSR is a G protein-coupled receptor (GPCR) that couples to heterotrimeric G proteins. CaSR
has an unusually large extracellular domain that has significant homology to bacterial amino acid-
binding proteins that form a so-called venus flytrap domain (VFTD) [4]. CaSR therefore belongs
to GPCR class 3 or class C. The VFTD domain of CaSR binds calcium and positive allosteric
modulators, such as L-amino acids [5], and is linked to a classic seven-transmembrane domain
via a cysteine-rich domain whose precise role in signal transduction is not well understood. The
seven transmembrane-spanning domain of CaSR binds positive and negative allosteric mod-
ulators, and presumably undergoes conformational changes of the membrane-spanning helices
upon ligand binding [6]. This promotes activation of the small G proteins Gq/11, G12/13, and Gi
that regulate the cytosolic calcium, MAP kinase, and cAMP signaling pathways [7,8]. Finally,
CaSR has a large C-terminal intracellular domain. This domain is involved in the regulation of cell
surface expression, receptor trafficking to and from the membrane, and rate of receptor
desensitization [9–11], processes that involve PKC-dependent phosphorylation of T888, b-
arrestin, the adaptor protein 2 (AP2), and other membrane-associated proteins [12–15].
Remarkably, large parts of this domain can be deleted without substantial negative effects
on cell-surface expression and function [16].

Physiology of Calcium Homeostasis


CaSR serves as the sensor for extracellular calcium in the homeostatic serum calcium feedback
loop. Activation of CaSR suppresses PTH release (Figure 1), but the precise mechanism of this
regulation and the involvement of [Ca2+]i, ERK1/2, or other signaling mechanisms is not completely
understood. Lowered serum calcium levels thus lead to increased PTH secretion; PTH in turn
stimulates calcium reabsorption in the kidney and thereby inhibits calcium disposal from the body.
PTH also regulates skeletal calcium release and uptake (Figure 1). Both in the kidney and in bone
the effects of PTH are mediated via the type-1 PTH receptor (PTH1R) that couples to the small
G protein Gs and activates cAMP signaling [17]. PTH also stimulates the 1/-hydroxylation of
vitamin D compounds in the kidney, and thereby indirectly stimulates calcium uptake in the gut
(Figure 1). Together these effects of PTH and vitamin D in kidney, bone, and gut elevate serum
calcium levels and complete the classic regulatory feedback loop of the calcium homeostasis.

CaSR is, however, not only expressed in the parathyroid, but also present in kidney, bone, gut,
and many other tissues [18]. In most of these tissues CaSR seems to be mainly involved in tissue
development, cell growth, and differentiation [19]. In the kidney, however, CaSR is expressed in
various cell types along the nephron and is directly involved in calcium homeostasis [20–22].
Animal studies indicate that activation of CaSR appears to directly enhance urinary calcium
excretion via reduced tubular calcium reabsorption independently of PTH [23,24] (Figure 1).

The effects and the signaling of CaSR and PTH in the kidney also interact with each other. PTH
activates tubular calcium reabsorption via PTH1R, the small G protein Gs and cAMP signaling.
By contrast, CaSR can activate Gi, suppress cAMP signaling, and inhibit PTH-induced trans-
cellular divalent cation reabsorption in the thick ascending limb of the loop of Henle (TAL) [21,25–
27]. In addition, PTH-stimulated synthesis of active vitamin D compounds can enhance CaSR
expression in the kidney [28].

When serum calcium levels are decreased, CaSR and PTH act synergistically on the kidney to
raise serum calcium as follows: PTH release from the parathyroid is enhanced and PTH1R
activation reduces urinary calcium excretion. Reduced activation of CaSR in the kidney eases

644 Trends in Endocrinology & Metabolism, September 2016, Vol. 27, No. 9
Absorpon

1,25 Vit D

Secreon Resorpon
Serum
PTH
calcium

CaSR

Excreon

Figure 1. Calcium Homeostasis and Calcium-Sensing Receptor (CaSR) Signaling. The serum calcium set-point
is determined by the CaSR, which negatively regulates parathyroid hormone (PTH) secretion from the parathyroid gland.
PTH in turn regulates calcium handling in bone and kidney, and enhances intestinal calcium absorption via increased renal
formation of 1,25-dihydroxyvitamin D (1,25 Vit D). In addition, CaSR inhibits PTH effects on the kidney and directly
stimulates urinary calcium excretion. The thick grey arrow indicates PTH secretion from the parathyroid gland. Thick blue
arrows indicate calcium absorption in the intestine, resorption from bone, and excretion in the kidney. The thin grey arrows
indicate regulatory actions of PTH. The thin red arrows indicate regulatory actions of CaSR. Plus and minus signs indicate
positive and negative effects.

suppression of PTH-activated calcium reabsorption and directly enhances tubular reabsorption


of calcium [21,23–27]. In this complex interaction (Figure 1) the effects of CaSR and PTH,
however, can also be conflicting. For example, excessive PTH levels in primary hyperparathy-
roidism (pHPT) would be expected to lead to increased tubular reabsorption so as to reduce
renal calcium excretion. However, a frequent clinical characteristic of pHPT is hypercalciuria, and
this contrasts with the hypocalciuria seen in familial hypocalciuric hypercalcemia (FHH) [2,29,30]
despite a presumably similarly elevated filtered calcium load in the kidney. This indicates that
CaSR predominates over PTH because wild-type CaSR correctly senses the elevated calcium
levels in pHPT and promotes renal calcium excretion.

Mutations Causing Gain of CaSR Function


Clinical Features of Patients with Activating CaSR Mutations
Activating CASR mutations lower the set-point for serum calcium and lead to hypocalcemic
diseases called autosomal dominant hypocalcemia (ADH) and Bartter syndrome type 5 (BS 5).
ADH patients have mild to moderate hypocalcemia with low to normal PTH levels and increased
serum phosphate levels. Inappropriately normal or high urinary calcium excretion is present in
almost all ADH patients [31,32], and is a consequence of the specific pathophysiologic distur-
bance of the calcium homeostasis by activating mutations (see below). In Bartter syndrome type
5, incompletely understood molecular mechanisms lead to additional renal wasting of sodium
and other salts [33–35]. The main clinical problems in all these patients are symptomatic

Trends in Endocrinology & Metabolism, September 2016, Vol. 27, No. 9 645
hypocalcemia and an elevated risk of tissue calcifications in the kidney, the basal ganglia, and
eye lenses, which may cause renal failure, neurologic deficits, and cataract formation [32,36–41].
Long-term complications develop both in symptomatic patients who receive treatment as well
as in asymptomatic untreated patients [32,39]. As discussed below, the main therapeutic
challenge with all available therapies is that symptom relief further aggravates the already
elevated risk of complications.

Molecular Mechanisms of Activating CaSR Mutations


Mutations that activate CaSR are spread throughout CaSR protein, with no apparent hotspot in
any particular functional domain.i Hence there appears to be no single unifying mechanism that
leads to an activated CaSR function.

Point mutations in the extracellular domain could alter binding of the ligand calcium but, because
the 3D structure of CaSR has not yet been determined, predictions of functional alterations from
structural changes are difficult and may be misleading [42,43]. Three amino acids (L173, P221,
E297) in the hinge region between loops 2 and 3 of the VFTD appear to be ‘toggle’ residues
where both activating and inactivating mutations have been described [2,42,44–48]. It is
believed that agonist binding affects oscillation of the VFTD between ‘open’ and ‘closed’
conformations [4], and these amino acids may be crucial for these conformational changes
and receptor activation. Remarkably, CaSR can be activated by large deletions in the intracellular
domain involving amino acid 895 and beyond [42,49,50], presumably by changes in cell-surface
expression, receptor trafficking, or desensitization. An almost identical phenotype (ADH type 2)
is caused by the activating point mutations of the small G protein /11 subunit (GNA11) [51–54].

Although CaSR activates multiple signaling pathways, very little is known about the differential
activation of these pathways. Some CaSR proteins exhibit biased but not grossly different
signaling via [Ca2+]i mobilization relative to ERK1/2 phosphorylation [55,56], but the pathophysi-
ological significance of this finding is not clear.

Pathophysiology of Activating CaSR Mutations


Gain-of-function mutations cause CaSR activation at lower than normal extracellular calcium
levels not only in the parathyroid but also in the kidney [21,31,32]. In the parathyroid glands, this
results in reduced PTH secretion and inappropriately low or even undetectable serum PTH levels
[32]. This expectedly leads to hypocalcemia, but the predominance of CaSR-mediated effects in
the interaction of PTH and CaSR in the kidney (Figure 1) creates specific problems in ADH that
are not present in patients with hypoparathyroidism caused by postsurgical or autoimmune
damage to the parathyroid glands.

If hypoparathyroidism is caused by loss of parathyroid tissue, the lack of PTH action in the kidney
would be expected to result in increased calcium excretion, but there is usually hypocalciuria (i.
e., reduced urinary calcium excretion). This supports results from animal studies that the intact
CaSR correctly senses lowered serum calcium levels independently of PTH, and that the direct
effects of CaSR on renal tubular reabsorption of calcium predominate over the effects of PTH
[21,23,24].

Patients with activating mutations therefore have absent or reduced PTH owing to CaSR-
mediated suppression of PTH secretion. In addition, the inappropriately activated CaSR further
suppresses PTH effects in the kidney and directly prevents the kidney from conserving calcium
[21,23–27] (Figure 1). This results in inappropriate normo- or even hypercalciuria, and elevates
the risk of kidney stones or calcifications in affected patients even without any therapeutic
intervention [32,39]. This also counteracts all conventional therapeutic attempts to raise serum
calcium levels because in ADH the kidney responds with increased calcium excretion to an

646 Trends in Endocrinology & Metabolism, September 2016, Vol. 27, No. 9
increased filtered calcium load produced by any rise of serum calcium levels. This most likely
further elevates the risk of renal complications [32,39,57,58].

Nephrocalcinosis and nephrolithiasis are frequently found in patients with ADH [39] and hypo-
parathyroidism from other causes [59,60]. An increased risk of calcifications of the basal ganglia
and cataract formation is documented only in ADH and non-surgical hypoparathyroidism
[39,60]. Local effects of an activated CaSR itself may contribute to tissue calcifications in
ADH [40,61] beyond the effects of general hypocalcemia and lack of PTH by incompletely
understood mechanisms.

Currently Available Therapies for Patients with Activating CaSR Mutations


Vitamin D analogs and calcium supplementation is the standard treatment to raise serum
calcium levels in ADH patients [32,39,40,57,58]. However, this inevitably aggravates hyper-
calciuria [32,39,57,58], which frequently limits attempts to normalize serum calcium. As in
hypoparathyroidism from other causes, there are no studies relating target calcium levels with
clinically relevant endpoints [60]. It is not known if symptom relief is sufficient or whether targeting
higher serum calcium levels in the low or mid-normal range has long-term clinical benefits.

In some patients, hydrochlorothiazide was found to temporarily reduce hypercalciuria [62].


Replacement of PTH1–34 (teriparatide) has been reported in five patients with ADH [40,62–65]
and was generally effective in reducing symptoms and/or raising serum calcium. Although urinary
calcium excretion declined or was constant despite rising serum calcium levels in most patients,
nephrocalcinosis can develop despite continuous PTH therapy [65], and there may be considerable
variability between individual patients or mutated CaSR proteins in the response to PTH1–34 [64].

At best, the available therapeutic options reduce symptoms without further increasing the
elevated risk of tissue complications. Complete symptom relief and the reduction of long-term
complications caused by the underlying disease itself are currently very difficult to achieve. The
reason for this therapeutic dilemma is the predominance of the effects of CaSR over those of
PTH on tubular reabsorption of calcium in the kidney. The activated CaSR in ADH and BS5
reduces the desired effects, and increases the adverse effects, of conventional therapeutic
attempts [32,39,57,58]

The defect in CaSR itself therefore needs to be targeted to correct the underlying pathophysi-
ology. An increasing body of evidence from in vitro and in vivo studies suggests that this could
become possible with calcilytics, drugs that directly inhibit CaSR [3].

Calcilytics as a Therapeutic Perspective for ADH and BS5 Patients


Effect of Calcilytics In Vitro on Mutant Proteins Causing Gain of CaSR Function
Most drugs have been designed to target normally functioning wild-type receptors or enzymes.
To pharmacologically correct the molecular cause of ADH and BS5, calcilytics will need to rectify
the function of altered CaSR or G protein /11 (GNA11), and the efficacy of calcilytics under
these circumstances needs to be established. As recently reviewed, about 70 naturally-occur-
ring activating CaSR and GNA11 mutations that cause ADH or BS5 have been described [1].
Functional in vitro tests with calcilytics have been reported for 32 activating CaSR and two
activating GNA11 point mutants, and all were sensitive to at least one calcilytic in vitro [1,66,67].
Four of these mutants (A840 V, Q245R, E228K, and E228A) were also sensitive in vivo [67] (see
below). It therefore appears highly likely that, in principle, the majority of point-mutants in ADH
and BS5 patients will be responsive to calcilytics. Calcilytics may also work on activating
deletion-mutant proteins because these usually retain the amino acids between 600 and
850 that appear to be crucial for the binding of calcilytics [6,68], but this has not yet been
studied (see Outstanding Questions).

Trends in Endocrinology & Metabolism, September 2016, Vol. 27, No. 9 647
Effect of Calcilytics in Animal Models of ADH
The in vivo effects of calcilytics have been studied in CaSR C129S and A843E knock-in mouse
models [61,69] and the Nuf mouse model (CaSR L723Q) generated by random mutagenesis
[69,70]. All three mouse models replicate the human ADH phenotype, with decreased serum Ca
and PTH, and increased serum phosphate levels. The C129S and A843E knock-in mouse
models also showed absolute or relative hypercalciuria and reduced urinary cAMP excretion [61].
A843E causes BS5 in humans; however, renal salt-wasting was not studied in this mouse model,
but the A843E mice had a more severe phenotype and higher mortality than C129S mice [61].

Administration of the antagonists JTT-305/MK-5442 [61] or NPS 2143 [69] increased serum
calcium and PTH but stabilized or even decreased urinary calcium excretion in all three models.
In C129S and A843E mice, JTT-305/MK-5442 also increased urinary cAMP excretion, and
administration for 3 months reduced calcium excretion and prevented renal calcification as
determined by histological von Kossa staining [61]. By contrast, long-term subcutaneous PTH
injections also increased serum calcium but failed to reduce calcium excretion and did not
prevent renal calcification in the C129S and A843E models [61].

Effect of Calcilytics in Humans


Calcilytics with a particularly short serum half-life have been developed to stimulate endogenous
PTH secretion as an oral alternative to subcutaneous PTH injections to enhance bone formation
in osteoporosis. There are numerous registered Phase I and Phase II clinical trialsii in which a total
of more than 1000 individuals, mostly postmenopausal osteopenic or osteoporotic women, but
also almost 100 healthy males, were treated with calcilytics to study their clinical effects (Table 1).
Because serum calcium levels were normal at baseline in these subjects, all these individuals
presumably had wild-type CaSR. In six of these registered and three unregistered trials changes
in serum calcium, phosphate, and PTH levels were determined, as well as renal calcium
excretion. This allows conclusion to be drawn about the in vivo effects of calcilytics in humans
with wild-type CaSR and, by extension, about possible therapeutic effects in ADH patients with
mutated CaSR proteins sensitive to calcilytics. In addition, results from a small Phase II trial with
the calcilytic NPSP795 for the treatment of five ADH patients harboring the A840 V, Q245R,
E228K, and E228A mutations have been reported recently [67]. To our knowledge no clinical
trials have been performed in patients with BS5 (see Outstanding Questions).

In three of these studies the amino-alcohol calcilytic ronacaleret (also known as SB-751689) was
studied [71,72]; in two studies the closely related compound NPSP795 (also known as SB-
423562) [67,73] was used, and one study employed SB-423557, a prodrug of NPSP795 [73].
Three studies used the amino-alcohol MK-5442 [74] (also known as JTT-305 [75]) and one study
utilized the quinazolinone AXT914 [76] (Table 1). All drugs were orally administered once daily,
except for NPSP795 which was given by short intravenous infusion.

No trial in osteopenic women was able to demonstrate a significantly increase in bone mineral
density. However, eight of ten studies demonstrated a dose-dependent increase in serum
calcium levels (Table 1); in ADH1 patients serum calcium levels were maintained despite fasting
and no calcium or vitamin D supplementation [67]. One study even reported that 44% of study
participants in the highest dose group became hypercalcemic [74]. Beyond this, the most
common adverse effects reported were mild neurological and gastrointestinal symptoms such
as fatigue, headache, constipation, diarrhea, nausea, and dyspepsia. Overall, however, calci-
lytics were well tolerated [67,71–74,76–80].

Nine trials studied the effects of calcilytics on PTH levels and all reported dose-dependent
increases in PTH. The effect on PTH was relatively short-lived, and after a rapid increase PTH
levels returned to baseline 6–12 h after administration [71,73,74,80]. Serum calcium showed a

648 Trends in Endocrinology & Metabolism, September 2016, Vol. 27, No. 9
Table 1. ADH1 patients, Healthy Subjects, and Osteopenic Postmenopausal Women, treated with Calcilyticsa,b
Effect of Calcilytic Treatment Compared to Baseline and/or Control Group

Clinicaltrials.gov Study Subjects Receiving Calcilytic/ Dose S-Ca S-PTH S-PO4 U-Ca U-cAMP Refs
Type drug Excretion

NCT02204579 Phase II 5 ADH1 patients NPSP795 5–50 mg i.v. N.c.c " N.r. # [67]
(3 male, 2 female)

NCT00532077 Phase I 34 Healthy Ronacaleret 100 or "–"" " # Acute " [71,77]
postmenopausal 400 mg/ administration: ##
women day p.o. Chronic
administration: ""

NCT00471237 Phase II 264/238 Osteopenic Ronacaleret 100–400 mg/ "–"" " N.r. " N.r. [72,78]
postmenopausal day p.o.
women

NCT00548496 58 Male and female Ronacaleret 200 or " " N.r. N.r. N.r. [79]
subjects with a radial 400 mg/
fracture day p.o.

NCT00996801 Phase II 351 Postmenopausal MK-5442 5–15 mg/ "–"" " N.r. N.r. N.r. [74]
women with day p.o.
osteoporosis

N.r. 100 MK-5442 10 or " N.r. N.r. N.r. N.r. [88]


Postmenopausal 20 mg p.o.
women

NCT00960934 Phase II 319 Postmenopausal MK-5442 2.5–15 mg/ " " N.r. N.r. N.r. [80]
women with day p.o.
osteoporosis

NCT00417261 Phase I 30 Healthy volunteers AXT914 4–120 mg/ " " # N.r. N.r. [76]
and (12 male) and day p.o.
Phase II osteopenic
postmenopausal
women

N.r. Phase I 28 Healthy men NPSP795 20 mg to N.c. " N.r. N.c. N.r. [73]
5 mg i.v.

N.r. Phase I 54 Healthy men SB-423557 5–500 mg p.o. " " N.r. N.c. N.r. [73]

a
Effects of NPSP795, ronacaleret, MK-5442, AXT914, and SB-423557 on serum calcium and PTH levels, and on urinary calcium excretion in clinical studies with ADH1
patients and in subjects with normal CaSR.
b
Abbreviations: ", increase; "", strong increase; #, decrease; ##, strong decrease; i.v., administered by short intravenous infusion; N.c., no change; N.r., not reported;
p.o., orally administered; S-Ca, total serum calcium; S-PTH, serum parathyroid hormone; U-Ca excretion, urinary calcium excretion.
c
No change in serum calcium despite fasting and no calcium or vitamin D supplementation.

similar pattern at the beginning of calcilytic treatment, with a return to pre-treatment levels after
12 h. These results are in line with the rapid pharmacokinetic profiles of ronacaleret [71] and
AXT914 [76] (tmax 1–2 h, t½ 4–5 h) and NPSP795 (t½ <1 h after i.v. administration; tmax 2–3 h, t½
1.4–3.7 h after oral administration of the prodrug SB-423557) [73]. After once-daily calcilytic
treatment continously for 2 weeks or longer, serum calcium levels still increased acutely and
decreased later, but remained above pre-treatment levels, and overall serum calcium levels
gradually increased. The PTH response pattern, however, was remarkably stable until the end of
the observation periods, indicating no loss of effect for MK-5442 [74,80] or ronacaleret [72] for
up to 6 or 12 months of treatment, respectively. The fact that four different calcilytics (SB-423557
is rapidly metabolized to NPSP795 in vivo) from two chemical classes (amino-alcohol and
quinazolinones) with partially different binding modes to CaSR [68] exert very similar biochemical
effects is a strong indicator that these effects in patients are specific to calcilytics, and are not a
solitary effect of a single compound or chemical class.

In principle, similar effects on serum calcium and phosphate might be obtained with PTH
treatment. In three study cohorts, calcilytics were therefore compared to subcutaneous

Trends in Endocrinology & Metabolism, September 2016, Vol. 27, No. 9 649
injections of 20 mg recombinant human (rh)PTH [71,72,76–78]. In all three studies, rhPTH indeed Outstanding Questions
raised serum calcium [71,72,76–78], but ronacaleret and AXT914 were more efficacious in How efficacious are calcilytics in patients
terms of raising serum calcium and yielded a greater cumulative increase in serum PTH levels with activating CaSR mutations?

than rhPTH injections [71,72,76–78].


Will calcilytics reduce renal calcium
excretion in patients and tissue com-
In two study cohorts, effects on urinary calcium excretion were also studied. In the time intervals plications in the long term?
of 0–4 h, 4–8 h, and 8–12 h after administration, both ronacaleret and rhPTH decreased urinary
calcium excretion, but even 100 mg ronacaleret was more efficacious than rhPTH [77]. With Are there differences in the efficacy of
different calcilytics on different mutant
400 mg ronacaleret, fractional calcium excretion dropped from 1.5% to 0.2% on average, in
proteins?
other words calcium excretion dropped by 80–90% in the first 12 h [77] but returned to baseline
12–24 h after administration. This is in perfect agreement with the temporal serum concentration Are calcilytics efficacious in patients
profile of PTH and of calcilytics designed to have a short serum half-life. This may also explain harboring deletion mutations?
why, after chronic administration of ronacaleret and PTH once daily for several weeks, serum
calcium levels and total urinary calcium excretion increase [71]. Presumably, raised serum Are calcilytics efficacious against renal
salt-wasting in BS5?
calcium levels lead to greater filtered calcium load, which increases urinary calcium excretion
once the effects of calcilytics or PTH have abated after 12 h. This suggests that administration of
Are long-acting calcilytics more effica-
available substances two or three times a day, sustained release dosage forms, or longer-acting cious than current short-acting
calcilytics may be more effective in the long term (see Outstanding Questions). In line with calcilytics?
previous results, the overall effect of chronic calcilytic treatment on urinary excretion was
stronger with calcilytics than with rhPTH [77]. What are the appropriate therapeutic
target levels for serum calcium and uri-
nary calcium excretion in ADH and BS5
Concluding Remarks and Future Perspectives patients?
The unique advantage of calcilytics over other drugs is that calcilytics can break the detrimental
predominance of activated CaSR over other mechanisms that act on calcium excretion. Are calcilytics safe in patients with acti-
However, it will be crucial to determine in clinical trials whether calcilytics will indeed reduce vating CaSR mutations in the long
term?
renal calcium excretion and complications in the long term and whether longer or continuously
acting modes of calcilytic administration will have benefits over the existing relatively short-acting
substances (see Outstanding Questions).

These clinical trials will also need to establish the safety and possible adverse effects of calcilytic
treatment of ADH and BS5. In clinical trials with healthy individuals the most common adverse
effects of calcilytics were mild disorders of the gastrointestinal tract and nervous system, such as
fatigue, headache, constipation, diarrhea, nausea, and dyspepsia [76,81], which might be related
to CaSR expressed in gut [82] and brain [83]. CaSR is also present in skin, lung, heart, mammary
glands, and numerous other tissues [18,84–86], but the physiological and pharmacological
significance of CaSR expression in these tissues has been disputed [87]. Potential short- and
long-term adverse effects of calcilytics cannot be ruled out, and will need to be evaluated in clinical
studies before calcilytics can be used for routine medical treatment of ADH and BS5 patients.

Taken together, the available data suggest that calcilytics could correct the problem of activating
CaSR mutations at the root. Redeployment of these already existing drugs to treat patients with
ADH or BS5 appears to be the most promising future perspective for these rare but severe and
difficult to treat diseases.

Resources
i
www.casrdb.mcgill.ca
ii
clinicaltrials.gov

References
1. Mayr, B.M. et al. (2015) Genetics in endocrinology: gain and loss of 2. Pollak, M.R. et al. (1993) Mutations in the human Ca2+-sensing
function mutations of the calcium sensing receptor and associated receptor gene cause familial hypocalciuric hypercalcemia
proteins: current treatment concepts. Eur. J. Endocrinol. 174, and neonatal severe hyperparathyroidism. Cell 75, 1297–1303
R189–R208

650 Trends in Endocrinology & Metabolism, September 2016, Vol. 27, No. 9
3. Nemeth, E.F. and Goodman, W.G. (2016) Calcimimetic and calci- express functional calcium-sensing receptor. Am. J. Physiol. Renal
lytic drugs: feats, flops, and futures. Calcif. Tissue Int. 98, 341–358 Physiol. 308, F1200–F1206
4. Hu, J. and Spiegel, A.M. (2007) Structure and function of the 26. Bapty, B.W. et al. (1998) Extracellular Mg2+- and Ca2+-sensing in
human calcium-sensing receptor: insights from natural and engi- mouse distal convoluted tubule cells. Kidney Int. 53, 583–592
neered mutations and allosteric modulators. J. Cell. Mol. Med. 11, 27. De Jesus Ferreira, M.C. and Bailly, C. (1998) Extracellular Ca2+
908–922 decreases chloride reabsorption in rat CTAL by inhibiting cAMP
5. Mun, H.C. et al. (2004) The venus fly trap domain of the extracel- pathway. Am. J. Physiol. 275, F198–F203
lular Ca2+-sensing receptor is required for L-amino acid sensing. J. 28. Canaff, L. and Hendy, G.N. (2002) Human calcium-sensing recep-
Biol. Chem. 279, 51739–51744 tor gene. Vitamin D response elements in promoters P1 and P2
6. Leach, K. et al. (2016) Towards a structural understanding of confer transcriptional responsiveness to 1,25-dihydroxyvitamin D.
allosteric drugs at the human calcium-sensing receptor. Cell J. Biol. Chem. 277, 30337–30350
Res. 26, 574–592 29. Wills, M.R. (1969) The urinary calcium–creatinine ratio as a
7. Hofer, A.M. and Brown, E.M. (2003) Extracellular calcium sensing measure of urinary calcium excretion. J. Clin. Pathol. 22,
and signalling. Nat. Rev. Mol. Cell Biol. 4, 530–538 287–290
8. Ward, D.T. and Riccardi, D. (2012) New concepts in calcium- 30. Marx, S.J. et al. (1977) Family studies in patients with primary
sensing receptor pharmacology and signalling. Br. J. Pharmacol. parathyroid hyperplasia. Am. J. Med. 62, 698–706
165, 35–48 31. Pollak, M.R. et al. (1994) Autosomal dominant hypocalcaemia
9. Bai, M. et al. (1996) Expression and characterization of inactivating caused by a Ca2+-sensing receptor gene mutation. Nat. Genet.
and activating mutations in the human Ca2+-sensing receptor. J. 8, 303–307
Biol. Chem. 271, 19537–19545 32. Pearce, S.H. et al. (1996) A familial syndrome of hypocalcemia with
10. Breitwieser, G.E. (2013) The calcium sensing receptor life cycle: hypercalciuria due to mutations in the calcium-sensing receptor.
trafficking, cell surface expression, and degradation. Best Pract. N. Engl. J. Med. 335, 1115–1122
Res. Clin. Endocrinol. Metab. 27, 303–313 33. Seyberth, H.W. and Schlingmann, K.P. (2011) Bartter- and Gitel-
11. Nesbit, M.A. et al. (2013) Mutations in AP2S1 cause familial man-like syndromes: salt-losing tubulopathies with loop or DCT
hypocalciuric hypercalcemia type 3. Nat. Genet. 45, 93–97 defects. Pediatr. Nephrol. 26, 1789–1802
12. Lorenz, S. et al. (2007) Functional desensitization of the extracel- 34. Choi, K.H. et al. (2015) Autosomal dominant hypocalcemia with
lular calcium-sensing receptor is regulated via distinct mecha- Bartter syndrome due to a novel activating mutation of calcium
nisms: role of G protein-coupled receptor kinases, protein sensing receptor, Y829C. Korean J. Pediatr. 58, 148–153
kinase C and beta-arrestins. Endocrinology 148, 2398–2404 35. Watanabe, S. et al. (2002) Association between activating muta-
13. Breitwieser, G.E. (2012) Minireview: the intimate link between cal- tions of calcium-sensing receptor and Bartter's syndrome. Lancet
cium sensing receptor trafficking and signaling: implications for 360, 692–694
disorders of calcium homeostasis. Mol. Endocrinol. 26, 1482–1495 36. Raue, F. et al. (2011) Activating mutations in the calcium-sensing
14. Grant, M.P. et al. (2011) Agonist-driven maturation and plasma receptor: genetic and clinical spectrum in 25 patients with auto-
membrane insertion of calcium-sensing receptors dynamically somal dominant hypocalcaemia – a German survey. Clin. Endo-
control signal amplitude. Sci. Signal. 4, ra78 crinol. 75, 760–765
15. Lazarus, S. et al. (2011) A novel mutation of the primary protein 37. Sayer, J.A. and Pearce, S.H. (2003) Extracellular calcium-sensing
kinase C phosphorylation site in the calcium-sensing receptor receptor dysfunction is associated with two new phenotypes. Clin.
causes autosomal dominant hypocalcemia. Eur. J. Endocrinol. Endocrinol. 59, 419–421
164, 429–435 38. Kurozumi, A. et al. (2013) Extrapyramidal symptoms and
16. Ray, K. et al. (1997) The carboxyl terminus of the human calcium advanced calcification of the basal ganglia in a patient with auto-
receptor. Requirements for cell-surface expression and signal somal dominant hypocalcemia. Intern. Med. 52, 2077–2081
transduction. J. Biol. Chem. 272, 31355–31361 39. Sorheim, J.I. et al. (2010) Phenotypic variation in a large family with
17. Abou-Samra, A.B. et al. (1992) Expression cloning of a common autosomal dominant hypocalcaemia. Horm. Res. Paediatr. 74,
receptor for parathyroid hormone and parathyroid hormone- 399–405
related peptide from rat osteoblast-like cells: a single receptor 40. Thim, S.B. et al. (2014) Activating calcium-sensing receptor gene
stimulates intracellular accumulation of both cAMP and inositol variants in children: a case study of infant hypocalcaemia and
trisphosphates and increases intracellular free calcium. Proc. Natl. literature review. Acta Paediatr. 103, 1117–1125
Acad. Sci. U.S.A. 89, 2732–2736
41. Kinoshita, Y. et al. (2014) Functional activities of mutant calcium-
18. Magno, A.L. et al. (2011) The calcium-sensing receptor: a molec- sensing receptors determine clinical presentations in patients with
ular perspective. Endocr. Rev. 32, 3–30 autosomal dominant hypocalcemia. J. Clin. Endocrinol. Metab.
19. Riccardi, D. et al. (2013) The extracellular calcium-sensing recep- 99, E363–E368
tor, CaSR, in fetal development. Best Pract. Res. Clin. Endocrinol. 42. Hannan, F.M. et al. (2012) Identification of 70 calcium-sensing
Metab. 27, 443–453 receptor mutations in hyper- and hypo-calcaemic patients: evi-
20. Graca, J.A. et al. (2016) Comparative expression of the extracel- dence for clustering of extracellular domain mutations at calcium-
lular calcium-sensing receptor in the mouse, rat, and human binding sites. Hum. Mol. Genet. 21, 2768–2778
kidney. Am. J. Physiol. Renal Physiol. 310, F518–F533 43. Hannan, F.M. et al. (2012) A calcium-sensing receptor (CaSR)
21. Motoyama, H.I. and Friedman, P.A. (2002) Calcium-sensing variant, Glu250Lys, present in familial hypocalciuric hypercalcae-
receptor regulation of PTH-dependent calcium absorption by mia (FHH) and autosomal dominant hypocalcaemic hypercalciuria
mouse cortical ascending limbs. Am. J. Physiol. Renal Physiol. (ADHH) probands represents a functionally neutral polymorphism:
283, F399–F406 lessons for CaSR mutational analysis. Endocrine Abstracts 28,
22. Riccardi, D. and Valenti, G. (2016) Localization and function of the OC2.8
renal calcium-sensing receptor. Nat. Rev. Nephrol. http://dx.doi. 44. Conley, Y.P. et al. (2000) Three novel activating mutations in the
org/10.1038/nrneph.2016.59 calcium-sensing receptor responsible for autosomal dominant
23. Kantham, L. et al. (2009) The calcium-sensing receptor (CaSR) hypocalcemia. Mol. Genet. Metab. 71, 591–598
defends against hypercalcemia independently of its regulation of 45. Nissen, P.H. et al. (2007) Molecular genetic analysis of the calcium
parathyroid hormone secretion. Am. J. Physiol. Endocrinol. sensing receptor gene in patients clinically suspected to have
Metab. 297, E915–E923 familial hypocalciuric hypercalcemia: phenotypic variation and
24. Loupy, A. et al. (2012) PTH-independent regulation of blood mutation spectrum in a Danish population. J. Clin. Endocrinol.
calcium concentration by the calcium-sensing receptor. J. Clin. Metab. 92, 4373–4379
Invest. 122, 3355–3367 46. Pearce, S.H. et al. (1996) Calcium-sensing receptor mutations in
25. Di Mise, A. et al. (2015) Conditionally immortalized human proximal familial hypocalciuric hypercalcaemia with recurrent pancreatitis.
tubular epithelial cells isolated from the urine of a healthy subject Clin. Endocrinol. 45, 675–680

Trends in Endocrinology & Metabolism, September 2016, Vol. 27, No. 9 651
47. Silve, C. et al. (2005) Delineating a Ca2+ binding pocket within the 68. Widler, L. (2011) Calcilytics: antagonists of the calcium-sensing
venus flytrap module of the human calcium-sensing receptor. J. receptor for the treatment of osteoporosis. Future Med. Chem. 3,
Biol. Chem. 280, 37917–37923 535–547
48. Zhang, C. et al. (2014) Role of Ca2+ and L-Phe in regulating 69. Hannan, F.M. et al. (2015) The calcilytic agent NPS 2143 rectifies
functional cooperativity of disease-associated ‘toggle’ calcium- hypocalcemia in a mouse model with an activating calcium-sens-
sensing receptor mutations. PloS One 9, e113622 ing receptor (CaSR) mutation: relevance to autosomal dominant
49. Lienhardt, A. et al. (2000) A large homozygous or heterozygous in- hypocalcemia type 1 (ADH1). Endocrinology 156, 3114–3121
frame deletion within the calcium-sensing receptor's carboxylter- 70. Hough, T.A. et al. (2004) Activating calcium-sensing receptor
minal cytoplasmic tail that causes autosomal dominant hypocal- mutation in the mouse is associated with cataracts and ectopic
cemia. J. Clin. Endocrinol. Metab. 85, 1695–1702 calcification. Proc. Natl. Acad. Sci. U.S.A. 101, 13566–13571
50. Obermannova, B. et al. (2016) Novel calcium-sensing receptor 71. Caltabiano, S. et al. (2013) Characterization of the effect of chronic
cytoplasmic tail deletion mutation causing autosomal dominant administration of a calcium-sensing receptor antagonist, ronaca-
hypocalcemia: molecular and clinical study. Eur. J. Endocrinol. leret, on renal calcium excretion and serum calcium in postmeno-
174, K1–K11 pausal women. Bone 56, 154–162
51. Mannstadt, M. et al. (2013) Germline mutations affecting Galpha11 72. Fitzpatrick, L.A. et al. (2011) The effects of ronacaleret, a calcium-
in hypoparathyroidism. N. Engl. J. Med. 368, 2532–2534 sensing receptor antagonist, on bone mineral density and bio-
52. Li, D. et al. (2014) Autosomal dominant hypoparathyroidism chemical markers of bone turnover in postmenopausal women
caused by germline mutation in GNA11: phenotypic and molecular with low bone mineral density. J. Clin. Endocrinol. Metab. 96,
characterization. J. Clin. Endocrinol. Metab. 99, E1774–E1783 2441–2449

53. Nesbit, M.A. et al. (2013) Mutations affecting G-protein subunit 73. Kumar, S. et al. (2010) An orally active calcium-sensing receptor
alpha11 in hypercalcemia and hypocalcemia. N. Engl. J. Med. antagonist that transiently increases plasma concentrations of
368, 2476–2486 PTH and stimulates bone formation. Bone 46, 534–542
54. Piret, S.E. et al. (2016) Identification of a G-protein subunit- 74. Cosman, F. et al. (2016) A phase 2 study of MK-5442, a calcium-
alpha11 gain-of-function mutation, Val340Met, in a family with sensing receptor antagonist, in postmenopausal women with
autosomal dominant hypocalcemia type 2 (ADH2). J. Bone Miner. osteoporosis after long-term use of oral bisphosphonates. Oste-
Res. http://dx.doi.org/10.1002/jbmr.2797 oporosis Int. 27, 377–386

55. Leach, K. et al. (2012) Identification of molecular phenotypes and 75. Fukumoto, S. (2011) Antagonist for calcium-sensing receptor:
biased signaling induced by naturally occurring mutations of the JTT-305/MK-5442. Clin. Calcium 21, 89–93 (article in Japanese)
human calcium-sensing receptor. Endocrinology 153, 4304–4316 76. John, M.R. et al. (2014) AXT914 a novel, orally-active parathyroid
56. Leach, K. et al. (2013) Impact of clinically relevant mutations on the hormone-releasing drug in two early studies of healthy volunteers
pharmacoregulation and signaling bias of the calcium-sensing and postmenopausal women. Bone 64C, 204–210
receptor by positive and negative allosteric modulators. Endocri- 77. Caltabiano, S. et al. (2009) Characterization of the effect of rona-
nology 154, 1105–1116 caleret, a calcium-sensing receptor antagonist, on renal calcium
57. Chikatsu, N. et al. (2003) A family of autosomal dominant hypo- excretion. J. Bone Miner. Res. 24 (Suppl. 1), A1051
calcemia with an activating mutation of calcium-sensing receptor 78. Fitzpatrick, L.A. et al. (2012) Ronacaleret, a calcium-sensing
gene. Endocr. J. 50, 91–96 receptor antagonist, increases trabecular but not cortical bone
58. Tan, Y.M. et al. (2003) Autosomal dominant hypocalcemia: a novel in postmenopausal women. J. Bone Miner. Res. 27, 255–262
activating mutation (E604K) in the cysteine-rich domain of the 79. Fitzpatrick, L.A. et al. (2011) Ronacaleret, a calcium-sensing
calcium-sensing receptor. J. Clin. Endocrinol. Metab. 88, 605–610 receptor antagonist, has no significant effect on radial fracture
59. Mitchell, D.M. et al. (2012) Long-term follow-up of patients with healing time: results of a randomized, double-blinded, placebo-
hypoparathyroidism. J. Clin. Endocrinol. Metab. 97, 4507–4514 controlled Phase II clinical trial. Bone 49, 845–852
60. Bollerslev, J. et al. (2015) European Society of Endocrinology 80. Halse, J. et al. (2014) A phase 2, randomized, placebo-controlled,
clinical guideline: treatment of chronic hypoparathyroidism in dose-ranging study of the calcium-sensing receptor antagonist
adults. Eur. J. Endocrinol. 173, G1–G20 MK-5442 in the treatment of postmenopausal women with oste-
oporosis. J. Clin. Endocrinol. Metab. 99, E2207–E2215
61. Dong, B. et al. (2015) Calcilytic ameliorates abnormalities of
mutant calcium-sensing receptor (CaSR) knock-in mice mimicking 81. Fitzpatrick, L. et al. (2009) Ronacaleret, a calcium-sensing recep-
autosomal dominant hypocalcemia (ADH). J. Bone Miner. Res. 30, tor antagonist: results of a 1 year double-blind, placebo-con-
1980–1993 trolled, dose-ranging phase II study. J. Bone Miner. Res. 24
(Suppl. 1), A1130
62. Mittelman, S.D. et al. (2006) A hypocalcemic child with a novel
activating mutation of the calcium-sensing receptor gene: suc- 82. Macleod, R.J. (2013) CaSR function in the intestine: hormone
cessful treatment with recombinant human parathyroid hormone. secretion, electrolyte absorption and secretion, paracrine non-
J. Clin. Endocrinol. Metab. 91, 2474–2479 canonical Wnt signaling and colonic crypt cell proliferation. Best
Pract. Res. Clin. Endocrinol. Metab. 27, 385–402
63. Gonzales, M.C. et al. (2013) Recombinant human parathyroid
hormone therapy (1-34) in an adult patient with a gain-of-function 83. Liu, X.L. et al. (2013) Calcium sensing receptor absence delays
mutation in the calcium-sensing receptor-a case report. Endocr. postnatal brain development via direct and indirect mechanisms.
Pract. 19, e24–e28 Mol. Neurobiol. 48, 590–600

64. Schellhaas, B. et al. (2015) Teriparatide as a novel therapeutic 84. Brennan, S.C. et al. (2016) The extracellular calcium-sensing
option in the treatment of ADH – a case report. Exp. Clin. Endo- receptor regulates human fetal lung development via CFTR. Sci.
crinol. Diabetes 123, P01_01 Rep. 6, 21975

65. Theman, T.A. et al. (2009) PTH(1-34) replacement therapy in a 85. Ahearn, T.U. et al. (2016) Calcium sensing receptor tumor expres-
child with hypoparathyroidism caused by a sporadic calcium sion and lethal prostate cancer progression. J. Clin. Endocrinol.
receptor mutation. J. Bone Miner. Res. 24, 964–973 Metab. http://dx.doi.org/10.1210/jc.2016-1082

66. Babinsky, V.N. et al. (2016) Allosteric modulation of the cal- 86. Cheng, S.X. (2016) Calcium-sensing receptor: a new target for
cium-sensing receptor rectifies signaling abnormalities associ- therapy of diarrhea. World J. Gastroenterol. 22, 2711–2724
ated with G-protein alpha-11 mutations causing hypercalcemic 87. Nemeth, E.F. (2006) Misconceptions about calcimimetics. Ann. N.
and hypocalcemic disorders. J. Biol. Chem. 291, 10876– Y. Acad. Sci. 1068, 471–476
10885 88. Fukumoto, S. et al. (2009) Randomized, single-blinded placebo-
67. Ramnitz, M. et al. (2015) Treatment of autosomal dominant hypo- controlled study of a novel calcilytic, JTT-305, in patients
calcemia with the calcilytic NPSP795. J. Bone Miner. Res. 30 with postmenopausal osteoporosis. J. Bone Miner. Res. 24
(Suppl. 1), SA0002 (Suppl. 1), A1131

652 Trends in Endocrinology & Metabolism, September 2016, Vol. 27, No. 9

Das könnte Ihnen auch gefallen