Skip to main content
The Journal of Clinical Endocrinology and Metabolism logoLink to The Journal of Clinical Endocrinology and Metabolism
. 2012 Jan 11;97(2):311–325. doi: 10.1210/jc.2011-2332

Update on Bone Anabolics in Osteoporosis Treatment: Rationale, Current Status, and Perspectives

Roland Baron 1,, Eric Hesse 1
PMCID: PMC3275361  PMID: 22238383

Abstract

Osteoporosis is defined as low bone mineral density associated with skeletal fractures secondary to minimal or no trauma, most often involving the spine, the hip, and the forearm. The decrease in bone mineral density is the consequence of an unbalanced bone remodeling process, with higher bone resorption than bone formation. Osteoporosis affects predominantly postmenopausal women, but also older men. This chronic disease represents a considerable medical and socioeconomic burden for modern societies. The therapeutic options for the treatment of osteoporosis have so far comprised mostly antiresorptive drugs, in particular bisphosphonates and more recently denosumab, but also calcitonin and, for women, estrogens or selective estrogen receptor modulators. These drugs have limitations, however, in particular the fact that they lead to a low turnover state where bone formation decreases with the decrease in bone-remodeling activity. In this review, we discuss the alternative class of osteoporosis drugs, i.e. bone anabolics, their biology, and the perspectives they offer for our therapeutic armamentarium. We focus on the two main osteoanabolic pathways identified as of today: PTH, the only anabolic drug currently on the market; and activation of canonical Wnt signaling through inhibition of the endogenous inhibitors sclerostin and dickkopf1. Each approach is based on a different molecular mechanism, but most recent evidence suggests that these two pathways may actually converge, at least in part. Whereas recombinant human PTH treatment is being revisited with different formulations and attempts to regulate endogenous PTH secretion via the calcium-sensing receptor, antibodies to sclerostin and dickkopf1 are currently in clinical trials and may prove to be even more efficient at increasing bone mass, possibly independent of bone turnover. Each of these anabolic approaches has its own limitations and safety issues, but the prospects of effective anabolic therapy for osteoporosis are indeed bright.


Osteoporosis is the result of a dysfunction of endocrine and/or autocrine/paracrine factors and/or their target cells in bone, leading to the inability to reach a proper peak bone mass and/or to maintain skeletal homeostasis. These alterations, together with genetic determinants and mechanical and nutritional cues, cause a decrease in bone density, alterations in bone microarchitecture, and ultimately fractures. Osteoporosis is predominantly a disease of aging, affecting particularly postmenopausal women but also older men. The coordinated actions of bone cells that become disturbed in osteoporosis occur according to two main biological principles, bone modeling and remodeling.

Bone Modeling and Remodeling

The development and maintenance of mammalian bones depends on the coordinated actions of matrix-resorbing hematopoietic lineage-derived osteoclasts and matrix-producing mesenchyme-derived osteoblasts. During bone modeling, the process that shapes skeletal elements at developmental stages but also at a low pace throughout life, bone resorption and formation occur in an uncoupled manner and on separate surfaces. In contrast, bone remodeling, the mechanism that ensures tissue turnover while maintaining bone mass in the adult, is based on the coupled and balanced activities of bone resorption and formation within each basic multicellular unit (BMU). BMUs are constituted of cells of both lineages, which are active at specific times during the remodeling cycle. These packages of cells are located along the bone surface, mostly at the interface with the hematopoietic bone marrow (endosteum) but also at the surface of bones (periosteum). BMUs are initiated through the activation of bone resorption, which is followed by bone formation. Within each BMU, activities are “coupled” if bone formation follows bone resorption, and activities are “balanced” if the amount of bone formed by osteoblasts equals and compensates for the amount of bone that was previously resorbed by osteoclasts (Fig. 1A). Stimulating bone remodeling increases bone turnover through an increase in the number of BMUs per bone surface area, also called activation frequency (1). In osteoporosis, within a BMU both coupled but unbalanced or uncoupled bone remodeling can cause severe alterations in bone mass, which will increase in severity proportionally to the activation frequency, i.e. with the turnover rate (Fig. 1B).

Fig. 1.

Fig. 1.

Schematic of the remodeling and modeling activities under physiological conditions, in osteoporosis, and during anabolic treatment. A, Within an active BMU under physiological conditions, bone is constantly removed by osteoclasts (OCs) during the resorption phase of the remodeling cycle. After the reversal phase, new bone matrix is produced by osteoblasts (OBs) during the formation phase at sites where bone resorption has occurred with the amount of bone formed being equal of the amount of bone resorbed, thereby maintaining bone mass. Once the BMU is completed, osteoblasts become entrapped as osteocytes (OCYs) into the newly formed matrix, remain on the bone surface as lining cells (LCs), or undergo apoptosis. Bone then remains in the quiescence phase until a new BMU is initiated. B, In osteoporosis, bone resorption is increased and bone formation is decreased, resulting in a loss of bone. C, Administration of recombinant human PTH (rhPTH) stimulates both osteoclast-mediated bone resorption and osteoblast-mediated bone formation, resulting in a high bone turnover with a net gain in bone mass. In addition to its remodeling-based bone anabolic effect, rhPTH also induces bone formation on surfaces around the resorption sites that were not previously subject to bone resorption (modeling). D, Activation of the canonical Wnt signaling pathway tends to decrease bone resorption but mostly increases both remodeling-based and modeling-based bone formation, thereby causing a striking increase in bone formation, particularly in areas that were not previously resorbed (modeling).

During a remodeling cycle, preosteoclasts are activated, migrate, and fuse to mature osteoclasts at sites where bone matrix needs to be replaced due to diminished matrix quality, cell viability/metabolism, or microfractures. At the end of the resorption phase (approximately 1–2 wk in humans), osteoclasts recruit and are replaced by osteoblasts through active cross talk between these two cell lineages, and bone formation begins. During the bone formation phase (approximately 2–3 months in humans), osteoblasts lay down bone matrix, which then mineralizes. The rate at which this occurs is the mineral apposition rate (MAR), which reflects the activity of individual osteoblasts. The bone formation rate (BFR) is the MAR multiplied by the surfaces undergoing bone formation. Both are true measures of the bone-forming activity in an individual (1). At the end of the bone formation phase, osteoblasts become quiescent as bone-lining cells on the surface of the newly formed bone, die by apoptosis, or become included within the matrix as osteocytes (Fig. 1A). Osteocytes are not merely “old” osteoblasts but have emerged as key cells that contribute to the regulation of calcium (Ca2+) and phosphorus metabolism through the control of bone remodeling and Ca2+ fluxes and the secretion of fibroblast growth factor 23, respectively. Osteocytes also secrete sclerostin, a protein that inhibits bone formation, and sense compromised bone matrix, thereby stimulating osteoclast recruitment and the generation of a new remodeling cycle. Furthermore, two recent studies demonstrate that osteocytes are an important source of receptor activator of NF-κB ligand (RANKL). RANKL binds to the RANK receptor on osteoclast precursors and mature osteoclasts and stimulates osteoclastogenesis and bone resorption (101, 102). Thus, osteocytes regulate bone resorption and formation in the context of both bone modeling and remodeling (2).

Osteoporosis

Osteoporosis is a systemic skeletal disease characterized by an unbalanced and/or uncoupled bone-remodeling activity leading to bone loss (Fig. 1B), microarchitectural deterioration of bone, and ultimately fractures at typical sites such as the lumbar spine, the femoral neck, and the distal radius. These fractures are often associated with an increase in morbidity and mortality. Because of its widespread nature, with a 50% fracture risk in all women after the age of 50 yr and a 25% risk in men, osteoporosis is a global public health concern and a great socioeconomic burden (3).

The goal of any osteoporosis therapy is the prevention of both vertebral (mostly dependent on trabecular bone density and architecture) and nonvertebral (mostly dependent on cortical thickness and porosity) fractures, which in principle can be achieved by inhibiting bone resorption and/or by stimulating bone formation. Yet, the dependence of trabecular and cortical bone on remodeling or modeling activity is different, with cortical bone being more susceptible to modeling activity, particularly along its periosteal surface. This difference may in part be responsible for the relative lack of efficacy of antiresorptive drugs on nonvertebral fractures because their effects are restricted to remodeling-based activities. Current antiresorptive drugs decrease the activation frequency, thereby causing a secondary decrease in BFR. This culminates in a low bone turnover, which in turn limits further increases in trabecular bone mass. In addition, questions have been raised about the association of long-term treatment in osteoporosis and high-dose use of these agents in oncology and clinical complications such as osteonecrosis of the jaw and so-called “atypical” subtrochanteric fractures (4).

Anabolic therapies are dependent on increasing the activation frequency and favoring bone formation within the BMU, on directly stimulating bone formation through activation of bone modeling, or on a combination of both (Fig. 1, C and D). Thus, true bone anabolics are defined by their ability to increase bone formation, as measured by biochemical markers procollagen type 1 amino-terminal propeptide (P1NP) and bone-specific alkaline phosphatase, and histomorphometric parameters (MAR and BFR) on bone biopsies.

The two main bone anabolic pathways are PTH signaling and canonical wingless-int (Wnt) signaling. Of the two, the canonical Wnt pathway might be more dependent on increasing bone modeling, potentially increasing bone mass in patients independent of bone resorption and activation frequency/bone turnover. In contrast, PTH anabolic function is more dependent on increasing the activation frequency, which may in part limit its therapeutic window (see below).

Given the limitations of current antiosteoporosis drugs, a search for new therapeutics has focused in the last few years on also identifying novel antiresorptives that prevent the decrease in activation frequency and bone formation and on bone anabolics that increase bone formation directly without affecting bone resorption. In this review, we will focus on bone anabolics and discuss their mode of action, limitations, and promises for the near future.

Although other biological agents, such as bone morphogenetic proteins (5) or IGF (6), are theoretically capable of increasing bone formation, either locally or systemically, practical limitations of their use and/or systemic effects outside of the skeleton have so far prevented their development in osteoporosis. In this review, we will focus only on the approaches that are currently in the clinic or in clinical trials.

Intermittent PTH: the Current Anabolic Option

PTH is secreted by the parathyroid glands in response to a reduced serum Ca2+ concentration and normalizes the Ca2+ levels by enhancing Ca2+ uptake in the intestine, Ca2+ re-absorption in the kidney, and by stimulating the osteoclast- and osteocyte-mediated Ca2+ release from bone. Mechanistically, PTH binds to the PTH receptor (PTH1R), a class II G protein-coupled receptor that activates several signaling pathways, including the Gsα-linked cAMP-dependent protein kinase A signaling pathway and the Gq/11-linked phosphatidyl inositol-specific phospholipase C (PLC)-protein kinase C signaling pathway. In vivo studies have been conducted to determine the specific role of these distinct PTH1R signaling pathways in bone. For instance, Guo et al. (7) reported a mouse that expresses a mutant PTH1R (DSEL), which stimulates adenylyl cyclase but is unable to activate PLC signaling. At 10 wk of age, DSEL-mutant mice demonstrate a low trabecular bone mass, showing the importance of the PTH1R-mediated activation of the PLC signaling pathway for normal bone homeostasis (7). As revealed recently, binding of PTH to the PTH1R also activates the canonical Wnt-signaling pathway, which is discussed below in greater detail. In addition, PTH increases the commitment of mesenchymal precursor cells to the osteoblast lineage, promotes osteoblast maturation, and inhibits osteoblast apoptosis, thereby increasing osteoblast number and function (Fig. 2) (8).

Fig. 2.

Fig. 2.

Effects of the two main anabolic pathways, PTH and Wnt signaling, on osteoblasts, and indirectly on osteoclasts. PTH and Wnt both stimulate the proliferation of mesenchymal stem cells (MSCs) and the commitment of these cells into the osteoblast (OB) lineage, whereas the differentiation into chondrocytes and adipocytes is prevented by canonical Wnt signaling. In the late OB, both pathways increase the mineral apposition and bone formations rate (MAR, BFR). In addition, the Wnt pathway stimulates the production of osteoprotegerin (OPG), a soluble decoy receptor for the RANKL, preventing osteoclast (OC) differentiation and function. In contrast, PTH stimulates the secretion of RANKL, which binds to its receptor RANK on OC precursor cells (OC-pre) and mature OCs, thereby stimulating OC differentiation, function, and ultimately bone resorption. Wnt activity is inhibited by sclerostin and Dkk1, both secreted by late OBs and osteocytes. PTH represses the expression of both sclerostin and Dkk1, whereas Dkk1 expression is increased by Wnt activity, establishing a negative feedback loop. Thus, PTH and Wnt signaling pathways increase bone formation through several mechanisms, but only the Wnt pathway represses bone resorption, whereas PTH stimulates OCs via the induction of RANKL production by OBs.

Analysis of biopsies from patients with primary hyperparathyroidism called the attention of the field to the effects of PTH on bone remodeling. Bone exposed to sustained high levels of PTH show a marked increase in activation frequency and bone resorption but also in osteoblast numbers and BFR. Although trabecular bone density is often unchanged or even slightly increased, the enhanced bone resorption leads to an increased cortical porosity. Osteoblasts, but not osteoclasts, were found to express the PTH1R and to respond with an increase not only in proliferation and differentiation, but also in the secretion of RANKL. RANKL binds to the RANK receptor on osteoclast precursors and mature osteoclasts and stimulates osteoclastogenesis and bone resorption. This led to the conclusion that the mechanism of action is primarily an increase in bone formation and only secondarily, through the cross talk between osteoblasts and osteoclasts, an increase in bone resorption. Animal studies then demonstrated that short exposures to recombinant human PTH (rhPTH), as opposed to sustained increases, could dissociate the positive bone anabolic response from the negative bone catabolic response, and this led to the development and marketing of rhPTH.

To date, injectable forms of rhPTH are the only approved osteoanabolic drugs on the market for the treatment of osteoporosis. Although the intact hormone rhPTH 1–84 (Preotact) is approved only by European regulatory agencies, the bioactive N-terminal 34-amino acid fragment rhPTH 1–34 (teriparatide, Forsteo, Forteo) is available in the United States and Europe. Upon sc injection, both forms rapidly reach peak concentrations and are degraded in about 1 h. rhPTH reduces the fracture risk of the spine much greater than that of nonvertebral bones, perhaps owing to site-specific differences, with cortical bone being less positively affected than trabecular bone (9). Indeed, rhPTH typically does not increase bone mineral density (BMD), which even often decreases, in the distal 1/3 of the forearm. This is due to an increase in cortical porosity secondary to enhanced endocortical remodeling (9, 10). In contrast, rhPTH increases bone formation along the periosteum, a primarily bone modeling surface, perhaps contributing to improving the trabecular and cortical architecture (11). Thus, in addition to the remodeling-based increase in bone formation, rhPTH also induces modeling-based bone formation, and this also occurs on surfaces adjacent to the BMU (Fig. 1C) (12). Clinical approval of teriparatide by the U.S. Food and Drug Administration was based on clinical trials including more than 2800 osteoporosis patients. In the pivotal clinical trial of Neer et al. (9), rhPTH (1–34) lowered the risk of vertebral fractures by 65% and that of nonvertebral fractures by 40% compared with placebo during the 19 months of treatment.

Several factors seem to limit the effectiveness of rhPTH. As mentioned above, in response to rhPTH, osteoblasts not only produce bone matrix but also secrete growth factors and cytokines including RANKL, thereby stimulating osteoclastogenesis. Thus, even if administered intermittently, chronic use of rhPTH increases bone formation in part through an increase in the activation frequency (remodeling-based anabolic), and this ultimately leads also to an increase in bone resorption. Although the net effect is still a gain in cancellous bone mass at early time points, it appears that bone resorption slowly catches up with bone formation, leading to a plateauing of the net anabolic effect after 18–24 months (9). Another possible reason that limits the use of rhPTH therapy is the progressive decrease in responsiveness secondary to tachyphylaxis, or a depletion of the pool of mesenchymal osteoblast precursors, or both (13).

Thus, administration of an antiresorptive drug combined with rhPTH could further increase bone mass by blunting the rhPTH-activated bone resorption. Although clinical studies have generated inconsistent results, with Black et al. (14) and Finkelstein et al. (15) reporting blunting of the anabolic response to rhPTH in patients under alendronate treatment, a recent report demonstrates that administration of a single dose of zoledronic acid combined with daily sc injections of rhPTH increased hip and spine BMD greater and more rapidly than either drug alone, suggesting that this combination therapy could be beneficial for patients with a high fracture risk (16). Furthermore, sequential administration of antiresorptive drugs after rhPTH is already an established treatment protocol because bone resorption continues after cessation of the treatment, causing a 4% bone loss in the first year after rhPTH withdrawal (17).

Potential Concerns

Other factors that have limited the use of rhPTH are its cost and concerns about its potential link to osteosarcoma. Indeed, treatment of osteoporosis with rhPTH is limited to 24 months in the United States and 18 months in Europe due to the risk of cancer because treatment of rats with high doses of rhPTH 1–34 increased the prevalence of osteosarcoma (18). It should, however, be noted that at present no connection has been demonstrated between elevated PTH serum levels in the context of hyperparathyroidism or rhPTH treatment and the occurrence of osteosarcoma in humans (19).

Although rhPTH is usually well tolerated, a few adverse effects are observed in patients, including hypercalcemia, nausea, headache, dizziness, and leg cramps (9). Both forms of rhPTH have the same adverse effects, but rhPTH 1–84 has been reported to have a lower risk of hypercalcemia (17). Despite all efforts made with rhPTH, the limited effect on nonvertebral fractures, the costs, the inconvenient route of administration, the activation of bone resorption, and the loss of efficacy with time suggest that rhPTH, although the best anabolic option today, will ultimately only partially meet the medical needs. Reducing the impact of some of these limitations constitutes the basis for current attempts to develop small molecules affecting the secretion of endogenous PTH and to use different routes of rhPTH administration.

New Approaches to PTH

Novel formulations

Transdermal application of rhPTH

One attractive option for the alternative delivery of rhPTH is transdermal self-administration using coated microneedle patches (20). In a randomized study, 165 postmenopausal women at a mean age of 64 yr were treated daily with either 20 μg rhPTH 1–34 sc or with a patch (Zosano Pharma) for the transdermal application of 20, 30, or 40 μg rhPTH 1–34 or placebo control over a period of 6 months. Interestingly, whereas the transdermal application of 40 μg rhPTH 1–34 increased the BMD in the lumbar spine to the same extent as the sc administered rhPTH 1–34, the gain of total hip BMD was much greater (20), suggesting that the pharmacokinetics of this formulation may be more beneficial to cortical bone than the daily injections.

Oral and inhaled delivery of rhPTH

A phase I randomized, placebo-controlled study reported that PTH serum concentrations rose in a dose-dependent manner upon oral administration of rhPTH 1–34 (1, 2.5, 5, and 10 mg) formulated with 100 or 200 mg of the absorption enhancer 5-CNAC [N-(5-chlorosalicyloyl)-8-aminocaprylic acid] and similarly to sc administered rhPTH. The pulmonary route of delivery is another option currently being explored (21). A phase I clinical trial has recently been performed to compare this mode of administration to sc rhPTH (www.clinicaltrials.gov).

Using an alternative protein: PTHrP

PTHrP is highly homologous to PTH in its first 36 amino acids, binds to and activates PTH1R, and increases bone mass to an extent similar to rhPTH in rats and in humans, improving mechanical strength of the spine, femur, and tibia in rats (22). In a double-blind, placebo-controlled, randomized clinical pilot study, 16 women between the ages of 50 and 75 yr were tested for the bone anabolic effect of PTHrP. Half of the study subjects were given placebo, whereas the others received approximately 400 μg PTHrP 1–36 sc daily. No adverse effects were observed and the BMD in the lumbar spine increased by 4.7% in response to the treatment (23).

A more recent study aimed to define the therapeutic window and the dose-limiting toxicities of PTHrP and to determine whether PTHrP acts as a pure anabolic agent (24). The study included 41 healthy postmenopausal women between the ages of 45 and 75 yr that were given either placebo or increasing doses of PTHrP 1–36. As reported previously, PTHrP did not cause severe adverse effects, despite a mild hypercalcemia in some subjects that were given the maximal tolerable dose of 750 μg/d. However, this was thought to be an indirect effect due to activation of 1,25-dihydroxyvitamin D production with a resulting increase in intestinal Ca2+ absorption rather than a direct effect of PTHrP on the bone-resorbing osteoclasts. Unlike rhPTH, PTHrP appeared to act as a pure bone anabolic agent without concomitant activation of bone resorption because no changes in the bone resorption markers C-telopeptide of type I collagen and N-telopeptide of type I collagen were found, but the markers of bone formation osteocalcin and P1NP were increased (24). This suggests distinct mechanisms of action for PTHrP and PTH, possibly related to differences in the on-off kinetics of the ligands on their common receptor, affecting different aspects of its downstream signaling (25). Although the effect of PTHrP on BMD was not investigated in this small study, these and previous data suggest that PTHrP might be a promising pure anabolic agent for the treatment of osteoporosis.

Based on a successful phase II, showing that sc injection of the PTHrP analog BA058 for 1 yr increased BMD at critical sites such as the spine and hip faster and greater than rhPTH, and with a reduced risk of hypercalcemia, BA058 is entering phase III. In addition, a patch for the transdermal delivery of BA058 using a microstructured transdermal system microneedle technology is currently being developed. This product is currently in phase I (www.radiuspharm.com). Whether the chronic use of PTHrP will, like rhPTH, lead to the same efficacy and/or safety limitations remains to be determined.

Increasing Endogenous PTH Secretion with Calcilytics

The costs and modes of administration of large peptides such as rhPTH and PTHrP have motivated a search for alternative approaches to the manipulation of this anabolic pathway with small molecules. Because the search for small molecule agonist of the PTH1R has so far proven unsuccessful, targeting the CaSR offers a possible alternative.

Cells of the parathyroid glands synthesize and store PTH and express on their surface the G protein-coupled seven-transmembrane-spanning CaSR, which inhibits PTH secretion when Ca2+ is bound. A low concentration of Ca2+ in the blood decreases the activity of the CaSR, thereby stimulating parathyroid glands to release PTH into the blood stream. The resulting rise in Ca2+ activates the CaSR and terminates the secretion of PTH by the parathyroid gland cells (26). This system is amenable to pharmacological manipulation using allosteric modulators of the CaSR, i.e. calcimimetics to treat hypercalcemia and secondary hyperparathyroidism or calcilytics to induce PTH secretion (27). Calcilytics mimic the state of hypocalcemia and, if short-acting, provoke a transient burst of endogenous PTH independently of extracellular Ca2+. In addition to a rapid onset of action, calcilytics should have a sufficiently high clearance and a low volume of distribution, allowing a timely clearance of the drug to mimic the pharmacokinetics of injected PTH and thereby reduce the risk of severe adverse effects such as hypercalcemia and the onset of bone resorption. If administered orally, calcilytics might be more convenient than rhPTH and therefore a possible alternative to the prevailing treatment (28).

Daily delivery of NPS 2143, the first calcilytic, to ovariectomized rats for 5 wk caused a prolonged increase in PTH plasma concentration for at least 4 h due to its long half-life. This provoked a strong increase in bone turnover but no rise in BMD. Based on the concomitantly activated bone formation and bone resorption, a concurrent administration of the antiresorptive drug 17β-estradiol caused a net increase in bone mass by decreasing bone resorption (29, 30). Another study demonstrated the short-acting compound SB-423562 and its precursor SB-423557 to promote the transient release of endogenous PTH in rats, dogs, monkeys, and humans. Although human studies have been somewhat limited so far, a single dose of SB-423562 iv or of SB-423557 orally caused a rise of endogenous PTH with a kinetic profile similar to that of injected rhPTH (31). In ovariectomized rats, orally administered SB-423557 increased bone formation and improved bone strength. ATF936 is another calcilytic that efficiently triggered the rapid elevation of endogenous PTH in rats and dogs after a single oral dose (26). In aged rats, the daily oral administration of ATF936 for 8 wk increased BMD in the proximal tibia, including trabecular- and cortical thickness. In humans, ATF936 was also well tolerated and caused a transient increase in PTH levels similar to the profile seen after rhPTH administration (26).

The first randomized, placebo-controlled, multicenter, dose-ranging trial reporting the use of a calcilytic tested Ronacaleret, a molecule similar to NPS 2143 but with reduced off-target effects and half-life. The trial included 569 postmenopausal women with low BMD. Although the markers of bone turnover increased with both rhPTH and Ronacaleret and decreased in subjects treated with alendronate after 12 month, the Ronacaleret-induced rise in serum PTH was prolonged compared with that found in subjects treated with rhPTH. In turn, the gain in lumbar spine BMD after Ronacaleret treatment was significantly below the increase seen with rhPTH or alendronate. Moreover, total hip BMD was even slightly decreased after 12 months of treatment with Ronacaleret, compared with the increase seen with rhPTH or alendronate (32). These disappointing results have led to the interruption of the clinical development of this compound for osteoporosis treatment. JTT-305/MK-5442 is another calcilytic that increased lumbar BMD in a placebo-controlled trial. However, more studies are needed to determine its clinical usefulness (33).

Besides the need for an appropriate pharmacokinetic profile, additional challenges exist with the use of calcilytics. First, the therapeutic window between the desired effects on bone and the unwanted hypercalcemia is quite small. Second, off-target effects limit the usefulness of calcilytics because CaSR are also expressed in other organs besides the parathyroid glands, including kidney, brain, and endothelial cells. Third, the secretory vesicles of the parathyroid glands that are induced to release their content upon inactivation of the CaSR by calcilytics contain not only PTH but also other products, such as chromogranin, which may negatively affect PTH secretion (34) and have other adverse effects (35). Lastly, in addition to the indirect effects of calcilytics on bone homeostasis via the CaSR in the parathyroid gland, calcilytics and/or calcimimetics may also directly affect bone cells because CaSR are expressed by osteoblasts and osteoclasts and might regulate cell recruitment, differentiation, and survival (36). Thus, although the results of ongoing clinical trials are still pending, the use of calcilytics to activate the PTH pathway to induce bone anabolism has so far not been convincing.

Modulating the Canonical Wnt-Signaling Pathway

About a decade ago, the identification of human mutations in the low-density lipoprotein receptor-related protein 5 (Lrp5) linked to low bone mass (osteoporosis pseudoglioma syndrome) or high bone mass (HBM) (3739) called the attention of the field to the Wnt pathway as a strong regulator of bone density and a potential alternative target to the PTH signaling pathway. At about the same time, the mutations causing HBM in sclerosteosis and van Buchem syndrome were identified and shown to affect the expression of another component of the Wnt signaling pathway, the Wnt antagonist sclerostin (40, 41). Interestingly, the single point HBM mutation initially identified in lrp5 (G171V) was shown to decrease the affinity of Lrp5 for the inhibitors sclerostin (42) and dickkopf1 (Dkk1) (39). Thus, these rare human genetic mutations demonstrated that Wnt signaling is a dominant regulator of bone density in humans. This was then confirmed in mouse genetic studies and became a major focus in our field for the discovery of new bone anabolic therapeutics. More recently, large genome-wide association studies identified lrp5 as one of the highly significant genes associated with BMD (43, 44), confirming in large populations the observations made in the rare genetic conditions described above, and validating the Wnt signaling pathway as a major regulator of bone mass.

Canonical Wnt activation occurs upon simultaneous binding of the secreted Wnt glycoproteins to one of the seven-helix-receptors of the frizzled family and the coreceptors Lrp5 or Lrp6. Although partially redundant and widely expressed, Lrp5 and Lrp6 are also expressed in cells of the osteoblast lineage (37, 45), and Lrp5 may be preferentially expressed and active in osteocytes (46, 47). In the absence of Wnt, intracellular Axin forms the destruction complex together with adenomatous polyposis coli, β-catenin, glycogen synthase kinase 3β (GSK3β), casein kinase 1a, and protein phosphatase 2A, leading to phosphorylation and subsequent degradation of β-catenin. Binding of Wnt to Lrp5/6 causes a conformational change of the cytoplasmic receptor domain, followed by the recruitment of Axin2 and the disassembly of the destruction complex. This prevents the phosphorylation of β-catenin by GSK3β and the proteasomal degradation of β-catenin, which accumulates in the cytosol and translocates into the nucleus, stimulating the expression of Wnt target genes including several osteoblast marker genes and osteoprotegerin (OPG), an osteoblast-derived inhibitor of osteoclast differentiation (4850), thus potentially activating bone formation and decreasing bone resorption at the same time (Figs. 2 and 3).

Fig. 3.

Fig. 3.

Signaling and cross talk of the PTH and Wnt signaling pathways in osteocytes. In the osteocyte (and late osteoblasts), activation of the canonical Wnt signaling pathway occurs upon simultaneous binding of the secreted glycoprotein Wnt3a (or other Wnts like Wnt 10b for instance) to the seven-helix-receptor frizzled (Fz) family and the coreceptors Lrp 5/6. Binding of Wnt3a to Lrp5/6 changes the conformation of the cytoplasmic receptor domain, causing the recruitment of Axin2 and preventing the phosphorylation of β-catenin by GSK3β and its proteasomal degradation. β-Catenin accumulates in the cytosol and translocates into the nucleus, thereby stimulating the expression of the Lrp5/6 antagonists Dkk1 and sclerostin, and the RANKL inhibitor OPG, via the T-cell factor/lymphoid enhancer factor (Tcf/Lef). PTH binds to its seven-transmembrane-spanning receptor and activates phosphatidyl inositol-specific phospholipase C (PLC), cAMP-dependent protein kinase A (PKA), and the protein kinase C (PKC) downstream signaling cascades, all contributing to the bone anabolic effect of PTH. In addition, PTH signaling cross talks with the Wnt signaling pathway by associating with Lrp6, inhibiting GSK3β, stabilizing β-catenin, and inhibiting the expression of both sclerostin and Dkk1.

The greatest therapeutic opportunity offered by the Wnt signaling pathway is the fact that it is under negative control by endogenous secreted factors like sclerostin and proteins of the Dkk family. Both bind to Lrp5/6 and interfere with Wnt binding (Fig. 3). In addition, Dkk1 interacts with Lrp5/6 and the Kremen receptors and initiate internalization of the receptor ligand complex (48). Other endogenous antagonists such as soluble frizzled receptor proteins (sFrp) or Wnt inhibitory factors bind the Wnt ligands, thereby decreasing not only canonical Wnt signaling, like sclerostin and Dkk1, but also its noncanonical signaling effects, the role of which is still poorly defined in bone.

Inhibiting the endogenous inhibitors

Although attractive, targeting Lrp5 or frizzled-receptors with small agonists or targeting intracellular components of the canonical Wnt pathway with small molecules could be challenging for bone anabolic therapies. Besides the molecular challenges, the lack of bone specificity has somewhat limited the enthusiasm for these approaches. In contrast, and possibly explaining the lack of extraskeletal side effects in patients with aberrant sclerostin expression, which has been shown to be exquisitely, albeit not entirely, restricted to late osteoblasts and osteocytes (51), pointing to its suitability as a therapeutic target of choice. Furthermore, any agent that would antagonize sclerostin would be active not only almost exclusively in bone but also only in areas of the skeleton where this inhibitor of bone formation is produced, possibly targeting therapy to specific areas of bone.

Sclerostin antibodies

Mouse genetics demonstrated that targeted deletion of sclerostin leads to HBM, due to a massive increase in BFR affecting not only trabecular but also cortical bone (52). In female rats, antibody-based sclerostin inhibition not only increased bone mass and strength in healthy animals but also reversed ovariectomy-induced bone loss (53). Furthermore, injection of sclerostin antibodies in aged male rats caused an increase in bone formation, bone mass, and strength of the long bones and the lumbar spine (54). In a model of hindlimb immobilization, antibody-mediated blockade of sclerostin in adult female rats resulted in a rapid increase in cortical and trabecular bone mass in both ambulated and immobilized bones. This effect was dominated by high bone formation and a decrease in bone resorption, suggesting that inhibition of sclerostin might be useful for the treatment of immobilization-induced osteopenia (55). In a model of bone healing, sclerostin-neutralizing antibodies increased the amount of bone and mechanical strength (56). Furthermore, sclerostin antibodies significantly improved the healing of fractures in rats and of osteotomies in monkeys accompanied with improvements in bone formation, bone mass, and bone strength at nonfractured cortical and trabecular sites in both models (57). Consistent with these data, injection of humanized sclerostin-neutralizing antibodies once a month for 2 months in gonad-intact female monkeys had a marked dose-dependent effect on bone formation (51). This increase was predominantly due to direct stimulation of bone formation along resting surfaces (modeling effect) with no increase in bone resorption but a stimulation of remodeling-based bone formation (58). Similar findings were made in sclerostin knockout mice and in ovariectomized rats treated with sclerostin antibodies (53, 54). Thus, the use of antagonists to endogenous inhibitors of the Wnt pathway seems to stimulate bone formation directly, through bone modeling, i.e. at least in part independent of bone remodeling and activation frequency and therefore without prior activation of bone resorption. This mode of action may open interesting therapeutic applications not only in osteoporosis but also in bone repair and in low turnover conditions where increasing bone mass is desired.

Recently, Padhi et al. (59) reported the first human phase I randomized, double-blind, placebo-controlled clinical trial testing ascending single doses of AMG 785, a humanized monoclonal sclerostin antibody, in healthy men and postmenopausal women. Bone formation markers increased within 1 month after a single sc dose of 10 mg/kg AMG 785 to levels similar to daily injections of rhPTH for 6 months, and markers of bone resorption decreased. Although this antiresorptive effect was expected due to the known Wnt-mediated increase in OPG expression in mice and due to the preclinical data, it remained a surprising finding in humans, particularly in its amplitude. Likewise, the gain in BMD at the lumbar spine and total hip was comparable or even greater than with rhPTH (59).

More recently Amgen/UCB reported in a press release (www.amgen.com) some of the results from the phase II study comparing the sclerostin-antibody AMG 785/CDP7851 to placebo in approximately 400 postmenopausal women with low BMD for the treatment of postmenopausal osteoporosis. AMG 785/CDP7851 was given at 70, 140, and 210 mg sc once a month and at 140 and 210 mg every 3 months. At 12 months, AMG 785/CDP7851 significantly increased BMD in the lumbar spine compared with placebo and teriparatide. Injection site reactions were the most frequently reported adverse events. These studies point to the promising future of sclerostin antibodies for the treatment of low bone mass diseases.

Dkk1 antagonists

Dkk1 is also an endogenous inhibitor of Wnt signaling, and the HBM mutations in Lrp5 negatively affect not only sclerostin but also Dkk1 affinity (39). Confirmation that Dkk1 plays a critical role in the regulation of bone mass was provided using mice in which Dkk1 was overexpressed or deleted, displaying severe loss (60) or gain (61) of bone mass, respectively. Ovariectomy-induced loss of bone and of mechanical strength were abrogated by Dkk1 antisense oligonucleotides (62), and inhibiting Dkk1 expression retards glucocorticoid-induced osteopenia (63). Antibody-mediated Dkk1 neutralization also protected against inflammatory bone loss in mice overexpressing TNF by preventing TNF-mediated impaired bone formation, osteocyte death, and enhanced osteoclast activity (64). In addition, Dkk1 inhibition reversed rheumatoid arthritis-associated bone destruction into a bone-forming osteoarthritis, with the formation of osteophytes, attributed to an enhanced bone formation when Wnt signaling is stimulated locally (65). In multiple myeloma, Dkk1 serum level correlated with focal bone lesions, and its inhibition increased osteoblast number and cancellous bone mass (66).

Based on this information, and although sclerostin antibodies are probably the preferred and most advanced therapeutic option for osteoporosis, antibodies to Dkk1 are also being developed, in particular for the treatment of multiple myeloma. If proven safe and efficacious, these antibodies could also find their way to a more general indication in low bone mass diseases, although the possibility that Dkk1 is less restricted to the bone microenvironment than sclerostin may raise more concerns about off-target effects.

Other potential approaches to enhance Wnt signaling

Neutralizing sFrps

Secreted frizzled-related proteins are a group of physiological antagonists of Wnt signaling. Effective inhibition of these Wnt antagonists causes a net activation of the Wnt pathway, both canonical and noncanonical, thereby potentially increasing bone mass in osteoporotic patients. This concept is supported by the presence of polymorphisms in the Sfrp1 gene that are associated with BMD and bone mineral content (67). In addition, fracture repair was accelerated in adult mice germline-deleted of Sfrp1 (68), and systemic overexpression of sFrp1 inhibited osteoblast function and the anabolic effect of rhPTH (69). Furthermore, mice overexpressing sFrp4 also exhibit a low BFR and bone mass phenotype (70). sFrps may therefore also be valid targets for enhancement of Wnt signaling, but the fact that antagonists would most likely activate both the canonical and noncanonical Wnt signaling pathways may lead to different clinical responses, both in terms of bone density and side effects.

Lithium and GSK3β inhibitors

Wnt activation by GSK3β inhibition could be useful for bone anabolic treatment, although the lack of skeletal specificity would be a challenge because it is expressed and functional in all cell types. Lithium is a nonspecific GSK3β inhibitor commonly used for the treatment of bipolar disease. Lithium activates Wnt signaling in a receptor-independent manner and increases bone mass in mice (71). Interestingly, in humans lithium decreased markers of bone formation and resorption but resulted in a net increase in bone density (72). Furthermore, lithium lowered the risk of fracture (73, 74), and the fracture risk increased after lithium withdrawal (75). It is not clear, however, whether this can be attributed to the direct effect of lithium on bone or to its effect on the mental disorder for which it was primarily given (75). Lithium also has a complex influence on Ca2+ homeostasis and exerts nonskeletal adverse effects. Thus, the potential of lithium as a bone anabolic agent may be limited, and it is not clear whether it is currently considered as a therapeutic option for osteoporosis treatment.

GSK3β inactivation can also be achieved using other, more specific small molecule inhibitors such as 603281-31-8, which increased bone formation markers, femoral bone mass, and bone strength in ovariectomized rats (76). Another study demonstrated the abrogation of glucocorticoid-induced bone loss by the GSK3β inhibitor 6-bromoindirubin-3′-oxim (77). Furthermore, the novel GSK3β inhibitor AR28 induced β-catenin nuclear translocation and increased bone mass after 2 wk, possibly due to an amplification of mesenchymal stem cells that became osteoblasts at the expense of adipocytes (78).

Potential concerns

Although activation of the Wnt signaling pathway is a very promising approach for the development of bone anabolic drugs, safety concerns exist, in particular regarding possible oncogenic effects and uncontrolled formation of bone leading to increased intracranial pressure, osteophyte formation, and/or closure of skeletal foramen, affecting hearing or vision for instance. The oncogenic concerns are based upon the fact that there is abundant literature linking aberrant activation of Wnt signaling to a variety of tumors, in particular colon cancer (79). Furthermore, recent studies have indicated that deletion of some of the endogenous inhibitors of Wnt signaling, such as Wnt inhibitory factor 1, can lead to the development of osteosarcoma (80). It is, however, reassuring that so far no increase in tumors has been reported in sost, dkk1, or Sfrp knockout mice. Similarly, the pharmaceutical companies that have been testing sclerostin or Dkk1 antibodies in clinical trials have performed extensive preclinical safety studies, which did not show any obvious oncogenicity. Finally, after long-term treatment with LiCl and HBM (sclerosteosis or van Buchem's disease), patients have not been noted to have an increased susceptibility to cancer (3739, 41). Oncogenicity will nevertheless have to be one of the major adverse effects that will need to be closely monitored in clinical trials and in postmarketing studies.

If uncontrolled bone formation leads to adverse effects in homozygous sclerosteosis or van Buchem patients (intracranial pressure, loss of hearing or vision), it is most reassuring that heterozygous carriers of these mutations do not demonstrate any adverse effects but exhibit significant increases in bone density (81). It is therefore important to mitigate these potential concerns with the fact that therapeutic intervention will not eliminate entirely the endogenous inhibitor and will occur only over a limited period of time.

The PTH and Wnt Pathways Converge on the Same Anabolic Mechanism

In the last few years, it has become apparent that Wnt signaling contributes significantly to the anabolic effects of PTH, raising the question whether we are really activating two different osteoanabolic pathways or whether these two pathways converge to eventually become only one. Indeed, it has been reported that PTH inactivates GSK3β (82) and stabilizes β-catenin (83). In addition, PTH1R and Lrp6 can form a complex when PTH binds to its receptor, leading to the disassembly of the destruction complex (84). PTH also targets the osteocytes and reduces the expression of the Wnt inhibitors sclerostin (47) and Dkk1 (Fig. 3) (85). This local decrease in endogenous inhibitors leads to enhancement of an autocrine Wnt signaling loop in these cells, and deletion of lrp5 in mice blunts the osteoanabolic effect of PTH (47).

Thus, whereas other aspects of PTH signaling may also contribute to its effects on activation frequency and anabolic responses, the role of Wnt signaling downstream of the PTH1R has clearly emerged as a very significant contributor to the observed effects of PTH on bone formation.

Antagonizing activin: another approach for bone anabolic therapy

Activin elicits its signaling by binding to two type I receptors (ActRIA and -IB) and two type II receptors (ActRIIA and -IIB). Preclinical studies have shown that activin acts as an antagonist to human osteoblast differentiation (86) and as an agonist to osteoclast formation and bone resorption (87). These findings indicated that blocking the endogenous effects of activin could favor bone formation and block bone resorption. ACE-011, a soluble form of the extracellular domain of the ActRIIA fused to the Fc domain of murine IgG (ActRIIA-IgG1-Fc), acts as an activin decoy receptor and causes bone anabolic effects in healthy and ovariectomized mice (88). Furthermore, sc injection of ACE-011 for 3 months in monkeys promoted bone formation and inhibited bone resorption, thereby acting as a dual anabolic-antiresorptive compound (89). In a phase I trial, ACE-011 was well tolerated and increased markers of bone formation (90). Thus, in principle this soluble form of the ActRIIa receptor could be developed for the anabolic treatment of osteoporosis, provided its safety profile is acceptable, especially because it increases the hematocrit (91), and its efficacy is comparable to rhPTH or to antagonists of Wnt signaling inhibitors.

Can novel antiresorptives reveal an “anabolic” component?

During bone remodeling, osteoclasts and/or the resorption they exert on bone matrix have the capacity to activate osteoblasts and bone formation, a process called coupling. This process ensures the succession of bone formation to bone resorption in the remodeling cycle within each BMU. In the last few years, a few coupling factors have been identified, and analysis of osteopetrotic mutants has illustrated the fact that it is possible to reduce the activity of osteoclasts without automatically reducing bone formation. Indeed, osteopetrotic conditions in animals and humans revealed that an impaired osteoclast activity due to mutations in or deletions of chloride channels, components of the vacuolar ATPase, or of cathepsin K, decreases bone resorption with a paradoxical increase in the number of osteoclasts and activation frequency and with a maintained or even increased BFR (92, 93).

This has led to the search for compounds that could indeed mimic such conditions. Such antiresorptive compounds could in principle meet the criteria for being classified as bone anabolics, provided they increase bone formation markers (P1NP and bone-specific alkaline phosphatase), MAR, and/or BFR while reducing markers of bone resorption (C-telopeptide of type I collagen, N-telopeptide of type I collagen). Although not meeting these criteria in humans, pharmacological inhibition of cathepsin K in animals has been reported to reduce bone resorption markers to an extent comparable to oral bisphosphonates (approximately 40%) while increasing bone formation markers in rabbits and monkeys (94, 95). In human clinical trials, bone formation markers returned to near baseline (96) or were unchanged (97, 98). Thus, although not yet fully supported by current human studies, the possibility exists that novel antiresorptives may also fulfill the criteria of bone anabolics. Furthermore, ongoing research on the mechanisms regulating the cross talk between osteoclasts and osteoblasts has identified some bone formation-stimulating osteoclast-derived factors (“clastokines”) and matrix-derived growth factors (99, 100), which may in the future help to design novel osteoanabolic compounds.

Conclusion and Perspectives

The future of osteoporosis therapy is full of exciting prospects. In addition to the extraordinary improvement in our ability to reduce the incidence of fractures since the introduction of the first bisphosphonates on the market, new and more potent bisphosphonates have allowed the use of less frequent administrations and safer routes. Moreover, the identification of RANKL as an essential cytokine in osteoclast differentiation and the development of efficient antibodies to block its action have further improved our ability to counter the devastating effects of uncontrolled bone resorption. If both bisphosphonates and RANKL antibodies also decrease bone formation, new antiresorptives may succeed at avoiding this antianabolic action, thereby possibly improving the fracture outcomes.

Yet, the identification of agents that can stimulate bone formation has been recognized as a priority in our field to treat severe osteoporosis cases and to potentially improve upon the limited efficacy of antiresorptive drugs on nonvertebral fractures. rhPTH has definitely proven its ability to increase bone formation and significantly increase bone density in severe patients, but its ability to reduce nonvertebral fractures has also been modest and its effects are limited in time, possibly because its mode of action is mostly based on bone remodeling. In this context, the discovery of Wnt signaling as a major osteoanabolic pathway, which seems to exert its effects mostly through a bone remodeling-independent mechanism (modeling-based) opens tremendous possibilities to improve bone density not only in trabecular bone but also in cortical bone, which should reduce the incidence of nonvertebral fractures.

Several clinical trials for osteoporosis treatment are ongoing, testing new antiresorptives, different forms of rhPTH, or agents that activate Wnt signaling. In many of these trials combinations and sequences of these agents with various antiresorptives are also being tested. The next few years will therefore be very exciting for osteoporosis treatment.

Acknowledgments

We are grateful to Sundeep Khosla for critical reading of the manuscript.

Roland Baron acknowledges the support of National Institutes of Health (National Institute of Arthritis and Musculoskeletal and Skin Diseases and National Institute of Dental and Craniofacial Research).

Disclosure Summary: R.B. has served as consultant for Merck, Novartis, Amgen, Takeda, Chugai, Teijin, Japan Tobacco, Ono, Servier, and GSK and is on the scientific advisory boards of Bone Therapeutics and Arcarios. E.H. has nothing to disclose.

Footnotes

Abbreviations:
ActR
Activin receptor
BFR
bone formation rate
BMD
bone mineral density
BMU
basic multicellular unit
CaSR
calcium-sensing receptor
Dkk1
Dickkopf homolog 1
GSK3β
glycogen synthase kinase 3β
HBM
high bone mass
Lrp
low-density lipoprotein receptor-related protein
MAR
mineral apposition rate
OPG
osteoprotegerin
PLC
phospholipase C
P1NP
procollagen type 1 amino-terminal propeptide
PTH1R
PTH receptor
RANKL
receptor activator of nuclear factor-κB ligand
rhPTH
recombinant human PTH
sFrp
secreted frizzled-related protein
Wnt
wingless-int.

References

  • 1. Parfitt AM, Drezner MK, Glorieux FH, Kanis JA, Malluche H, Meunier PJ, Ott SM, Recker RR. 1987. Bone histomorphometry: standardization of nomenclature, symbols, and units. Report of the ASBMR Histomorphometry Nomenclature Committee. J Bone Miner Res 2:595–610 [DOI] [PubMed] [Google Scholar]
  • 2. Bonewald LF. 2011. The amazing osteocyte. J Bone Miner Res 26:229–238 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 3. Harvey N, Dennison E, Cooper C. 2010. Osteoporosis: impact on health and economics. Nat Rev Rheumatol 6:99–105 [DOI] [PubMed] [Google Scholar]
  • 4. Hollick RJ, Reid DM. 2011. Role of bisphosphonates in the management of postmenopausal osteoporosis: an update on recent safety anxieties. Menopause Int 17:66–72 [DOI] [PubMed] [Google Scholar]
  • 5. Spiro AS, Beil FT, Timo Beil F, Schinke T, Schilling AF, Eulenburg C, Rueger JM, Amling M. 2010. Short-term application of dexamethasone enhances bone morphogenetic protein-7-induced ectopic bone formation in vivo. J Trauma 69:1473–1480 [DOI] [PubMed] [Google Scholar]
  • 6. Yakar S, Canalis E, Sun H, Mejia W, Kawashima Y, Nasser P, Courtland HW, Williams V, Bouxsein M, Rosen C, Jepsen KJ. 2009. Serum IGF-1 determines skeletal strength by regulating subperiosteal expansion and trait interactions. J Bone Miner Res 24:1481–1492 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 7. Guo J, Liu M, Yang D, Bouxsein ML, Thomas CC, Schipani E, Bringhurst FR, Kronenberg HM. 2010. Phospholipase C signaling via the parathyroid hormone (PTH)/PTH-related peptide receptor is essential for normal bone responses to PTH. Endocrinology 151:3502–3513 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 8. Bellido T, Ali AA, Plotkin LI, Fu Q, Gubrij I, Roberson PK, Weinstein RS, O'Brien CA, Manolagas SC, Jilka RL. 2003. Proteasomal degradation of Runx2 shortens parathyroid hormone-induced anti-apoptotic signaling in osteoblasts. A putative explanation for why intermittent administration is needed for bone anabolism. J Biol Chem 278:50259–50272 [DOI] [PubMed] [Google Scholar]
  • 9. Neer RM, Arnaud CD, Zanchetta JR, Prince R, Gaich GA, Reginster JY, Hodsman AB, Eriksen EF, Ish-Shalom S, Genant HK, Wang O, Mitlak BH. 2001. Effect of parathyroid hormone (1–34) on fractures and bone mineral density in postmenopausal women with osteoporosis. N Engl J Med 344:1434–1441 [DOI] [PubMed] [Google Scholar]
  • 10. Girotra M, Rubin MR, Bilezikian JP. 2006. The use of parathyroid hormone in the treatment of osteoporosis. Rev Endocr Metab Disord 7:113–121 [DOI] [PubMed] [Google Scholar]
  • 11. Ma YL, Zeng Q, Donley DW, Ste-Marie LG, Gallagher JC, Dalsky GP, Marcus R, Eriksen EF. 2006. Teriparatide increases bone formation in modeling and remodeling osteons and enhances IGF-II immunoreactivity in postmenopausal women with osteoporosis. J Bone Miner Res 21:855–864 [DOI] [PubMed] [Google Scholar]
  • 12. Jilka RL. 2007. Molecular and cellular mechanisms of the anabolic effect of intermittent PTH. Bone 40:1434–1446 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 13. Drake MT, Srinivasan B, Mödder UI, Ng AC, Undale AH, Roforth MM, Peterson JM, McCready LK, Riggs BL, Khosla S. 2011. Effects of intermittent parathyroid hormone treatment on osteoprogenitor cells in postmenopausal women. Bone 49:349–355 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 14. Black DM, Greenspan SL, Ensrud KE, Palermo L, McGowan JA, Lang TF, Garnero P, Bouxsein ML, Bilezikian JP, Rosen CJ. 2003. The effects of parathyroid hormone and alendronate alone or in combination in postmenopausal osteoporosis. N Engl J Med 349:1207–1215 [DOI] [PubMed] [Google Scholar]
  • 15. Finkelstein JS, Wyland JJ, Lee H, Neer RM. 2010. Effects of teriparatide, alendronate, or both in women with postmenopausal osteoporosis. J Clin Endocrinol Metab 95:1838–1845 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 16. Cosman F, Eriksen EF, Recknor C, Miller PD, Guañabens N, Kasperk C, Papanastasiou P, Readie A, Rao H, Gasser JA, Bucci-Rechtweg C, Boonen S. 2011. Effects of intravenous zoledronic acid plus subcutaneous teriparatide [rhPTH(1–34)] in postmenopausal osteoporosis. J Bone Miner Res 26:503–511 [DOI] [PubMed] [Google Scholar]
  • 17. Black DM, Bilezikian JP, Ensrud KE, Greenspan SL, Palermo L, Hue T, Lang TF, McGowan JA, Rosen CJ. 2005. One year of alendronate after one year of parathyroid hormone (1–84) for osteoporosis. N Engl J Med 353:555–565 [DOI] [PubMed] [Google Scholar]
  • 18. Vahle JL, Sato M, Long GG, Young JK, Francis PC, Engelhardt JA, Westmore MS, Linda Y, Nold JB. 2002. Skeletal changes in rats given daily subcutaneous injections of recombinant human parathyroid hormone (1–34) for 2 years and relevance to human safety. Toxicol Pathol 30:312–321 [DOI] [PubMed] [Google Scholar]
  • 19. Subbiah V, Madsen VS, Raymond AK, Benjamin RS, Ludwig JA. 2010. Of mice and men: divergent risks of teriparatide-induced osteosarcoma. Osteoporos Int 21:1041–1045 [DOI] [PubMed] [Google Scholar]
  • 20. Cosman F, Lane NE, Bolognese MA, Zanchetta JR, Garcia-Hernandez PA, Sees K, Matriano JA, Gaumer K, Daddona PE. 2010. Effect of transdermal teriparatide administration on bone mineral density in postmenopausal women. J Clin Endocrinol Metab 95:151–158 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 21. Shoyele SA, Sivadas N, Cryan SA. 2011. The effects of excipients and particle engineering on the biophysical stability and aerosol performance of parathyroid hormone (1–34) prepared as a dry powder for inhalation. AAPS PharmSciTech 12:304–311 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 22. Stewart AF, Cain RL, Burr DB, Jacob D, Turner CH, Hock JM. 2000. Six-month daily administration of parathyroid hormone and parathyroid hormone-related protein peptides to adult ovariectomized rats markedly enhances bone mass and biomechanical properties: a comparison of human parathyroid hormone 1–34, parathyroid hormone-related protein 1–36, and SDZ-parathyroid hormone 893. J Bone Miner Res 15:1517–1525 [DOI] [PubMed] [Google Scholar]
  • 23. Horwitz MJ, Tedesco MB, Gundberg C, Garcia-Ocana A, Stewart AF. 2003. Short-term, high-dose parathyroid hormone-related protein as a skeletal anabolic agent for the treatment of postmenopausal osteoporosis. J Clin Endocrinol Metab 88:569–575 [DOI] [PubMed] [Google Scholar]
  • 24. Horwitz MJ, Tedesco MB, Garcia-Ocaña A, Sereika SM, Prebehala L, Bisello A, Hollis BW, Gundberg CM, Stewart AF. 2010. Parathyroid hormone-related protein for the treatment of postmenopausal osteoporosis: defining the maximal tolerable dose. J Clin Endocrinol Metab 95:1279–1287 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 25. Dean T, Vilardaga JP, Potts JT, Jr, Gardella TJ. 2008. Altered selectivity of parathyroid hormone (PTH) and PTH-related protein (PTHrP) for distinct conformations of the PTH/PTHrP receptor. Mol Endocrinol 22:156–166 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 26. John MR, Widler L, Gamse R, Buhl T, Seuwen K, Breitenstein W, Bruin GJ, Belleli R, Klickstein LB, Kneissel M. 2011. ATF936, a novel oral calcilytic, increases bone mineral density in rats and transiently releases parathyroid hormone in humans. Bone 49:233–241 [DOI] [PubMed] [Google Scholar]
  • 27. Saidak Z, Brazier M, Kamel S, Mentaverri R. 2009. Agonists and allosteric modulators of the calcium-sensing receptor and their therapeutic applications. Mol Pharmacol 76:1131–1144 [DOI] [PubMed] [Google Scholar]
  • 28. Widler L, Altmann E, Beerli R, Breitenstein W, Bouhelal R, Buhl T, Gamse R, Gerspacher M, Halleux C, John MR, Lehmann H, Kalb O, Kneissel M, Missbach M, Müller IR, Reidemeister S, Renaud J, Taillardat A, Tommasi R, Weiler S, Wolf RM, Seuwen K. 2010. 1-Alkyl-4-phenyl-6-alkoxy-1H-quinazolin-2-ones: a novel series of potent calcium-sensing receptor antagonists. J Med Chem 53:2250–2263 [DOI] [PubMed] [Google Scholar]
  • 29. Gowen M, Stroup GB, Dodds RA, James IE, Votta BJ, Smith BR, Bhatnagar PK, Lago AM, Callahan JF, DelMar EG, Miller MA, Nemeth EF, Fox J. 2000. Antagonizing the parathyroid calcium receptor stimulates parathyroid hormone secretion and bone formation in osteopenic rats. J Clin Invest 105:1595–1604 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 30. Nemeth EF. 2002. The search for calcium receptor antagonists (calcilytics). J Mol Endocrinol 29:15–21 [DOI] [PubMed] [Google Scholar]
  • 31. Kumar S, Matheny CJ, Hoffman SJ, Marquis RW, Schultz M, Liang X, Vasko JA, Stroup GB, Vaden VR, Haley H, Fox J, DelMar EG, Nemeth EF, Lago AM, Callahan JF, Bhatnagar P, Huffman WF, Gowen M, Yi B, Danoff TM, Fitzpatrick LA. 2010. An orally active calcium-sensing receptor antagonist that transiently increases plasma concentrations of PTH and stimulates bone formation. Bone 46:534–542 [DOI] [PubMed] [Google Scholar]
  • 32. Fitzpatrick LA, Dabrowski CE, Cicconetti G, Gordon DN, Papapoulos S, Bone HG, 3rd, Bilezikian JP. 2011. The effects of ronacaleret, a calcium-sensing receptor antagonist, on bone mineral density and biochemical markers of bone turnover in postmenopausal women with low bone mineral density. J Clin Endocrinol Metab 96:2441–2449 [DOI] [PubMed] [Google Scholar]
  • 33. Fukumoto S. 2011. [Antagonist for calcium-sensing receptor. JTT-305/MK-5442]. Clin Calcium 21:89–93 [PubMed] [Google Scholar]
  • 34. Cohn DV, Fasciotto BH, Reese BK, Zhang JX. 1995. Chromogranin A: a novel regulator of parathyroid gland secretion. J Nutr 125(7 Suppl):2015S–2019S [DOI] [PubMed] [Google Scholar]
  • 35. Di Comite G, Morganti A. 2011. Chromogranin A: a novel factor acting at the cross road between the neuroendocrine and the cardiovascular systems. J Hypertens 29:409–414 [DOI] [PubMed] [Google Scholar]
  • 36. Marie PJ. 2010. The calcium-sensing receptor in bone cells: a potential therapeutic target in osteoporosis. Bone 46:571–576 [DOI] [PubMed] [Google Scholar]
  • 37. Gong Y, Slee RB, Fukai N, Rawadi G, Roman-Roman S, Reginato AM, Wang H, Cundy T, Glorieux FH, Lev D, Zacharin M, Oexle K, Marcelino J, Suwairi W, Heeger S, Sabatakos G, Apte S, Adkins WN, Allgrove J, Arslan-Kirchner M, Batch JA, Beighton P, Black GC, Boles RG, Boon LM, Borrone C, Brunner HG, Carle GF, Dallapiccola B, De Paepe A, Floege B, Halfhide ML, Hall B, Hennekam RC, Hirose T, Jans A, Jüppner H, Kim CA, Keppler-Noreuil K, Kohlschuetter A, LaCombe D, Lambert M, Lemyre E, Letteboer T, Peltonen L, Ramesar RS, Romanengo M, Somer H, Steichen-Gersdorf E, Steinmann B, Sullivan B, Superti-Furga A, Swoboda W, van den Boogaard MJ, Van Hul W, Vikkula M, Votruba M, Zabel B, Garcia T, Baron R, Olsen BR, Warman ML. 2001. LDL receptor-related protein 5 (LRP5) affects bone accrual and eye development. Cell 107:513–523 [DOI] [PubMed] [Google Scholar]
  • 38. Little RD, Carulli JP, Del Mastro RG, Dupuis J, Osborne M, Folz C, Manning SP, Swain PM, Zhao SC, Eustace B, Lappe MM, Spitzer L, Zweier S, Braunschweiger K, Benchekroun Y, Hu X, Adair R, Chee L, FitzGerald MG, Tulig C, Caruso A, Tzellas N, Bawa A, Franklin B, McGuire S, Nogues X, Gong G, Allen KM, Anisowicz A, Morales AJ, Lomedico PT, Recker SM, Van Eerdewegh P, Recker RR, Johnson ML. 2002. A mutation in the LDL receptor-related protein 5 gene results in the autosomal dominant high-bone-mass trait. Am J Hum Genet 70:11–19 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 39. Boyden LM, Mao J, Belsky J, Mitzner L, Farhi A, Mitnick MA, Wu D, Insogna K, Lifton RP. 2002. High bone density due to a mutation in LDL-receptor-related protein 5. N Engl J Med 346:1513–1521 [DOI] [PubMed] [Google Scholar]
  • 40. van Bezooijen RL, Roelen BA, Visser A, van der Wee-Pals L, de Wilt E, Karperien M, Hamersma H, Papapoulos SE, ten Dijke P, Löwik CW. 2004. Sclerostin is an osteocyte-expressed negative regulator of bone formation, but not a classical BMP antagonist. J Exp Med 199:805–814 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 41. Loots GG, Kneissel M, Keller H, Baptist M, Chang J, Collette NM, Ovcharenko D, Plajzer-Frick I, Rubin EM. 2005. Genomic deletion of a long-range bone enhancer misregulates sclerostin in Van Buchem disease. Genome Res 15:928–935 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 42. Ellies DL, Viviano B, McCarthy J, Rey JP, Itasaki N, Saunders S, Krumlauf R. 2006. Bone density ligand, Sclerostin, directly interacts with LRP5 but not LRP5G171V to modulate Wnt activity. J Bone Miner Res 21:1738–1749 [DOI] [PubMed] [Google Scholar]
  • 43. Richards JB, Rivadeneira F, Inouye M, Pastinen TM, Soranzo N, Wilson SG, Andrew T, Falchi M, Gwilliam R, Ahmadi KR, Valdes AM, Arp P, Whittaker P, Verlaan DJ, Jhamai M, Kumanduri V, Moorhouse M, van Meurs JB, Hofman A, Pols HA, Hart D, Zhai G, Kato BS, Mullin BH, Zhang F, Deloukas P, Uitterlinden AG, Spector TD. 2008. Bone mineral density, osteoporosis, and osteoporotic fractures: a genome-wide association study. Lancet 371:1505–1512 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 44. van Meurs JB, Trikalinos TA, Ralston SH, Balcells S, Brandi ML, Brixen K, Kiel DP, Langdahl BL, Lips P, Ljunggren O, Lorenc R, Obermayer-Pietsch B, Ohlsson C, Pettersson U, Reid DM, Rousseau F, Scollen S, Van Hul W, Agueda L, Akesson K, Benevolenskaya LI, Ferrari SL, Hallmans G, Hofman A, Husted LB, Kruk M, Kaptoge S, Karasik D, Karlsson MK, Lorentzon M, Masi L, McGuigan FE, Mellström D, Mosekilde L, Nogues X, Pols HA, Reeve J, Renner W, Rivadeneira F, van Schoor NM, Weber K, Ioannidis JP, Uitterlinden AG. 2008. Large-scale analysis of association between LRP5 and LRP6 variants and osteoporosis. JAMA 299:1277–1290 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 45. Babij P, Zhao W, Small C, Kharode Y, Yaworsky PJ, Bouxsein ML, Reddy PS, Bodine PV, Robinson JA, Bhat B, Marzolf J, Moran RA, Bex F. 2003. High bone mass in mice expressing a mutant LRP5 gene. J Bone Miner Res 18:960–974 [DOI] [PubMed] [Google Scholar]
  • 46. Cui Y, Niziolek PJ, MacDonald BT, Zylstra CR, Alenina N, Robinson DR, Zhong Z, Matthes S, Jacobsen CM, Conlon RA, Brommage R, Liu Q, Mseeh F, Powell DR, Yang QM, Zambrowicz B, Gerrits H, Gossen JA, He X, Bader M, Williams BO, Warman ML, Robling AG. 2011. Lrp5 functions in bone to regulate bone mass. Nat Med 17:684–691 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 47. O'Brien CA, Plotkin LI, Galli C, Goellner JJ, Gortazar AR, Allen MR, Robling AG, Bouxsein M, Schipani E, Turner CH, Jilka RL, Weinstein RS, Manolagas SC, Bellido T. 2008. Control of bone mass and remodeling by PTH receptor signaling in osteocytes. PloS One 3:e2942. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 48. MacDonald BT, Tamai K, He X. 2009. Wnt/β-catenin signaling: components, mechanisms, and diseases. Dev Cell 17:9–26 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 49. Glass DA, 2nd, Bialek P, Ahn JD, Starbuck M, Patel MS, Clevers H, Taketo MM, Long F, McMahon AP, Lang RA, Karsenty G. 2005. Canonical Wnt signaling in differentiated osteoblasts controls osteoclast differentiation. Dev Cell 8:751–764 [DOI] [PubMed] [Google Scholar]
  • 50. Holmen SL, Zylstra CR, Mukherjee A, Sigler RE, Faugere MC, Bouxsein ML, Deng L, Clemens TL, Williams BO. 2005. Essential role of β-catenin in postnatal bone acquisition. J Biol Chem 280:21162–21168 [DOI] [PubMed] [Google Scholar]
  • 51. van Bezooijen RL, ten Dijke P, Papapoulos SE, Löwik CW. 2005. SOST/sclerostin, an osteocyte-derived negative regulator of bone formation. Cytokine Growth Factor Rev 16:319–327 [DOI] [PubMed] [Google Scholar]
  • 52. Li X, Ominsky MS, Niu QT, Sun N, Daugherty B, D'Agostin D, Kurahara C, Gao Y, Cao J, Gong J, Asuncion F, Barrero M, Warmington K, Dwyer D, Stolina M, Morony S, Sarosi I, Kostenuik PJ, Lacey DL, Simonet WS, Ke HZ, Paszty C. 2008. Targeted deletion of the sclerostin gene in mice results in increased bone formation and bone strength. J Bone Miner Res 23:860–869 [DOI] [PubMed] [Google Scholar]
  • 53. Li X, Ominsky MS, Warmington KS, Morony S, Gong J, Cao J, Gao Y, Shalhoub V, Tipton B, Haldankar R, Chen Q, Winters A, Boone T, Geng Z, Niu QT, Ke HZ, Kostenuik PJ, Simonet WS, Lacey DL, Paszty C. 2009. Sclerostin antibody treatment increases bone formation, bone mass, and bone strength in a rat model of postmenopausal osteoporosis. J Bone Miner Res 24:578–588 [DOI] [PubMed] [Google Scholar]
  • 54. Li X, Warmington KS, Niu QT, Asuncion FJ, Barrero M, Grisanti M, Dwyer D, Stouch B, Thway TM, Stolina M, Ominsky MS, Kostenuik PJ, Simonet WS, Paszty C, Ke HZ. 2010. Inhibition of sclerostin by monoclonal antibody increases bone formation, bone mass, and bone strength in aged male rats. J Bone Miner Res 25:2647–2656 [DOI] [PubMed] [Google Scholar]
  • 55. Tian X, Setterberg RB, Li X, Paszty C, Ke HZ, Jee WS. 2010. Treatment with a sclerostin antibody increases cancellous bone formation and bone mass regardless of marrow composition in adult female rats. Bone 47:529–533 [DOI] [PubMed] [Google Scholar]
  • 56. Agholme F, Li X, Isaksson H, Ke HZ, Aspenberg P. 2010. Sclerostin antibody treatment enhances metaphyseal bone healing in rats. J Bone Miner Res 25:2412–2418 [DOI] [PubMed] [Google Scholar]
  • 57. Ominsky MS, Li C, Li X, Tan HL, Lee E, Barrero M, Asuncion FJ, Dwyer D, Han CY, Vlasseros F, Samadfam R, Jolette J, Smith SY, Stolina M, Lacey DL, Simonet WS, Paszty C, Li G, Ke HZ. 2011. Inhibition of sclerostin by monoclonal antibody enhances bone healing and improves bone density and strength of nonfractured bones. J Bone Miner Res 26:1012–1021 [DOI] [PubMed] [Google Scholar]
  • 58. Ominsky MS, Vlasseros F, Jolette J, Smith SY, Stouch B, Doellgast G, Gong J, Gao Y, Cao J, Graham K, Tipton B, Cai J, Deshpande R, Zhou L, Hale MD, Lightwood DJ, Henry AJ, Popplewell AG, Moore AR, Robinson MK, Lacey DL, Simonet WS, Paszty C. 2010. Two doses of sclerostin antibody in cynomolgus monkeys increases bone formation, bone mineral density, and bone strength. J Bone Miner Res 25:948–959 [DOI] [PubMed] [Google Scholar]
  • 59. Padhi D, Jang G, Stouch B, Fang L, Posvar E. 2011. Single-dose, placebo-controlled, randomized study of AMG 785, a sclerostin monoclonal antibody. J Bone Miner Res 26:19–26 [DOI] [PubMed] [Google Scholar]
  • 60. Li J, Sarosi I, Cattley RC, Pretorius J, Asuncion F, Grisanti M, Morony S, Adamu S, Geng Z, Qiu W, Kostenuik P, Lacey DL, Simonet WS, Bolon B, Qian X, Shalhoub V, Ominsky MS, Zhu Ke H, Li X, Richards WG. 2006. Dkk1-mediated inhibition of Wnt signaling in bone results in osteopenia. Bone 39:754–766 [DOI] [PubMed] [Google Scholar]
  • 61. Morvan F, Boulukos K, Clément-Lacroix P, Roman Roman S, Suc-Royer I, Vayssière B, Ammann P, Martin P, Pinho S, Pognonec P, Mollat P, Niehrs C, Baron R, Rawadi G. 2006. Deletion of a single allele of the Dkk1 gene leads to an increase in bone formation and bone mass. J Bone Miner Res 21:934–945 [DOI] [PubMed] [Google Scholar]
  • 62. Wang FS, Ko JY, Lin CL, Wu HL, Ke HJ, Tai PJ. 2007. Knocking down dickkopf-1 alleviates estrogen deficiency induction of bone loss. A histomorphological study in ovariectomized rats. Bone 40:485–492 [DOI] [PubMed] [Google Scholar]
  • 63. Wang FS, Ko JY, Yeh DW, Ke HC, Wu HL. 2008. Modulation of Dickkopf-1 attenuates glucocorticoid induction of osteoblast apoptosis, adipocytic differentiation, and bone mass loss. Endocrinology 149:1793–1801 [DOI] [PubMed] [Google Scholar]
  • 64. Heiland GR, Zwerina K, Baum W, Kireva T, Distler JH, Grisanti M, Asuncion F, Li X, Ominsky M, Richards W, Schett G, Zwerina J. 2010. Neutralisation of Dkk-1 protects from systemic bone loss during inflammation and reduces sclerostin expression. Ann Rheum Dis 69:2152–2159 [DOI] [PubMed] [Google Scholar]
  • 65. Diarra D, Stolina M, Polzer K, Zwerina J, Ominsky MS, Dwyer D, Korb A, Smolen J, Hoffmann M, Scheinecker C, van der Heide D, Landewe R, Lacey D, Richards WG, Schett G. 2007. Dickkopf-1 is a master regulator of joint remodeling. Nat Med 13:156–163 [DOI] [PubMed] [Google Scholar]
  • 66. Fulciniti M, Tassone P, Hideshima T, Vallet S, Nanjappa P, Ettenberg SA, Shen Z, Patel N, Tai YT, Chauhan D, Mitsiades C, Prabhala R, Raje N, Anderson KC, Stover DR, Munshi NC. 2009. Anti-DKK1 mAb (BHQ880) as a potential therapeutic agent for multiple myeloma. Blood 114:371–379 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 67. Sims AM, Shephard N, Carter K, Doan T, Dowling A, Duncan EL, Eisman J, Jones G, Nicholson G, Prince R, Seeman E, Thomas G, Wass JA, Brown MA. 2008. Genetic analyses in a sample of individuals with high or low BMD shows association with multiple Wnt pathway genes. J Bone Miner Res 23:499–506 [DOI] [PubMed] [Google Scholar]
  • 68. Gaur T, Wixted JJ, Hussain S, O'Connell SL, Morgan EF, Ayers DC, Komm BS, Bodine PV, Stein GS, Lian JB. 2009. Secreted frizzled related protein 1 is a target to improve fracture healing. J Cell Physiol 220:174–181 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 69. Yao W, Cheng Z, Shahnazari M, Dai W, Johnson ML, Lane NE. 2010. Overexpression of secreted frizzled-related protein 1 inhibits bone formation and attenuates parathyroid hormone bone anabolic effects. J Bone Miner Res 25:190–199 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 70. Cho HY, Choi HJ, Sun HJ, Yang JY, An JH, Cho SW, Kim SW, Kim SY, Kim JE, Shin CS. 2010. Transgenic mice overexpressing secreted frizzled-related proteins (sFRP)4 under the control of serum amyloid P promoter exhibit low bone mass but did not result in disturbed phosphate homeostasis. Bone 47:263–271 [DOI] [PubMed] [Google Scholar]
  • 71. Clément-Lacroix P, Ai M, Morvan F, Roman-Roman S, Vayssière B, Belleville C, Estrera K, Warman ML, Baron R, Rawadi G. 2005. Lrp5-independent activation of Wnt signaling by lithium chloride increases bone formation and bone mass in mice. Proc Natl Acad Sci USA 102:17406–17411 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 72. Zamani A, Omrani GR, Nasab MM. 2009. Lithium's effect on bone mineral density. Bone 44:331–334 [DOI] [PubMed] [Google Scholar]
  • 73. Vestergaard P, Rejnmark L, Mosekilde L. 2005. Reduced relative risk of fractures among users of lithium. Calcif Tissue Int 77:1–8 [DOI] [PubMed] [Google Scholar]
  • 74. Bolton JM, Metge C, Lix L, Prior H, Sareen J, Leslie WD. 2008. Fracture risk from psychotropic medications: a population-based analysis. J Clin Psychopharmacol 28:384–391 [DOI] [PubMed] [Google Scholar]
  • 75. Wilting I, de Vries F, Thio BM, Cooper C, Heerdink ER, Leufkens HG, Nolen WA, Egberts AC, van Staa TP. 2007. Lithium use and the risk of fractures. Bone 40:1252–1258 [DOI] [PubMed] [Google Scholar]
  • 76. Kulkarni NH, Onyia JE, Zeng Q, Tian X, Liu M, Halladay DL, Frolik CA, Engler T, Wei T, Kriauciunas A, Martin TJ, Sato M, Bryant HU, Ma YL. 2006. Orally bioavailable GSK-3α/β dual inhibitor increases markers of cellular differentiation in vitro and bone mass in vivo. J Bone Miner Res 21:910–920 [DOI] [PubMed] [Google Scholar]
  • 77. Wang FS, Ko JY, Weng LH, Yeh DW, Ke HJ, Wu SL. 2009. Inhibition of glycogen synthase kinase-3β attenuates glucocorticoid-induced bone loss. Life Sciences 85:685–692 [DOI] [PubMed] [Google Scholar]
  • 78. Gambardella A, Nagaraju CK, O'Shea PJ, Mohanty ST, Kottam L, Pilling J, Sullivan M, Djerbi M, Koopmann W, Croucher PI, Bellantuono I. 2011. Glycogen synthase kinase-3α/β inhibition promotes in vivo amplification of endogenous mesenchymal progenitors with osteogenic and adipogenic potential and their differentiation to the osteogenic lineage. J Bone Miner Res 26:811–821 [DOI] [PubMed] [Google Scholar]
  • 79. Clevers H. 2006. Wnt/β-catenin signaling in development and disease. Cell 127:469–480 [DOI] [PubMed] [Google Scholar]
  • 80. Kansara M, Tsang M, Kodjabachian L, Sims NA, Trivett MK, Ehrich M, Dobrovic A, Slavin J, Choong PF, Simmons PJ, Dawid IB, Thomas DM. 2009. Wnt inhibitory factor 1 is epigenetically silenced in human osteosarcoma, and targeted disruption accelerates osteosarcomagenesis in mice. J Clin Invest 119:837–851 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 81. Gardner JC, van Bezooijen RL, Mervis B, Hamdy NA, Löwik CW, Hamersma H, Beighton P, Papapoulos SE. 2005. Bone mineral density in sclerosteosis; affected individuals and gene carriers. J Clin Endocrinol Metab 90:6392–6395 [DOI] [PubMed] [Google Scholar]
  • 82. Suzuki A, Ozono K, Kubota T, Kondou H, Tachikawa K, Michigami T. 2008. PTH/cAMP/PKA signaling facilitates canonical Wnt signaling via inactivation of glycogen synthase kinase-3β in osteoblastic Saos-2 cells. J Cell Biochem 104:304–317 [DOI] [PubMed] [Google Scholar]
  • 83. Tobimatsu T, Kaji H, Sowa H, Naito J, Canaff L, Hendy GN, Sugimoto T, Chihara K. 2006. Parathyroid hormone increases β-catenin levels through Smad3 in mouse osteoblastic cells. Endocrinology 147:2583–2590 [DOI] [PubMed] [Google Scholar]
  • 84. Wan M, Yang C, Li J, Wu X, Yuan H, Ma H, He X, Nie S, Chang C, Cao X. 2008. Parathyroid hormone signaling through low-density lipoprotein-related protein 6. Genes Dev 22:2968–2979 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 85. Guo J, Liu M, Yang D, Bouxsein ML, Saito H, Galvin RJ, Kuhstoss SA, Thomas CC, Schipani E, Baron R, Bringhurst FR, Kronenberg HM. 2010. Suppression of Wnt signaling by Dkk1 attenuates PTH-mediated stromal cell response and new bone formation. Cell Metab 11:161–171 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 86. Eijken M, Swagemakers S, Koedam M, Steenbergen C, Derkx P, Uitterlinden AG, van der Spek PJ, Visser JA, de Jong FH, Pols HA, van Leeuwen JP. 2007. The activin A-follistatin system: potent regulator of human extracellular matrix mineralization. FASEB J 21:2949–2960 [DOI] [PubMed] [Google Scholar]
  • 87. Sugatani T, Alvarez UM, Hruska KA. 2003. Activin A stimulates IκB-α/NFκB and RANK expression for osteoclast differentiation, but not AKT survival pathway in osteoclast precursors. J Cell Biochem 90:59–67 [DOI] [PubMed] [Google Scholar]
  • 88. Pearsall RS, Canalis E, Cornwall-Brady M, Underwood KW, Haigis B, Ucran J, Kumar R, Pobre E, Grinberg A, Werner ED, Glatt V, Stadmeyer L, Smith D, Seehra J, Bouxsein ML. 2008. A soluble activin type IIA receptor induces bone formation and improves skeletal integrity. Proc Natl Acad Sci USA 105:7082–7087 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 89. Lotinun S, Pearsall RS, Davies MV, Marvell TH, Monnell TE, Ucran J, Fajardo RJ, Kumar R, Underwood KW, Seehra J, Bouxsein ML, Baron R. 2010. A soluble activin receptor type IIA fusion protein (ACE-011) increases bone mass via a dual anabolic-antiresorptive effect in cynomolgus monkeys. Bone 46:1082–1088 [DOI] [PubMed] [Google Scholar]
  • 90. Ruckle J, Jacobs M, Kramer W, Pearsall AE, Kumar R, Underwood KW, Seehra J, Yang Y, Condon CH, Sherman ML. 2009. Single-dose, randomized, double-blind, placebo-controlled study of ACE-011 (ActRIIA-IgG1) in postmenopausal women. J Bone Miner Res 24:744–752 [DOI] [PubMed] [Google Scholar]
  • 91. Raje N, Vallet S. 2010. Sotatercept, a soluble activin receptor type 2A IgG-Fc fusion protein for the treatment of anemia and bone loss. Curr Opin Mol Ther 12:586–597 [PubMed] [Google Scholar]
  • 92. Segovia-Silvestre T, Neutzsky-Wulff AV, Sorensen MG, Christiansen C, Bollerslev J, Karsdal MA, Henriksen K. 2009. Advances in osteoclast biology resulting from the study of osteopetrotic mutations. Hum Genet 124:561–577 [DOI] [PubMed] [Google Scholar]
  • 93. Karsdal MA, Martin TJ, Bollerslev J, Christiansen C, Henriksen K. 2007. Are nonresorbing osteoclasts sources of bone anabolic activity? J Bone Miner Res 22:487–494 [DOI] [PubMed] [Google Scholar]
  • 94. Pennypacker BL, Duong le T, Cusick TE, Masarachia PJ, Gentile MA, Gauthier JY, Black WC, Scott BB, Samadfam R, Smith SY, Kimmel DB. 2011. Cathepsin K inhibitors prevent bone loss in estrogen-deficient rabbits. J Bone Miner Res 26:252–262 [DOI] [PubMed] [Google Scholar]
  • 95. Jerome C, Missbach M, Gamse R. 5 March 2011. Balicatib, a cathepsin K inhibitor, stimulates periosteal bone formation in monkeys. Osteoporos Int 10.1007/s00198-011-1529-x [DOI] [PubMed] [Google Scholar]
  • 96. Eisman JA, Bone HG, Hosking DJ, McClung MR, Reid IR, Rizzoli R, Resch H, Verbruggen N, Hustad CM, DaSilva C, Petrovic R, Santora AC, Ince BA, Lombardi A. 2011. Odanacatib in the treatment of postmenopausal women with low bone mineral density: three-year continued therapy and resolution of effect. J Bone Miner Res 26:242–251 [DOI] [PubMed] [Google Scholar]
  • 97. Nagase S, Ohyama M, Hashimoto Y, Small M, Kuwayama T, Deacon S. 30 June 2011. Pharmacodynamic effects on biochemical markers of bone turnover and pharmacokinetics of the cathepsin K inhibitor, ONO-5334, in an ascending multiple-dose, phase 1 study. J Clin Pharmacol doi: 10.1177/0091270011399080 [DOI] [PubMed] [Google Scholar]
  • 98. Eastell R, Nagase S, Ohyama M, Small M, Sawyer J, Boonen S, Spector T, Kuwayama T, Deacon S. 2011. Safety and efficacy of the cathepsin K inhibitor ONO-5334 in postmenopausal osteoporosis: the OCEAN study. J Bone Miner Res 26:1303–1312 [DOI] [PubMed] [Google Scholar]
  • 99. Pederson L, Ruan M, Westendorf JJ, Khosla S, Oursler MJ. 2008. Regulation of bone formation by osteoclasts involves Wnt/BMP signaling and the chemokine sphingosine-1-phosphate. Proc Natl Acad Sci USA 105:20764–20769 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 100. Ryu J, Kim HJ, Chang EJ, Huang H, Banno Y, Kim HH. 2006. Sphingosine 1-phosphate as a regulator of osteoclast differentiation and osteoclast-osteoblast coupling. EMBO J 25:5840–5851 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 101. Nakashima T, Hayashi M, Fukunaga T, Kurata K, Oh-Hora M, Feng JQ, Bonewald LF, Kodama T, Wutz A, Wagner EF, Penninger JM, Takayanagi H. 2011. Evidence for osteocyte regulation of bone homeostasis through RANKL expression. Nat Med 17:1231–1234 [DOI] [PubMed] [Google Scholar]
  • 102. Xiong J, Onal M, Jilka RL, Weinstein RS, Manolagas SC, O'Brien CA. 2011. Matrix-embedded cells control osteoclast formation. Nat Med 17:1235–1241 [DOI] [PMC free article] [PubMed] [Google Scholar]

Articles from The Journal of Clinical Endocrinology and Metabolism are provided here courtesy of The Endocrine Society

RESOURCES