Osteoimmunology is an interdisciplinary research field combining the exciting fields of osteology and immunology. An observation that contributed enormously to the emergence of osteoimmunology was the accelerated bone loss caused by inflammatory diseases such as rheumatoid arthritis. Receptor activator of nuclear factor ĸB ligand (RANKL), which is the main regulator of osteoclastogenesis, was found to be the primary culprit responsible for the enhanced activation of osteoclasts: activated T cells directly and indirectly increased the expression of RANKL, and thereby promoted osteoclastic activity. Excessive bone loss is not only present in inflammatory diseases but also in autoimmune diseases and cancer. Furthermore, there is accumulating evidence that the very prevalent skeletal disorder osteoporosis is associated with alterations in the immune system. Meanwhile, numerous connections have been discovered in osteoimmunology beyond merely the actions of RANKL. These include the importance of osteoblasts in the maintenance of the hematopoietic stem cell niche and in lymphocyte development as well as the functions of immune cells participating in osteoblast and osteoclast development. Furthermore, research is being done investigating cytokines, chemokines, transcription factors and co-stimulatory molecules which are shared by both systems. Research in osteoimmunology promises the discovery of new strategies and the development of innovative therapeutics to cure or alleviate bone loss in inflammatory and autoimmune diseases as well as in osteoporosis. This review gives an introduction to bone remodeling and the cells governing that process and summarizes the most recent discoveries in the interdisciplinary field of osteoimmunology. Furthermore, an alternative large animal model will be discussed and the pathophysiological alterations of the immune system in osteoporosis will be highlighted.

The term ‘osteoimmunology’ was first used in the year 2000 as Aaron and Choi were highlighting the interdigitate communication between the immune and skeletal systems especially observed in autoimmune and other inflammatory diseases [1]. Major advances and discoveries in this interdisciplinary research field have lead to the revelation of molecular mechanisms as well as various cytokines and signaling transducers participating in the regulatory interplay between immune cells and bone cells. Furthermore, besides the arsenal of mutual signaling molecules, immune and bone cells also share a common site of origin, namely bone marrow. Due to the spatial proximity of the developing cells it is proposed that they influence each other not only after maturation and activation, as Kong and colleagues already observed in 1999 [2, 3], but also at the very beginning of their existence. Among others, Taichman and Emerson have described the important role of osteoblasts in the establishment of hematopoietic stem cell niches, as well as in the engraftment and maintenance of hematopoietic stem cells (HSC) in the bone marrow [4,5,6,7,8]. The development of bone cells has also been demonstrated to be supported by cells of the immune system; for instance, macrophages encourage osteoblastogenesis by the secretion of interleukin-18 [9], and T cells are capable of influencing osteoclastogenesis by the secretion of various cytokines such as interleukin-1, interleukin-6, interferon-γ or interleukin-4 [10, 11].

However, the most prominent osteoimmunological example arose from the observation of osteoclast-mediated bone loss in various inflammatory and autoimmune diseases such as rheumatoid arthritis, diabetes mellitus, lupus erythematosus, periodontal diseases and chronic viral infections (human immunodeficiency virus) [3,12,13,14,15,16]. Kong and colleagues were the first to notice that the osteoclast-inducing capacity of activated T cells in adjuvant arthritis was mediated through the RANKL/receptor activator of nuclear factor ĸB (RANK)/osteoprotegerin (OPG) axis. RANKL is a key regulator of osteoclastogenesis, thus contributing enormously to a process called bone remodeling. In this process, osteoclasts resorb old and damaged bone, which is then replaced by new bone material deposited by osteoblasts. As normal physiological bone remodeling is imperative for the maintenance of bone strength and integrity, imbalances either lead to increased or decreased bone mass the latter often being caused by inflammatory diseases.

Considering all of these intricate interrelationships between the immune and skeletal system and the high potential for developing innovative new therapeutic drugs targeting osteoclastogenesis-regulating cytokines, it is worth taking a closer look at the molecular mechanisms facilitating osteoclastogenesis and the environmental factors affecting that process.

Contrary to the general perception of bone being an inert, static material, it is a highly organized, living tissue and is the major constituent of the skeleton. The most prominent functions of bone are the protection of internal organs and the support of body structures. Beyond those functions bone additionally serves as an attachment site for muscles allowing locomotion and as an appropriate cavity for hematopoiesis in bone marrow. As a reservoir for inorganic ions, bone is responsible for the maintenance of calcium homeostasis and is able to rapidly mobilize mineral stores on metabolic demand.

Bone is composed of cells and extracellular matrix (ECM), the latter being further subdivided into an inorganic and organic part. The organic matrix is mainly constituted of type I collagen (approximately 95%), as well as other types of collagens, noncollagenous proteins and proteoglycans, whereas the inorganic matrix predominantly contains calcium and phosphorus, appearing as hydroxyapatite crystals ([3Ca3(PO4)2](OH)2) deposited into the collagenous matrix. This interdigitate organization confers rigidity and strength to the skeleton while maintaining a certain degree of elasticity. The major cells in bone are the osteoclasts, resorbing bone tissue, and osteoblasts (including osteocytes and bone lining cells), depositing bone tissue.

The main functions of osteoblasts are to synthesize the collagen-rich organic matrix and to provide optimal conditions for matrix mineralization by secreting numerous bone matrix proteins and matrix metalloproteinases (MMP) [17]. Furthermore, osteoblasts support hematopoiesis, notably osteoclastogenesis [18, 19]. Bone lining cells, one possible destiny of fully differentiated osteoblasts, are responsible for the initiation of bone remodeling by matrix degradation [20], whereas osteocytes, the other form of terminally differentiated osteoblasts, act as mechanosensors in bone tissue, thereby regulating bone mass and structure [21,22,23,24].

Osteoblasts, chondrocytes, adipocytes, stromal cells, myoblasts and tenocytes all originate from a common progenitor cell, the mesenchymal stem cell (MSC) [25,26,27]. The multipotent MSC undergoes several steps of commitment to give rise to progeny with more limited capacities until the differentiated end-stage cell is able to express particular functional markers and morphological traits. Core binding factor 1 (Cbfa1, also termed runt-related transcription factor 2, runx2) and the downstream factor osterix are crucial transcription factors for lineage commitment and osteoblast differentiation [28, 29]. The skeletons of Cbfa1 deficient mice only consist of cartilage, indicating its importance in osteoblast development [30, 31]. Recently, Jones et al. [32] identified the zinc-finger protein schnurri-3 (shn-3) which in association with WWP1 acts as a regulator of Cbfa1 expression by influencing cbfa1 degradation by ubiquitination.

Mature osteoblasts continue matrix deposition and start mineralization, expressing alkaline phosphatase and bone sialoprotein as well as osteocalcin and osteopontin [33]. Since osteoblasts produce a variety of proteins for bone matrix synthesis they are hallmarked by a prominent Golgi apparatus and rough endoplasmic reticulum. Osteoblasts turn into osteocytes after they have been surrounded by bone matrix. Osteocytes are poor in organelles, indicating other primary functions than matrix synthesis and mineralization. In fact, mature osteocytes alter their morphology by forming dendritic processes which enable them to communicate with other embedded osteocytes. These processes are believed to act as mechanosensors in bone tissue, which allow them to react to environmental changes [34,35,36,37]. Moreover, there is emerging evidence that apoptotic osteocytes increase the secretion of osteoclastogenic cytokines and thereby enhance bone resorption [38, 39]. Bone lining cells also arise from osteoblasts and are believed to be resting, inactive osteoblasts, covering the bone surface. As mentioned earlier, an important role in the initiation of bone remodeling has been assigned to them [40].

Osteoclasts are tissue-specific giant polykaryons derived from the monocyte/macrophage hematopoietic lineage and are the only cells capable of breaking down mineralized bone, dentine and calcified cartilage [41, 42]. The presence of RANKL and M-CSF (macrophage-colony-stimulating factor) are essential for the formation and fusion of multinucleated cells, expressing osteoclast-specific markers such as tartrate-resistant acid phosphatase (TRAP), cathepsin K, calcitonin receptor (CTR) and integrin receptors [43,44,45,46,47,48]. Via integrins, osteoclasts attach very tightly to the matrix, thereby creating an isolated lacuna (Howship’s lacuna) able to maintain an acidic environment necessary for matrix dissolution [49]. After attachment intracellular rearrangements lead to the polarization of the cell borders, whereas the sealing zone is adjacent to the basolateral domain and the ruffled border, respectively. At the opposite side of the ruffled border emerges the functional secretory domain (FSD) [50]. The ruffled border and the FSD are connected to each other via microtubules on which exocytotic vesicle traffic has been observed [51], suggesting the secretion of resorbed material into the extracellular space. In addition to the development of distinct membrane domains, the cytoskeleton undergoes organizational changes creating a dense actin ring in osteoclasts preparing for resorption.

The resorption of bone matrix takes place in the resorption lacuna [52]. The ruffled border is formed by the fusion of cytoplasmic acidic vacuoles, thereby releasing acid into the resorption lacuna and initiating rapid dissolution of the hydroxyapatite crystals [53]. Additionally, ATPases located in the ruffled border transport protons into the Howship’s lacuna. The protons are supplied by the reaction of water and carbon dioxide catalyzed by the enzyme carbonic anhydrase II resulting in the formation of protons and HCO3. Whereas the protons are pumped into the resorption lacuna HCO3 is transported into the extracellular space via HCO3/Cl exchanger. The imported chloride ions are also pumped into the resorption lacuna to form hydrochloric acid capable of dissolving the mineralized matrix. The organic matrix is degraded by various enzymes, including tartrate-resistant acid phosphatase (TRAP), cathepsin K and matrix metalloproteinase 9 (MMP-9) [20,54,55,56].

Continuously changing functional demands require permanent adaption of the bone structure and microarchitecture. Wolff [57 ]has observed this principle of functional adaptation already over 100 years ago. The process of where ‘form follows function’ [58] consists of two activities, namely, bone formation and bone resorption. While these processes are locally separated in modeling [59, 60], bone remodeling is characterized by the spatial and temporal coupling of bone formation by osteoblasts and bone resorption by osteoclasts [61]. Approximately 5–25% of bone surface is undergoing bone remodeling [62, 63], thereby restoring microdamages and ensuring mechanical integrity as well as regulating the release of calcium and phosphorus.

Bone remodeling involves four main processes: activation, resorption, reversal and formation [64]. The remodeling cycle is initiated by the activation of the quiescent bone surface, which is covered with bone lining cells [65, 66]. Osteoclast precursor cells are recruited to the activated surface and fuse to form mature, bone resorbing osteoclasts. The osteoclasts attach to the surface, dissolve the inorganic matrix by creating an acidic microenvironment, and degrade the organic matrix with specific enzymes. As bone resorption subsides and resorption pits remain, osteoclasts disappear and mononuclear cells prepare the surface for bone formation. The bone remodeling cycle is finished with the synthesis and deposition of bone matrix by osteoblasts, and bone lining cells building a canopy covering the surface, keeping the material dormant until the next cycle.

Prostaglandins, Leukotrienes and Hormones

Bone formation and resorption are under the subtle control of various local and systemic factors. Besides the multiple cytokines that participate in the regulatory system of bone homeostasis prostaglandins, leukotrienes and hormones additionally interact with bone cells, thereby affecting bone remodeling processes.

Prostaglandins and leukotrienes are metabolites of arachidonic acid and have stimulatory as well as inhibitory effects on bone [67]. In the presence of factors stimulating bone resorption, prostaglandin E2 (PGE2) is produced by the cyclo-oxygenase 2 (COX2)-mediated conversion of arachidonic acid in osteoblasts. Recently, Ha et al. [68] were able to demonstrate an inhibition of osteoclastogenesis primarily by the reduction of IL-1-induced COX2 activity and PGE2 production and only marginally by the suppression of RANKL by α-lipoic acid [69,70,71]. Additionally, leukotriene B4 was also shown to enhance osteoclastogenesis in a RANKL-independent manner [72,73,74].

The peptide hormone parathyroid hormone (PTH) is one of the most important regulators of calcium ion homeostasis [75, 76]. In response to low blood calcium levels, PTH is secreted into the circulation and acts on kidney, bone and intestine to maintain blood calcium concentrations. In bone, PTH upregulates the production of interleukin-6 and RANKL by osteoblasts, thereby facilitating the differentiation, activation and survival of osteoclasts [77,78,79,80,81,82]. Thus, PTH as well as PTHrP (PTH-related protein) promote bone resorption and consequently the release of calcium [83, 84].

The active hormonal form of vitamin D, calcitriol (1α,25-dihydroxyvitamin D3), is essential for the development and maintenance of the mineralized skeleton, as demonstrated in various studies using vitamin D receptor or 1α(OH)ase knock-out mice [85,86,87]. The impaired mineralization was normalized when a high-calcium, high-phosphate and high-lactose diet (rescue diet) was administered. However, Panda and colleagues were able to show that the administration of only 1,25(OH)2D3 to 1α(OH)ase knock-out mice was not sufficient to normalize the impaired mineralization if hypocalcemia was not corrected [85]. Furthermore, vitamin D-deficient mice showed an increase in osteoblast number, bone formation and bone volume as well as increased serum alkaline phosphatase levels. Additionally, osteoclast numbers were decreased due to a decreased production of RANKL and an enhanced production of OPG [88].

Concerning bone homeostasis, estrogen and androgens are the most intensively investigated sex steroids and, in contrast to PTH and 1,25-dihydroxyvitamin D3, enhance bone formation and inhibit bone resorption [89,90,91,92,93]. Estrogen as well as testosterone deficiency inevitably lead to an increased rate of bone turn-over, which has been demonstrated by Jilka et al.[ 94], who observed simultaneous increases in osteoclastic precursors as well as early osteoblastic precursors. Additionally, estrogen deficiency results in a net loss of bone as a consequence of an increased production of RANKL and a decreased production of OPG in osteoblastic cells as well as the enhancement of the secretion of pro-inflammatory and pro-resorptive cytokines in lymphocytes such as IL-1, IL-6 and tumor necrosis factor-α (TNF-α) [95,96,97,98,99,100]. Moreover, the bone-protective effect of estrogen has been demonstrated to be mediated by transforming growth factor-β (TGF-β), inducing apoptosis in osteoclasts [101,102,103]. In two randomized controlled trials from the Women’s Health Initiative, hormone replacement therapy (HRT) had been shown to decrease the incidence of major osteoporotic fractures [227, 228]. However, serious side effects such as cardiovascular disease and cancer have occurred and therefore other medications are used nowadays in the treatment of osteoporosis. Raloxifene is a selective estrogen receptor modulator (SERM) and is approved for the treatment of osteoporosis. Like estrogen, SERMs are known to mediate their effects through the estrogen receptor. While estrogen binds equally strongly to alpha and beta receptors, raloxifene preferentially binds to the alpha receptor. This specificity to a certain estrogen receptor allows a higher affinity to bone, and therefore the side effects of SERMs are less pronounced than those of HRT [229].

The discovery of RANKL and its receptors RANK and OPG has highlighted the molecular processes in osteoclastogenesis, raising the possibility to inhibit the development of osteoclasts, rescuing bone from exorbitant resorption. In 1997, Simonet et al.[ 104] discovered a protein which exposed an osteopetrotic phenotype when overexpressed in transgenic mice. Investigating further, they found that this protein was secreted by preosteoblasts/stromal cells and was capable to inhibit osteoclast development and activation. Due to its bone-protective effects they named it osteoprotegerin. OPG belongs to the TNF receptor superfamily, however lacking a transmembrane and cytoplasmic domain. OPG is expressed in a variety of tissues, including lung, heart, kidney, liver, stomach, intestine, brain, spinal cord, thyroid gland, smooth muscle tissue and bone, indicating multiple possible functions [105,106,107,108,109,110]. The most prominent role of OPG has been assigned to bone protection; however, recent investigations have also proposed important functions of OPG in endothelial cell survival [111, 112] and vascular calcification [113,114,115].

After the identification of OPG followed the discovery of RANKL, which possesses not only a huge repertoire of names (TRANCE: TNF-related activation-induced cytokine; ODF: osteoclast differentiating factor; OPGL: osteoprotegerin ligand; TNFSF11: TNF superfamily member 11) but also many faces regarding its structure, function and appearance in tissues. The names originated from the four discoverers, each one having used different approaches to identify the protein. They searched either for a ligand for OPG [106], screened for apoptosis-regulating genes in T cell hybridomas [116], or found RANKL to induce osteoclastogenesis [117] and enhance the life-span of dendritic cells [118]. Kartsogiannis et al. [119] detected RANKL protein and mRNA expression in a variety of tissues, including bone, brain, heart, kidney, liver, lung, intestine, skeletal muscle, mammary tissue, placenta, spleen, thymus and testis [reviewed in [120]. This extensive distribution of RANKL throughout the body already indicates its multiple functions, whereas the most important one is dedicated to the induction of osteoclastogenesis, thus, the regulation of bone remodeling. RANKL knock-out mice reveal a severe osteopetrotic phenotype due to the absence of osteoclasts. Furthermore, defects in tooth eruption, lymph node genesis, mammary gland and lymphocyte development were reported as well as disturbances in T cell/dendritic cell interactions [121, 122]. Recently, Jones et al.[ 123] reported the observation of RANKL triggering cell migration of cancer cells expressing the receptor RANK indicating a role for RANKL as a chemo-attractant for cancer cells.

RANKL is a member of the TNF superfamily and is mainly expressed in preosteoblasts/stromal cells as well as on activated T cells [106, 117, 118, 124] and is closely related to the TNF-related apoptosis-inducing ligand TRAIL (homology ∼30%) and FasL (homology ∼20%). The human RANKL gene has been localized to chromosome 13q14 and encodes for three isoforms: RANKL1 and RANKL2 are type II transmembrane proteins, whereas RANKL2 encodes for a shorter intracellular domain. RANKL3 is a soluble protein, partially produced by the cleavage of the membrane-bound form by TACE (TNFα-converting enzyme, a metalloprotease) or other MMPs and by the direct transcription of the alternatively spliced RANKL-encoding gene [125,126,127,128,129].

The expression of OPG and RANKL is highly inducible by various systemic and local factors. Among others, estrogens, bone morphogenetic protein (BMP) 2, INF-γ and TGF-β positively regulate OPG, whereas PTH, 1,25(OH)2 vitamin D3, glucocorticoids, prostaglandin E2, IL-6, IL-8 and IL-11 enhance the expression of RANKL [reviewed in [130]].

The third participant in the bone remodeling regulatory system is RANK (receptor activator of nuclear factor ĸB) and belongs to the TNFR superfamily like OPG [118]. RANK represents a type I transmembrane protein and is expressed in tissues as ubiquitously as RANKL, although most commonly found in osteoclasts and dendritic cells [131]. RANK-deficient mice show similar phenotypes to those of RANKL knock-out mice, including tooth eruptions, osteopetrosis and missing lymph nodes [132].

The RANK signaling cascade (fig. 1) is initiated when RANKL binds to the extracellular domain of RANK which passes the signal along to TRAF6 (TNF receptor-associated factor 6) [133,134,135,136,137,138]. By transfecting RANK deficient cells in RANK mutants that are incapable of binding either TRAF 1, 2, 3, 5, or 6 Armstrong et al.[ 134] were able to demonstrate that predominantly TRAF6 was essential for the induction of osteoclastogenesis. TRAF6 has various downstream mediators which control the expression of osteoclast-specific genes during differentiation and activation of osteoclasts. The two most investigated pathways are the activation of the transcription factors NF-ĸB and AP-1 (activated protein 1). Targeted disruptions of the p50/p52 component of NF-ĸB and the c-fos component of AP-1 resulted in impaired osteoclastogenesis and revealed an osteopetrotic phenotype [139,140,141]. AP-1 is activated by signaling cascades mediated by JNK (c-Jun N-terminal kinase) whereas the phosphorylation of the inhibitor of NF-ĸB kinase (IKK) leads to the activation of NF-ĸB [142,143,144,145,146,147]. Other cascades of mitogen-activated protein kinases such as the TGF-β-inducible kinase TAK1 and the p38 stress kinase have also been described to participate in RANK signal transduction [148,149,150]. p38 is activated via the phosphorylation by MKK6 and in turn activates the transcriptional regulator mi/Mitf, which is responsible for the transcriptional control of genes encoding for the osteoclast-specific enzymes TRAP and cathepsin K [151, 152]. ERK (extracellular signal-related kinase) is a downstream target of MEK1 and acts as a negative regulator of osteoclastogenesis for ERK inhibitors have shown to accelerate RANKL-induced osteoclastogenesis [153]. The serine/threonine kinase Akt and the phosphatidylinositol-3-OH kinase (PI(3)K) are downstream elements of src and are known to mediate cell survival, motility and cytoskeletal rearrangements by the activation of the MEK/ERK and Akt/NFĸB pathways [121,154]. Recently, AFX/FOXO4 was shown to be the key downstream mediator activated by Akt/PKB to modulate osteoclast survival [155].

Fig. 1

RANK signaling in osteoclasts. The RANK signaling cascade is initiated upon the binding of RANKL to the extracellular domain of RANK which passes the signal along to TRAF6. The activation of TRAF6 initiates pathways leading to the activation of the transcription factors NFAT, NFkB, the MAP kinase mediators jun, fos and p38 as well as the down-stream targets of Akt AFX/FOXO4, which contribute to osteoclast differentiation, activation and survival. The ITAM-containing co-stimulatory molecules DAP12 and FcRγ, respectively, initiate Ca2+-signaling leading to the activation of NFATc1. This schematic representation only focuses on the most important pathways, not illuminating further interactions of the signaling mediators.

Fig. 1

RANK signaling in osteoclasts. The RANK signaling cascade is initiated upon the binding of RANKL to the extracellular domain of RANK which passes the signal along to TRAF6. The activation of TRAF6 initiates pathways leading to the activation of the transcription factors NFAT, NFkB, the MAP kinase mediators jun, fos and p38 as well as the down-stream targets of Akt AFX/FOXO4, which contribute to osteoclast differentiation, activation and survival. The ITAM-containing co-stimulatory molecules DAP12 and FcRγ, respectively, initiate Ca2+-signaling leading to the activation of NFATc1. This schematic representation only focuses on the most important pathways, not illuminating further interactions of the signaling mediators.

Close modal

However, on the level of transcription factors nuclear factor of activated T cells c1 (NFATc1) has been elected as the master regulator of osteoclastogenesis. Many downstream effectors of RANK such as NF-ĸB and AP-1 contribute to the activation of NFATc1. Furthermore, RANKL-induced Ca2+-signaling, mediated by immunoreceptor tyrosine-based activation motifs (ITAMs), has been shown to be indispensable for osteoclastogenesis, since mice deficient in the ITAM-containing adaptor molecules DAP (DNAX-activating protein) 12 and Fc common receptor γ chain (FcRγ) are severely osteopetrotic. After retroviral transfer of DAP12 the osteopetrotic phenotype was rescued. The association of paired immunoglobulin-like receptor A (PIR-A) and OSCAR to FcRγ and triggering receptor expressed on myeloid cells (TREM) 2 and signal-regulatory protein β1 (SIRPβ1) to DAP12 in osteoclast precursors is considered to act as a co-stimulatory signal for RANKL, since one signal by its own is not able to induce osteoclastogenesis [156,157,158].

Recent studies dealt with the effects cytokines have on the generation of osteoblasts and osteoclasts. It is known that IL-1α, IL-1β, IL-6 and other members of the gp130 cytokine family, IL-7 and TNF-α directly or indirectly promote osteoclastogenesis [159,160,161,162], whereas interferon-beta (IFN-β), IFN-γ, IL-3, IL-4, IL-10, IL-13, and IL-12 alone and in synergy with IL-18 [163,164,165,166,167,168], amongst others, inhibit osteoclast formation. TGF-β was found to both induce, via suppressor of cytokine signaling 3 (SOCS3) [163], and suppress osteoclastogenesis [169]. For a detailed list, see the review by Theolyre et al. [120].

Among the osteoclastogenesis-inhibiting cytokines, interferons have attracted increasing attention. Thus, IFN-γ, a cytokine produced by activated T cells, was identified to strongly suppress osteoclastogenesis by inhibiting RANKL signaling by downregulation of the transcription factor TRAF6 expression via Stat1. The same applies for IFN-β, which is especially interesting in that this cytokine is induced by RANKL via a second down-stream molecule, namely c-Fos, but at the same time acts as negative regulator of RANKL signaling by inhibiting c-Fos expression in terms of a negative feedback loop. This interference of interferons with osteoclast differentiation was reviewed in detail by Takayanagi et al. [170].

However, there is little detailed information on the cytokine production pattern of osteoblasts. IL-6 was shown to be produced by stromal cells/osteoblasts [171]. A number of growth factors and hormones are known to promote proliferation and differentiation of osteoblasts, such as TGF-β (which is also assumed to depress osteoblast differentiation) [172], parathyroid hormone, its locally produced homologue parathyroid hormone-related peptide (PTHrP), low-density lipoprotein receptor-related protein-5 (LRP-5) [173], and osteopontin [174]. In recent research in osteology, much attention has been attributed towards bone morphogenic proteins (BMPs). So far, in pigs BMP-6 and BMP-7 have been shown to increase osteogenic differentiationin vitro [175], and BMP-4, besides IGF-1 and TGF-β, was found to be expressed during distraction osteogenesis in a pig model [176]. For human mesenchymal stem cells (MSCs), among BMP-2, -4, -6, and -7, BMP-6 was found to be the most consistent and potent regulator of osteoblast differentiation. Addition of BMP-6 to MSCs in vitro leads to the upregulation of type I collagen, osteocalcin, and bone sialoprotein [177]. Production of IL-6 and RANKL by osteoblasts is promoted by PTH and TNF-α, but with markedly different kinetics. Whereas PTH induces only a rapid, but transient elevation of both cytokines, TNF-α leads to a biphasic increase of these cytokines, thus indicating the potent role of TNF-α in pathologic conditions [77].

There is general agreement that lymphocytes influence bone remodeling by exerting an impact on osteoclastogenesis (fig. 2). Thus, T cells are assumed to be responsible for bone loss which occurs as a consequence of a series of pathological conditions, for example systemic viral infections and chronic local bone and joint diseases, such as rheumatoid arthritis or inflammatory bowel disease [3, 178]. Concerning the type of impact that T cells exert on osteoclastogenesis results from in vitro and in vivo experiments differ to a high degree. The same is true for different lymphocyte subpopulations, i.e. data concerning the effect of CD4 and CD8 lymphocytes on osteoclastogenesis, are not consistent. On the one hand, data from literature suggest an inhibitory effect of T cells. In one in vitro study, 1α,25(OH)2D3-stimulated osteoclast-like cell formation was enhanced after lymphocyte depletion. This was attributed to increased PGE2 production and consecutive upregulated RANKL and downregulated OPG expression [179]. IFN-γ was found to be the modulatory factor, which is produced by activated (by anti-CD3ε) T cells, and which interferes with TRAF6, thus strongly inhibiting the RANKL-induced activation of NF-ĸB and JNK in vitro. Resting T cells were found to exert no effect on osteoclastogenesis in the cited reference. These results were confirmed by another experiment, demonstrating that activated T cells have no effect in co-culture with IFN- γR–/– BMMs (bone marrow-derived monocyte/macrophage precursor cells) stimulated by RANKL [11]. In contrast to the above mentioned results, resting T cells were also found to negatively regulate osteoclastogenesis via production of granulocyte/monocyte colony-stimulating factor (GM-CSF) and IFN-γ by CD4 but not CD8 T cells [180]. Another in vitro study demonstrated that the downregulatory effect of lymphocytes is due to the CD8 T cell subset, and independent of IL-4 and TGF-β [181]. On the other hand, activated T cells were shown to promote osteoclastogenesis in vitro andin vivo. Activated (by anti-CD3ε and anti-CD28) CD4 T cells exerted their effect via membrane-bound and secretory RANKL. Transfer of ctla4–/– bone marrow cells in ragl–/– mice led to a significant decrease in bone mineral density. Consistent results were achieved by direct transfer of purified ctla4–/– T cells in ragl–/– and opgl–/– mice [3]. Recently, it was demonstrated that the effects of activated T cells on osteoclastogenesis depend on how they are activated [182]. Anti CD3ε- and anti CD28-Ab-activated T cells inhibited osteoclastogenesis, whereas T cells activated with staphylococcal enterotoxin A, PHA, and Con A had inconsistent effects. The osteoclastogenic effect was CD4+-dependent.

Fig. 2

Cellular regulation of osteoclastogenesis. Osteoblasts/stromal cells are the main regulators of osteoclastogenesis. They express the cytokines RANKL, which binds to RANK on osteoclast precursors and thereby induces osteoclastogenesis, and OPG, which is able to prevent that interaction. Amongst others, TGF-β and 17-β-estradiol stimulate the production of OPG whereas 1,25(OH)2D3, PTH and PGE2 promote the production of RANKL. Upon activation via dendritic cells T-cells activate osteoclasts directly through the secretion of sRANKL. Furthermore, T cells secrete INFg, which on the one hand stimulates macrophages to produce pro-inflammatory cytokines which in turn promote RANKL expression in osteoblasts/stromal cells, and on the other hand suppresses permanent osteoclast activation by the destruction of TRAF6. Furthermore, endothelial cells have been shown to express RANKL and OPG and might therefore also participate in the regulation of osteoclastogenesis. MyP = Myeloid progenitor; OC-P = osteoclast precursor; aOC = activated osteoclast; OB/SC = osteoblast/stromal cell; MΦ = macrophage; T-C = T cell; DC = dendritic cell; EC = endothelial cell. Modified from Yasuda et al. [106].

Fig. 2

Cellular regulation of osteoclastogenesis. Osteoblasts/stromal cells are the main regulators of osteoclastogenesis. They express the cytokines RANKL, which binds to RANK on osteoclast precursors and thereby induces osteoclastogenesis, and OPG, which is able to prevent that interaction. Amongst others, TGF-β and 17-β-estradiol stimulate the production of OPG whereas 1,25(OH)2D3, PTH and PGE2 promote the production of RANKL. Upon activation via dendritic cells T-cells activate osteoclasts directly through the secretion of sRANKL. Furthermore, T cells secrete INFg, which on the one hand stimulates macrophages to produce pro-inflammatory cytokines which in turn promote RANKL expression in osteoblasts/stromal cells, and on the other hand suppresses permanent osteoclast activation by the destruction of TRAF6. Furthermore, endothelial cells have been shown to express RANKL and OPG and might therefore also participate in the regulation of osteoclastogenesis. MyP = Myeloid progenitor; OC-P = osteoclast precursor; aOC = activated osteoclast; OB/SC = osteoblast/stromal cell; MΦ = macrophage; T-C = T cell; DC = dendritic cell; EC = endothelial cell. Modified from Yasuda et al. [106].

Close modal

In mice it was shown that T cells are not absolutely required for osteoclastogenesis in a rheumatoid arthritis model, although they form an important pathologic feature in arthritic joints [183]. T cells in rheumatic joints are known to be in a kind of frustrated state, which is characterized by a downregulation of IFN-γ-production [184, 185]. Therefore it seems as if proinflammatory cytokines, which are produced by macrophages upon stimulation by T cells, act on synovial fibroblasts. These cells are the main source of RANKL responsible for osteoclast differentiation, although RANKL has also been shown to be produced by T cells. This model may be sufficient for explaining why T cells, which on the one hand produce IFN-γ, act osteoclastogenically by interacting with other cells in the rheumatoid joint.

Besides lymphocyte-osteoclast interactions there is now increasing awareness of the importance of osteoblasts in osteoimmunology. These cells seem to have antigen-presenting properties, since osteoblast cell lines were shown to express MHCII molecules besides the adhesins CD54 (ICAM-1) and CD166 (ALCAM), which were upregulated through IFN-γ, and could thus also activate T cells. Furthermore, osteoblasts were shown to express members of the toll-like receptor family, in particular TLR-4, TLR-5 and TLR-9, indicating an active role in host immune response. Pattern recognition receptors were found not only on the surface of osteoblasts but also intracellularly, as was recently reported by Marriott et al.[ 186] who were able to demonstrate the expression of the nucleotide-binding oligomerization domain proteins NOD1 and NOD2 following bacterial challenge of the cells. On the other hand, osteoblasts can produce IL-6 upon encountering T cells and stimulation by IL-17 [187]. Stanley et al. also demonstrated the ability of osteoblastic cell lines to present superantigen to T cells. This is of importance as superantigens are implicated in a variety of autoimmune conditions, such as rheumatoid arthritis.

Hematopoietic stem cells (HSC) are located in the bone marrow and are responsible for the continuous production of blood cells in an adult organism. Their capacity for self-renewal and their ability to differentiate into multiple cell types is strongly dependent on their surrounding microenvironment, which is also referred to as stem cell niche. There, cells produce various signaling molecules, cell adhesion molecules and components of the extracellular matrix and thereby determine the long-term repopulating ability of stem cells. Taichman and Emerson were among the first to notice that osteoblasts play a crucial role in stem cell maintenance due to an intimate cell-to-cell contact via integrins [5,6,7,8]. Another interesting observation was made in cbfa1 deficient mice, which were devoid of osteoblasts. In addition, those mice were characterized by the absence of bone marrow, although they showed normal hematopoietic development in liver and spleen until day E17.5, suggesting an important role for osteoblasts in HSC homing into the bone marrow cavity [31, 32]. Using a chimeric mouse model Kronenberg [188] was able to demonstrate the inhibition of HSC homing into the bone marrow after the deletion of the G-protein Gsα. Furthermore, he and others reported the supporting effects of osteogenic PTHs in HSC maintenance by stimulating bone lining cells to produce N-cadherin, important for stem cell attachment, and jagged-1, activating notch receptors on HSC [188, 189]. Arai et al.[ 190] identified a quiescent and anti-apoptotic subpopulation of HSC adhering to osteoblasts via the receptor tyrosine kinase Tie2 on HSCs and angiopoietin-1 on osteoblasts. Furthermore, the interaction of Tie2 with angiopoietin-1 increased the cadherin and intergrin mediated cell adhesion to osteoblasts and maintained the long-term repopulating activity of HSCs.

Osteology is a fast developing branch in gerontological and rheumatological research, since postmenopausal osteoporosis as well as osteoporosis in elderly men is of great importance for individual well-being and public health. Most scientific work in medical research is performed in rodent models and hardly any experiments are conducted in larger animals such as cats, dogs or rabbits. Porcine cells should be paid more attention in osteology as porcine bone tissue is more closerly related to humans [191]. Thus, it would be of great benefit for osteological research to create a well characterized animal model of similar size and physiology to the human species. Like most other mammals the pig is a quadruped. As a consequence its skeleton is subjected to different forces when compared to humans, which is a weakness of this model from the biomechanical point of view. Nevertheless, static features of its bone apparatus, joints, muscles, and tendons exhibit more similarities to humans than do those of rodents. An additional major advantage of the porcine system over the rodent system is the opportunity to harvest a strikingly higher amount of bone marrow and blood cells per individual. Besides that, treatment of some osteopathologies in farm animals is also of economic interest. Well-fitting examples for such pathologic conditions in the porcine species are osteochondrosis dessicans and the humpback syndrome in fattening pigs. Many aspects of the etiology, pathogenesis, and treatment are still unclear. For that reason, in vitro models are still one important option to enlarge our knowledge concerning basic and therapeutic mechanisms in osteology. Since in pigs no information on the cytokine pattern of bone marrow-derived cells from healthy animals is currently available, our main interest was to collect basic data regarding the cytokine pattern of these cells cultured in and without presence of 1α,25(OH)2D3, which is known to promote osteoclastogenesis. This was considered an important issue, as bone marrow stromal cells (and lymphocytes) provide the milieu leading to increased or decreased osteoclastogenesis to a large extent by their cytokine production pattern. Since RANKL plays an important role in the generation of osteoclasts in humans and rodents [106], it was of particular interest to search for indications of the existence of a porcine homologue. At the mRNA level, cytokines with the most remarkable expression intensities were IL-1α, IL-6 and IL-8. Immunofluorescent staining with specific mAbs recognizing a panel of porcine cytokines revealed the presence of IL-1β and IL-6 in stromal cells and in osteoclasts. Since the production of IL-6 has predominantly been attributed to stromal cells [171], but was also shown to be produced by osteoclasts in our experiments, it appears that these cells could support their generation in a positive feedback fashion. Overall expression of TNF-α mRNA was low in the cultured bone marrow-derived cells. When investigating production of TNF-α at the protein level, stromal cells exhibited only faint immunofluorescent signals, but osteoclasts, in contrast, gave a bright fluorescence [192]. This is in accordance with the murine system, where osteoclasts are known to produce considerable amounts of TNF-α [193].

In the recent past, scientific interest in bone biology was focused on the RANK/RANKL/OPG osteoclastogenesis-regulatory cytokine system. We found that the porcine RANKL homologue not only was expressed in bone marrow-derived cells, but also could be detected in white blood cells by RT-PCR. The following sequencing step confirmed the PCR results by exhibiting a considerable nucleotide homology to human and murine sequences which was 79% in the case of the porcine sequence (GenBank Acc. No. AY606802) when compared to the human sequence. Then, we quantitatively analyzed RANKL production in cell culture supernatants by an ELISA system evaluated for the detection of human soluble RANKL. Indeed, RANKL was expressed to a significantly higher degree in cultures treated with 1α,25(OH)2D3 when compared to control cultures without 1α,25(OH)2D3[192]. This finding implicates that porcine RANKL is upregulated through 1α,25(OH)2D3, as is known from non-porcine systems [194], and might therefore be the cytokine with the highest osteoclastogenic activity in pigs. Furthermore, recent experiments demonstrated that porcine osteoblasts predominantly expressed soluble and membrane-bound RANKL besides other cytokines found in the murine osteoblast such as IL-1, IL-6, and TNF-α [195].

In conclusion, the cytokine pattern of the porcine system reflects those found in human and murine bone marrow cell cultures. Additionally, our findings provide strong evidence of the existence of the RANK/RANKL/OPG system in pigs. Further experiments should clarify the impact that 1α,25(OH)2D3 exerts on cytokine production by peripheral lymphocytes in pigs.

Currently, we are investigating the role of lymphocyte subsets on porcine osteoclast generation – seemingly peripheral blood mononuclear cells in general and CD4+ T cells in particular exert an osteoclastogenic effect in osteoclastogenesis – and immunophenotypic and cytokinologic properties of porcine osteoblasts.

In the near future it should be possible to use the ongoing increasing understanding of the osteoimmunological principles of porcine bone marrow-derived cells also to replace rodent models for osteological research by porcine in vivo models at a higher rate than today. Glucocorticoid-induced osteoporosis in minipigs may serve as an example of a porcine in vivo model of osteoporosis [196].

Osteoporosis is defined as a skeletal disorder characterized by compromised bone strength predisposing a person to an increased risk of fracture. Bone strength primarily reflects the integration of bone density and bone quality [197]. Osteoporosis is among the most important conditions associated with aging; the lifetime risk for a fragility fracture (vertebral fracture (fig. 3), distal forearm fracture, hip fracture) in a 50-year-old white US woman is approximately 40%, whereas that in a white US man is 13% [198].

Fig. 3

Magnetic resonance images of the thoracic and lumbar spine. Magnetic resonance study of the thoracic and lumbar spine of a man with multiple vertebral fractures due to osteoporosis. Courtesy of Prof. Dr. H. Resch, Department of Medicine 2, St. Vincent Hospital, Vienna.

Fig. 3

Magnetic resonance images of the thoracic and lumbar spine. Magnetic resonance study of the thoracic and lumbar spine of a man with multiple vertebral fractures due to osteoporosis. Courtesy of Prof. Dr. H. Resch, Department of Medicine 2, St. Vincent Hospital, Vienna.

Close modal

From a theoretical point of view, osteoporosis results from any imbalance of bone turnover that results in an excess of osteoclast activity (bone resorption) over osteoblast activity (bone formation). Nevertheless, it should be borne in mind that in an individual, bone mass is determined by the amount of bone mass achieved at skeletal maturity (‘peak bone mass’) and the velocity of subsequent bone loss. The risk of osteoporosis is strongly influenced by genetic components [199]; further determinants are hormonal and nutritional factors as well as exercise [200]. Generally, osteoporosis is classified as either primary (idiopathic) or secondary. A variety of diseases (e.g. rheumatoid arthritis [201, 202]), medications (e.g. glucocorticoids or cyclosporin A) and conditions such as alcohol abuse can result in secondary osteoporosis.

In 1998, Riggs et al.[ 203] proposed a unitary model of primary (involutional) osteoporosis in postmenopausal women and aging men. As in the model of Riggs and Melton [204], the existence of an early rapid and a late slow phase of bone loss in women is emphasized; in contrast, in aging men there is only one slow phase of continuous bone loss. The early rapid phase of bone loss in women clearly results from postmenopausal estrogen deficiency. In the late, slow phase of bone loss, vitamin D deficiency and extraskeletal consequences of estrogen deficiency lead to an increase of serum PTH levels; secondary hyperparathyroidism further stimulates bone resorption [203].

As early as 1987, Pacifici et al. [205] reported that monocytes from patients with idiopathic osteoporosis produced significantly more IL-1 than those from control subjects. In subsequent work [206], the authors demonstrated an increased IL-1 production by monocytes from postmenopausal when compared to premenopausal women or estrogen/progesterone treated women. Whereas nonosteoporotic postmenopausal women achieved premenopausal IL-1 levels within 8 years of menopause, in osteoporotic subjects an elevated IL-1 production was seen as long as 15 years after menopause. Surgically induced menopause was associated with a production of IL-1, TNF-α and GM-CSF by mononuclear cells; in women who received estrogen replacement therapy, simultaneous decreases of cytokine secretion and bone resorption were noted [207]. Whereas Zarrabeitia et al. [208] did not detect abnormalities of cytokine production in osteoporosis, Zheng et al. [209] found an increased production of IL-1, IL-6 and TNF-α by whole blood cells from patients with postmenopausal osteoporosis. Taken together these data indicate that an increased production of mononuclear cell immune products contributes to the postmenopausal enhancement of bone resorption. In this context we should like to mention that the aging process is characterized by a progressive proinflammatory status, a phenomenon referred to as ‘inflamm-aging’ by Franceschi et al. [210].

In addition to changes of cytokine production by monocytes, T cell abnormalities have been reported in patients with osteoporosis. In 1984, Fujita et al. [211] described an increased CD4+/CD8+ ratio in osteoporosis; these findings were corroborated by Imai et al. [212] and Rosen et al. [213]. Hustmyer et al. [214] described a negative correlation between the CD8+/CD56+ subset and bone mineral density. Data from our laboratory indicate that in postmenopausal women with osteoporotic fractures the CD8+/CD57+ subset is expanded; moreover, in the fracture patients the percentage of CD8+ cells that expressed TNF-α was augmented [215]. Thus, in addition to monocytes and their products, T cells appear to contribute to the pathogenesis of primary osteoporosis.

Quite surprisingly there are only few studies on RANKL or osteoprotegerin in patients with osteoporosis. Eghbali-Fatourechi et al. [216] determined the surface expression of RANKL on bone marrow mononuclear cells in premenopausal, early postmenopausal and estrogen treated postmenopausal women by flow cytometry. The surface expression of RANKL on marrow stromal cells, B cells and T cells was significantly higher in early postmenopausal when compared to premenopausal or estrogen-treated women. These findings suggest that upregulation of RANKL on stromal cells and lymphocytes in the bone marrow could mediate increased bone resorption consecutive to estrogen deficiency. Whereas the study of Eghbali-Fatourechi et al. [216] refers to the early, rapid phase of postmenopausal bone loss there are data that indicate a role of the RANKL/OPG pathway also in fracture susceptibility: Abdallah et al. [217] demonstrated an increased RANKL/OPG mRNA ratio in bone biopsies from women with hip fractures. In contrast to studies on surface expression or mRNA levels of RANKL and OPG, the measurement of these markers in serum has produced somewhat paradoxical results. With regard to OPG, most studies found elevated OPG serum levels in patients with osteoporosis [218,219,220,221] whereas one study reported decreased OPG levels in osteoporotic patients with vertebral fractures [222]. Liu et al. [223] found no differences of serum OPG and RANKL levels as well as the RANKL/OPG ratio among normal, osteopenic and osteoporotic women. Nevertheless, Schett et al. [224] showed that low levels of RANKL are a predictor of an increased risk of nontraumatic fracture. While it is possible that at least some of these observations could represent compensatory mechanisms to counteract enhanced bone degradation, the clinical utility of serum RANKL and OPG measurements still requires further investigation [225]. Further evidence for the critical role of the RANKL/OPG pathway comes from an intervention trial: the administration of denosumab, a monoclonal antibody against RANKL, in postmenopausal women with low bone mass decreased bone resorption and increased bone mineral density [226]. Thus, RANKL appears to be a promising target for the treatment of osteoporosis.

Contributing studies to this review were funded in part as Profillinienprojekt of the University of Veterinary Medicine, Vienna.

1.
Aaron J, Choi Y: Bone versus immune system. Nature 2000;408:535–536.
2.
Rifas L, Arackal S, Weitzmann MN: Inflammatory T cells rapidly induce differentiation of human bone marrow stromal cells into mature osteoblasts. J Cell Biochem 2003;88:650–659.
3.
Kong YY, Feige U, Sarosi I, Bolon B, Tafuri A, Morony S, Capparelli C, Li J, Elliott R, McCabe S, Wong T, Campagnuolo G, Moran E, Bogoch ER, Van G, Nguyen LT, Ohashi PS, Lacey DL, Fish E, Boyle WJ, Penninger JM: Activated T cells regulate bone loss and joint destruction in adjuvant arthritis through osteoprotegerin ligand. Nature 1999;402:304–309.
4.
Dazzi F, Ramasamy R, Glennie S, Jones SP, Roberts I: The role of mesenchymal stem cells in haemopoiesis. Blood Rev 2006;20:161–171.
5.
Jung Y, Wang J, Havens A, Sun Y, Wang J, Jin T, Taichman RS: Cell-to-cell contact is critical for the survival of hematopoietic progenitor cells on osteoblasts. Cytokine 2005;32:155–162.
6.
Neiva K, Sun YX, Taichman RS: The role of osteoblasts in regulating hematopoietic stem cell activity and tumor metastasis. Braz J Med Biol Res 2005;38:1449–1454.
7.
Taichman RS, Emerson SG: The role of osteoblasts in the hematopoietic microenvironment. Stem Cells 1998;16:7–15.
8.
Taichman RS, Reilly MJ, Emerson SG: The hematopoietic microenvironment: osteoblasts and the hematopoietic microenvironment. Hematology 2000;4:421–426.
9.
Cornish J, Gillespie MT, Callon KE, Horwood NJ, Moseley JM, Reid IR: Interleukin-18 is a novel mitogen of osteogenic and chondrogenic cells. Endocrinology 2005;144:1194–1201.
10.
Mirosavljevic D, Quinn JM, Elliott J, Horwood NJ, Martin TJ, Gillespie MT: T-cells mediate an inhibitory effect of interleukin-4 on osteoclastogenesis. J Bone Miner Res 2003;18:984–993.
11.
Takayanagi H, Ogasawara K, Hida S, Chiba T, Murata S, Sato K, Takaoka A, Yokochi T, Oda H, Tanaka K, Nakamura K, Taniguchi T: T-cell-mediated regulation of osteoclastogenesis by signaling cross-talk between RANKL and IFN-gamma. Nature 2000;408:600–605.
12.
Konishi M, Takahashi K, Yoshimoto E, Uno K, Kasahara K, Mikasa K: Association between osteopenia/osteoporosis and the serum RANKL in HIV-infected patients. AIDS 2005;19:2040–2041.
13.
Kotake S, Udagawa N, Hakoda M, Mogi M, Yano K, Tsuda E, Takahashi K, Furuya T, Ishiyama S, Kim KJ, Saito S, Nishikawa T, Takahashi N, Togari A, Tomatsu T, Suda T, Kamatani N: Activated human T cells directly induce osteoclastogenesis from human monocytes: possible role of T cells in bone destruction in rheumatoid arthritis patients. Arthritis Rheum 2001;44:1003–1012.
14.
Mundy GR: Metastasis to bone: causes, consequences and therapeutic opportunities. Nat Rev Cancer 2002;2:584–593.
15.
Shigeyama Y, Pap T, Kunzler P, Simmen BR, Gay RE, Gay S: Expression of osteoclast differentiation factor in rheumatoid arthritis. Arthritis Rheum 2000;43:2523–2530.
16.
Stellon AJ, Davies A, Compston J, Williams R: Bone loss in autoimmune chronic active hepatitis on maintenance corticosteroid therapy. Gastroenterology 1985;89:1078–1083.
17.
Fujisaki K, Tanabe N, Suzuki N, Mitsui N, Oka H, Ito K, Maeno M: The effect of IL-1α on the expression of matrix metalloproteinases, plasminogen activators, and their inhibitors in osteoblastic ROS 17/2.8 cells. Life Sci 2006;78:1975–1982.
18.
Takahashi N, Katsu T, Udagawa N, Sasaki T, Yamaguchi A, Moseley JM, Martin TJ, Suda T: Osteoblastic cells are involved in osteoclast formation. Endocrinology 1988;123:2600–2602.
19.
Gori F, Hofbauer LC, Dunstan CR, Spelsberg TC, Khosla S, Riggs BL: The expression of osteoprotegerin and RANK ligand and the support of osteoclast formation by stromal-osteoblast lineage cells is developmentally regulated. Endocrinology 2000;141:4768–4777.
20.
Everts V, Delaisse JM, Korper W, Niehof A, Vaes G, Beertsen W: Degradation of collagen in the bone-resorbing compartment underlying the osteoclast involves both cysteine-proteinases and matrix metalloproteinases. J Cell Physiol 1992;150:221–231.
21.
Power J, Loveridge N, Rushton N, Parker M, Reeve J: Osteocyte density in aging subjects is enhanced in bone adjacent to remodeling haversian systems. Bone 2002;30:859–865.
22.
Yeni Y, Vashishth D, Fyhrie D: Estimation of bone matrix apparent stiffness variation caused by osteocyte lacunar size and density. J Biomech Eng 2001;123:10–17.
23.
Burger E, Klein-Nulend J: Mechanotransduction in bone-role of the lacuno-canalicular network. FASEB J 1999;13:S101–S112.
24.
Cowin SC: Mechanosensation and fluid transport in living bone. J Musculoskelet Neuronal Interact 2002;2:256–266.
25.
Beneyahu D: The hematopoietic microenvironment: the osteogenic compartment of bone marrow: cell biology and clinical applications. Hematology 2000;4:427–435.
26.
Aubin J: Regulation of osteoblast formation and function. Rev Endocrinol Metab Disord 2001;2:81–94.
27.
Liu F, Malaval L, Aubin J: Global amplification polymerase chain reaction reveals novel transitional stages during osteoprogenitor differentiation. J Cell Sci 2003;116:1787–1796.
28.
Ducy P, Zhang R, Geoffroy V, Ridall AL, Karsenty G: Osf2/cbfa1: a transcriptional activator of osteoblast differentiation. Cell 1997;89:747–754.
29.
Nakashima K, et al: The novel zinc finger-containing transcription factor osterix is required for osteoblast differentiation and bone formation. Cell 2002;108:17–29.
30.
Komori T, Yagi H, Nomura S, Yamaguchi A, Sasaki K, Deguchi K, Shimizu Y, Bronson RT, Gao YH, Inada M, Sato M, Okamoto R, Kitamura Y, Yoshiki S, Kishimoto S: Targeted disruption of Cbfa1 results in a complete lack of bone formation owing to maturational arrest of osteoblasts. Cell 1997;89:755–764.
31.
Otto F, Thornell AP, Crompton T, Denzel A, Gilmour KC, Rosewell LR, Stamp GWH, Beddington RSP, Mundlos S, Olsen BP, Selby PB, Owen MJ: Cbfa1, a candidate gene for cleidocranial dysplasia syndrome, is essential for osteoblast differentiation and bone development. Cell 1997;89:765–777.
32.
Jones D, Wein M, Glimcher L: Schnurri-3: a key regulator of postnatal skeletal remodeling. 1st Int Conf Osteoimmunology, 2006, A2.
33.
Marom R, Shur I, Solomon R, Benayahu D: Characterization of adhesion and differentiation markers of osteogenic marrow stromal cells. J Cell Physiol 2005;202:41–48.
34.
Noble BS, Peet N, Stevens HY, Brabbs A, Mosley JR, Reilly GC, Reeve J, Skerry TM, Lanyon LE: Mechanical loading: biphasic osteocyte survival and targeting of osteoclasts for bone destruction in rat cortical bone. Am J Physiol 2003;284:C934–C944.
35.
Rawlinson SC, Pitsillides AA, Lanyon LE: Involvement of different ion channels in osteoblasts’ and osteocytes’ early responses to mechanical strain. Bone 1996;19:609–614.
36.
Marroti G, Cane V, Palazzini S, Lalumbo C: Structure function relationships in the osteocyte. Miner Electrolyte Metab 1990;4:93–106.
37.
Mullender MG, Huiskes R: Proposal for the regulatory mechanism of Wolff’s law. J Orthop Res 2005;3:503–511.
38.
Gu G, Mulari M, Peng Z, Hentunen T, Väänänen HK: Death of osteocytes turns off the inhibition of osteoclasts and triggers local bone resorption 2005;335:1095–1101.
39.
Kogianni G, Mann V, Noble BS: Apoptotic osteocytes induce the production of pro-inflammatory osteoclastogenic cytokines. 1st Int Conf Osteoimmunology, 2006, A37.
40.
Miller SC, de Saint-Georges L, Bowman BM, Jee WS: Bone lining cells: structure and function. Scanning Microsc 1989;3:953–960.
41.
Udagawa N, Takahashi N, Akatsu T, Tanaka H, Sasaki T, Nishihara T, Koga T, Martin TJ, Suda T: Origin of osteoclasts: Mature monocytes and macrophages are capable of differentiating into osteoclasts under a suitable microenvironment prepared by bone marrow-derived stromal cells. Proc Natl Acad Sci USA 1990;87:7260–7264.
42.
Kurihara N, Chenu C, Miller M, Civin C, Roodman GD: Identification of committed mononuclear precursors for osteoclast-like cells formed in long term human marrow cultures. Endocrinology 1990;126:2733–2741.
43.
Kodama H, Nose M, Niida S, Yamasaki A: Essential role of macrophage colony-stimulating factor in the osteoclast differentiation supported by stromal cells. J Exp Med 1991;173:1291–1294.
44.
Jacquin C, Gran DE, Lee SK, Lorenzo JA, Aguila HL: Identification of multiple osteoclast precursor populations in murine bone marrow. J Bone Miner Res 2006;21:67–77.
45.
Luchin A, Purdom G, Murphy K, Clark MY, Angel N, Cassady AI, Hume DA, Ostrowski MC: The microphthalmia transcription factor regulates expression of the tartrate-resistant acid phosphatase gene during terminal differentiation of osteoclasts. J Bone Miner Res 2000;15:451–460.
46.
Hattersley G, Chambers TJ: Calcitonin receptors as markers for osteoclastic differentiation: correlation between generation of bone-resorptive cells and cells that express calcitonin receptors in mouse bone marrow cultures. Endocrinology 1989;125:1606–1612.
47.
Hansen T, Otto M, Gaumann A, Eckardt A, Petrow PK, Delank KS, Kirkpatrick CJ, Kriegsmann J: Cathepsin K in aseptic hip prosthesis loosening: expression in osteoclasts without polyethylene wear particles. J Rheumatol 2001;28:1615–1619.
48.
Teti A, Grano M, Carano A, Colucci S, Zambonin Zallone A: Immunolocalization of beta 3 subunit of integrins in osteoclast membrane. Boll Soc Ital Biol Sper 1989;65:1031–1037.
49.
Yoshida S, Domon T, Wakita M: Studies of the clear zone of osteoclasts: immunohistological aspects of its form and distribution. Arch Histol Cytol 1989;52:513–520.
50.
Salo J, Metsikko K, Palokangas H, Lehenkari P, Vaananen HK: Bone-resorbing osteoclasts reveal a dynamic division of basal plasma membrane into two different domains. J Cell Sci 1996;109:301–307.
51.
Salo J, Lehenkari P, Mulari M, Metsikko K, Vaananen HK: Removal of osteoclast bone resorption products by transcytosis. Science 1997;276:270–273.
52.
Baron R, Neff L, Louvard D, Courtoy PJ: Cell-mediated extracellular acidification and bone resorption: Evidence for a low pH in resorbing lacunae and localization of a 100-kD lysosomal membrane protein at the osteoclast ruffled border. J Cell Biol 1985;101:2210–2222.
53.
Teti A, Blair HC, Schlesinger P, Grano M, Zambonin-Zallone A, Kahn AJ, Teitelbaum SL, Hruska KA: Extracellular protons acidify osteoclasts, reduce cytosolic calcium, and promote expression of cell-matrix attachment structures. J Clin Invest 1989;84:773–780.
54.
Delaisse JM, Andersen TL, Engsig MT, Henriksen K, Troen T, Blavier L: Matrix metalloproteinases (MMP) and cathepsin K contribute differently to osteoclastic activities. Microsc Res Tech 2000;61:504–513.
55.
Wucherpfennig AL, Li YP, Stetler-Stevenson WG, Rosenberg AE, Stashenko P: Expression of 92 kD type IV collagenase/gelatinase B in human osteoclasts. J Bone Miner Res 1994;9:549–556.
56.
Okada Y, Naka K, Kawamura K, Matsumoto T, Nakanishi I, Fujimoto N, Sato H, Seiki M: Localization of matrix metalloproteinase 9 (92-kilodalton gelatinase/type IV collagenase = gelatinase B) in osteoclasts: implications for bone resorption. Lab Invest 1995;72:311–322.
57.
Wolff J: Das Gesetz der Transformation des Knochens. Berlin, August Hirschwald, 1892.
58.
Sommerfeldt DW, Rubin CT: Biology of bone and how it orchestrates the form and function of the skeleton. Eur Spine J 2001;S86–S95.
59.
Frost HM: Skeletal structural adaptations to mechanical usage (SATMU). 1. Redefining Wolff’s law: the bone modeling problem. Anat Rec 1990;226:403–413.
60.
Frost HM: Skeletal structural adaptions to mechanical usage (SATMU): 2. Redefining Wolff’s law: the remodeling problem. Anat Rec 1990;226:414–422.
61.
Rodan GA, Martin TJ: Role of osteoblasts in hormonal control of bone resorption – a hypothesis. Calc Tissue Int 1981;33:349–351.
62.
Parfitt AM: Osteonal and hemi-osteonal remodeling: the spatial and temporal framework for signal traffic in adult human bone. J Cell Biochem 1994;55:273–286.
63.
Raisz LG: Local and systemic factors in the pathogenesis of osteoporosis. N Engl J Med 1988;318:818–828.
64.
Parfitt AM: Pharmacological manipulations of bone remodelling in calcium homeostasis; in Kanis (ed): Progress in Basic and Clinical Pharmacology. Basel, Karger, 1990, pp 1–27.
65.
Hauge EM, Qvesel D, Erikson EF, Moselkilde L, Melsen F: Cancellous bone remodeling occurs in specialized compartments lined by cells expressing osteoblastic markers. J Bone Miner Res 2001;16:1575–1582.
66.
Perez-Amodio S, Beertsen W, Everts V: (Pre-)osteoclasts induce retraction of osteoblasts before their fusion to osteoclasts. J Bone Miner Res 2004;19:1722–1731.
67.
Pilbeam CC, Harrison JR, Raisz LG: Prostaglandins and bone metabolism; in Bilezikian JP, Raisz LG, Rodan GA (eds): Principles of Bone Biology. San Diego, Academic Press, 2002, pp 979–994.
68.
Ha H, Lee JH, Kim HN, Kim HM, Kwak HB, Lee S, Kim HH, Lee ZH: alpha-Lipoic acid inhibits inflammatory bone resorption by suppressing prostaglandin E2 synthesis. J Immunol 2006;176:111–117.
69.
Hiraga T, Myoui A, Choi ME, Yoshikawa H, Yoneda T: Stimulation of cyclooxygenase-2 expression by bone-derived transforming growth factor-{beta} enhances bone metastases in breast cancer. Cancer Res 2006;66:2067–2073.
70.
Gregory LS, Kelly WL, Reid RC, Fairlie DP, Forwood MR: Inhibitors of cyclo-oxygenase-2 and secretory phospholipase A2 preserve bone architecture following ovariectomy in adult rats. Bone 2006;39:134–142.
71.
Shoji M, Tanabe N, Mitsui N, Tanaka H, Suzuki N, Takeichi O, Sugaya A, Maeno M: Lipopolysaccharide stimulates the production of prostaglandin E(2) and the receptor Ep4 in osteoblasts. Life Sci 2006;78:2012–2018.
72.
Jiang J, Lv HS, Lin JH, Jiang DF, Chen ZK: LTB4 can directly stimulate human osteoclast formation from PBMC independent of RANKL. Artif Cells Blood Substit Immobil Biotechnol 2005;33:391–403.
73.
Anderson GI, MacQuarrie R, Osinga C, Chen YF, Langman M, Gilbert R: Inhibition of leukotriene function can modulate particulate-induced changes in bone cell differentiation and activity. J Biomed Mater Res 2001;58:406–411.
74.
Traianedes K, Dallas MR, Garrett IR, Mundy GR, Bonewald LF: 5-Lipoxygenase metabolites inhibit bone formation in vitro. Endocrinology 1998;139:3178–3184.
75.
Kronenberg H, Bringhurst F, Nussbaum S, Jüppner H, Abou-Samra A, Segre G, Potts J: Parathyroid hormone: biosynthesis, secretion, chemistry, and action; in Mundy G, Martin J (eds): Handbook of Experimental Pharmacology: Physiology and Pharmacology of Bone. Heidelberg, Springer, 1993, pp 185–201.
76.
Potts J, Jüppner H: Parathyroid hormone and parathyroid hormone-related peptide in calcium homeostasis, bone metabolism, and bone development: the proteins, their genes, and receptors; in Avioli L, Krane S (eds): Metabolic Bone Disease. New York, Academic Press, 1997, pp 51–94.
77.
Dai J, He P, Chen X, Greenfield E: TNFa and PTH utilize distinct mechanisms to induce IL-6 and RANKL expression with markedly different kinetics. Bone 2006;38:509–520.
78.
Feyen J, Eldorf P, Padova F, Trechsel U: Interleukin-6 is produced by bone and modulated by parathyroid hormone. J Bone Miner Res 1989;4:633–638.
79.
Greenfield E, Horowitz M, Lavish S: Stimulation by parathyroid hormone of interleukin-6 and leukemia inhibitory factor expression in osteoblasts is an immediate-early gene response induced by cAMP signal transduction. J Biol Chem 1996;271:10984–10989.
80.
Greenfield E, Gornik S, Horowitz M, Donahue H, Shaw S: Regulation of cytokine expression in osteoblasts by parathyroid hormone: rapid stimulation of interleukin-6 and leukemia inhibitory factor mRNA. J Bone Miner Res 1993;8:1163–1171.
81.
Horowitz M, Brown M, Insogna K, Coleman D, Centrella M, Phillips J, et al: PTHrP and PTH induce the secretion of IL-6 by a clonal osteosarcoma cell line; in Oppenheim KK (ed): Molecular and Cellular Biology of Cytokines. New York, Wiley-Liss, 1990, pp 471–476.
82.
Huang Y, Harrison J, Lorenzo J, Kream B: Parathyroid hormone induces interleukin-6 heterogeneous nuclear and messenger RNA expression in murine calvarial organ cultures. Bone 1998;23:327–332.
83.
Onyia J, Libermann T, Bidwell J, Arnold D, Tu Y, McClelland P, et al: Parathyroid hormone (1–34)-mediated interleukin-6 induction. J Cell Biochem 1997;67:265–274.
84.
Pollock J, Blaha M, Gornik S, Stevenson S, Greenfield E: In vivo demonstration that parathyroid hormone and parathyroid hormone-related protein stimulate expression by osteoblasts of interleukin-6 and leukemia inhibitory factor. J Bone Miner Res 1996;11:754–759.
85.
Panda DK, Miao D, Bolivar I, Li J, Huo R, Hendy GN, Glotzman D: Inactivation of the 25-hydroxylase and vitamin D receptor demonstrates independent and interdependent effects of calcium and vitamin D on skeletal and mineral homeostasis. J Biol Chem 2004;179:16754–16766.
86.
Dardenne O, Prud’homme J, Arabian A, Glorieux R, St-Arnaud R: Targeted inactivation of the 25-hydroxyvitamin D(3)-1(alpha)-hydroxylase gene (CYP27B1) creates an animal model of pseudovitamin D-deficiency rickets. Endocrinology 2001;142:3135–3141.
87.
Li YC, Pirro AE, Amling M, Delling R, Baron R, Bronson R, Demay MB: Targeted ablation of the vitamin D receptor: an animal model of vitamin D-dependent rickets type II with alopecia. Proc Natl Acad Sci USA 1997;94:9831–9835.
88.
Kitazawa S, Kajimoto K, Kondo T, Kitazawa R: Vitamin D3 supports osteoclastogenesis via functional vitamin D response element of human RANKL gene promotor. J Cell Biochem 2003;89:771–777.
89.
Khosla S, Melton LJ, Riggs B: Estrogen and the male skeleton. J Clin Endocrinol Metab 2002;87:1443–1450.
90.
Carani C, Qin K, Simoni M, Faustini-Fustini M, Serpente S, Boyd J, Korach K, Simpson E: Effect of testosterone and estradiol in a man with aromatase deficiency. N Engl J Med 1997;337:91–95.
91.
Leder B, LeBlanc K, Schoenfeld D, Eastell R, Finkelstein J: Differential effects of androgens and estrogens on bone turnover in normal men. J Clin Endocrinol Metab 2003;88:204–210.
92.
Hofbauer LC, Hicok K, Chen S, Khosla S: Regulation of osteoprotegerin production by androgens and anti-androgens in human osteoblastic lineage cells. Eur J Endocrinol 2002;147:269–273.
93.
Hofbauer LC, Khosla S, Dunstan C, Lacey D, Spelsberg T, Riggs B: Estrogen stimulates gene expression and protein production of osteoprotegerin in human osteoblastic cells. Endocrinology 1999;140:4367–4370.
94.
Jilka RL, Weinstein RS, Bellido T, Parfitt AM, Manolagas SC: Osteoblast programmed cell death (apoptosis): modulation by growth factors and cytokine. J Bone Miner Res 1998;13:793–802.
95.
Eghbali-Fatourechi G, Khosla S, Sanyal A, Boyle W, Lacey D, Riggs B: Role of RANK ligand in mediating increased bone resorption in early postmenopausal women. J Clin Invest 2003;111:1221–1230.
96.
Pacifici R, Carano A, Santoro SA, Rifas L, Jeffrey JJ, Malone JD, McCracken R, Avioli LV: Bone matrix constituents stimulate interleukin-1 release from human blood mononuclear cells. J Clin Invest 1991;87:221–228.
97.
Tanaka S, Takahashi N, Udagawa N, Tamura T, Akatsu T, Stanely E, Kurokawa T, Suda T: Macrophage colony-stimulating factor is indispensable for both proliferation and differentiation of osteoclast progenitors. J Clin Invest 1993;91:257–263.
98.
Manolagas S, Jilka R: Bone marrow, cytokines, and bones remodeling: emerging insights into the pathophysiology of osteoporosis. N Engl J Med 1995;332:305–311.
99.
Jilka RL, Passeri G, Girasole G, Cooper S, Abrams J, Broxmeyer H, Manolagas SC: Estrogen loss upregulates hematopoiesis in the mouse: a mediating role of IL-6. Exp Hematol 1995;23:500–506.
100.
Jilka RL, Hangoc G, Girasole G, Passeri G, Williams DC, Abrams JS, Boyce B, Broxmeyer H, Manolagas SC: Increased osteoclast development after estrogen loss: mediation by interleukin-6. Science 1992;257:88–91.
101.
Oursler MJ, Cortese C, Keeting PE, Anderson MA, Bonde SK, Riggs BL, Spelsberg TC: Modulation of transforming growth factor-beta production in normal human osteblast-like cells by 17β-estrodiol and parathyroid hormone. Endocrinology 1991;129:3313–3320.
102.
Hughes DE, Dai A, Tiffee JC, Li HH, Mundy GR, Boyce BF: Estrogen promotes apoptosis of murine osteoclasts mediated by TGF-beta. Nat Med 1996;2:1132–1136.
103.
Fox SW, Lovibond AC: Current insights into the role of transforming growth factor-beta in bone resorption. Mol Cell Endocrinol 2005;243:19–26.
104.
Simonet WS, Lacey DL, Dunstan CR, Kelley M, Chang MS, Luthy R, Nguyen HQ, Wooden S, Bennett L, Boone T, Shimamoto G, DeRose M, Elliott R, Colombero A, Tan HL, Trail G, Sullivan J, Davey E, Bucay N, Renshaw-Gregg L, Hughes TM, Hill D, Pattison W, Campell P, Boyle WJ: Osteoprotegerin: a novel secreted protein involved in regulation of bone density. Cell 1997;89:309–319.
105.
Yasuda H, Shima N, Nakagawa N, Mochizuki SI, Yano K, Fujise N, Sato Y, Goto M, Yamaguchi K, Kuriyama M, Kanno T, Murakami A, Tsuda E, Morinaga T, Higashio K: Identity of osteoclastogenesis inhibitory factor (OCIF) and osteoprotegerin (OPG): a mechanism by which OPG/OCIF inhibits osteoclastogenesis in vitro. Endocrinology 1998;139:1329–1337.
106.
Yasuda H, Shima N, Nakagawa N, Yamaguchi K, Kinosaki M, Mochizuki S, Tomoyasu A, Yano K, Goto M, Murakami A, Tsuda E, Morinaga T, Higashio K, Udagawa N, Takahashi N, Suda T: Osteoclast differentiation factor is a ligand for osteoprotegerin/osteoclastogenesis-inhibitory factor and is identical to TRANCE/RANKL. Proc Natl Acad Sci USA 1998;95:3597–3602.
107.
Yun TJ, Chaudhary PM, Shu GL, Frazer JK, Ewings MK, Schwartz SM, Pascual V, Hood LE, Clark EA: OPG/FDCR-1, a TNF receptor family member, is expressed in lymphoid cells and is up-regulated by ligating CD40. J Immunol 1998;161:6113–6121.
108.
Kwon BS, Wang S, Udagawa N, Haridas V, Lee ZH, Kim KK, Oh KO, Greene J, Li Y, Su J, Gentz R, Aggarwal BB, Ni J: TR1, a new member of the tumor necrosis factor receptor superfamily, induces fibroblast proliferation and inhibits osteoclastogenesis and bone resorption. FASEB J 1998;12:845–854.
109.
Hofbauer LC, Shui C, Riggs BL, Dunstan CR, Spelsberg TC, O’Brien T, Khosla S: Effects of immunosuppressants on receptor activator of NF-kappaB ligand and osteoprotegerin production by human osteoblastic and coronary artery smooth muscle cells. Biochem Biophys Res Commun 2001;280:334–339.
110.
Srivastava S, Matsuda M, Hou Z, Bailey JP, Kitazawa R, Herbst MP, Horseman ND: Receptor activator of NF-kappaB ligand induction via Jak2 and Stat5a in mammary epithelial cells. J Biol Chem 2003;278:46171–46178.
111.
Cross SS, Yang Z, Brown NJ, Balasubramanian SP, Evans CA, Woodward JK, Neville-Webbe HL, Lippitt JM, Reed MW, Coleman RE, Holen I: Osteoprotegerin (OPG)-a potential new role in the regulation of endothelial cell phenotype and tumour angiogenesis? Int J Cancer 2006;118:1901–1908.
112.
Malyankar UM, Scatena M, Suchland KL, Yun TJ, Clark EA, Giachelli CM: Osteoprotegerin is an alpha vs. beta 3-induced, NF-kappa B-dependent survival factor for endothelial cells. J Biol Chem 2000;275:20959–20962.
113.
Al-Fakhri N, Hofbauer LC, Preissner KT, Franke FE, Schoppet M: Expression of bone-regulating factors osteoprotegerin (OPG) and receptor activator of NF-kappaB ligand (RANKL) in heterotopic vascular ossification. Thromb Haemost 2005;94:1335–1337.
114.
Rasmussen LM, Tarnow L, Hansen TK, Parving HH, Flyvbjerg A: Plasma osteoprotegerin levels are associated with glycaemic status, systolic blood pressure, kidney function and cardiovascular morbidity in type 1 diabetic patients. Eur J Endocrinol 2006;154:75–81.
115.
Bucay N, Sarosi I, Dunstan CR, Morony S, Tarpley J, Capparelli C, Scully S, Tan HL, Xu W, Lacey DL, Boyle WJ, Simonet WS: osteoprotegerin-deficient mice develop early onset osteoporosis and arterial calcification. Genes Dev 1998;12:1260–1268.
116.
Wong BR, Rho J, Arron J, Robinson E, Orlinick J, Chao M, Kalachikov E, Bartlett FS, Frankel WN, Lee SY, Choi Y: TRANCE is a novel ligand for the tumor necrosis factor receptor family that activates c-Jun N-terminal kinase in T cells. J Biol Chem 1997;272:25190–25194.
117.
Lacey DL, Timms E, Tan HL, Kelley MJ, Dunstan CR, Burgess T, Elliott R, Colombero A, Elliott G, Scully S, Hsu H, Sullivan J, Hawkins N, Davy E, Capparelli C, Eli A, Qian YX, Kaufman S, Sarosi I, Shalhoub V, Senaldi G, Guo J, Delaney J, Boyle WJ: Osteoprotegerin ligand is a cytokine that regulates osteoclast differentiation and activation. Cell 1998;93:165–178.
118.
Anderson DM, Maraskovosky E, Billingsley WL, Dougall WC, Temetsko ME, Roux ER, Teepe MC, Dubose RF, Cosman D, Galibert L: A homologue of the TNF receptor and its ligand enhance T-cell growth and dendritic-cell function. Nature 1997;390:175–179.
119.
Kartsogiannis V, Zhou H, Horwood NJ, Thomas RJ, Hards DK, Quinn JM, Niforas P, Ng KW, Martin TJ, Gillespie MT: Localization of RANKL (receptor activator of NF kappa B ligand) mRNA and protein in skeletal and extraskeletal tissues. Bone 1999;25:525–534.
120.
Theoleyre S, Wittrant Y, Tat SK, Fortun Y, Redini F, Heymann D: The molecular triad OPG/RANK/RANKL: involvement in the orchestration of pathophysiological bone remodeling. Cytokine Growth Factor Rev 2004;15:457–475.
121.
Wong BR, Besser D, Kim N, Arron JR, Vologodskaia M, Hanafusa H, et al: TRANCE, a TNF family member, activates Akt/PKB through a signaling complex involving TRAF6 and c-src. Mol Cell 1999;4:1041–1049.
122.
Martin TJ, Gillespie MT: Receptor activator of nuclear factor kappa B ligand (RANKL): another link between breast and bone. Trends Endocrinol Metab 2001;12:2–4.
123.
Jones DH, Nakashima T, Sanchez OH, Kozieradzki I, Komarova SV, Sarosi I, Morony S, Rubin E, Sarao R, Hojilla CV, Komnenovic V, Kong YY, Schreiber M, Dixon SJ, Sims SM, Khokha R, Wada T, Penninger JM: Regulation of cancer cell migration and bone metastasis by RANKL. Nature 2006;440:692–696.
124.
Collin-Osdoby P, Rothe L, Anderson F, Nelson M, Maloney W, Osdoby P: Receptor activator of NF-kappa B and osteoprotegerin expression by human microvascular endothelial cells, regulation by inflammatory cytokines, and role in human osteoclastogenesis. J Biol Chem 2001;276:20659–20672.
125.
Suzuki J, Ikeda T, Kuroyama H, Seki S, Kasai M, Utusayama M, Tatsumi M, Uematsu H, Hirokawa K: Regulation of osteoclastogenesis by three human RANKL isoforms expressed in NIH3T3 cells. Biochem Biophys Res Commun 2004;314:1021–1027.
126.
Ikeda T, Kasai M, Utsuyama M, Hirokawa K: Determination of three isoforms of the receptor activator of nuclear factor-kappaB ligand and their differential expression in bone and thymus. Endocrinology 2001;142:1419–1426.
127.
Lum L, Wong BR, Josien R, Becherer JD, Erdjument-Bromage H, Schlondorff J, Tempst P, Choi Y, Blobel CP: Evidence for a role of a tumor necrosis factor-alpha (TNF-alpha)-converting enzyme-like protease in shedding of TRANCE, a TNF family member involved in osteoclastogenesis and dendritic cell survival. J Biol Chem 1999;274:13613–13618.
128.
Kanamaru F, Iwai H, Ikeda T, Nakajima A, Ishikawa I, Azuma M: Expression of membrane-bound and soluble receptor activator of NF-κB ligand (RANKL) in human T-cells. Immunol Lett 2004;94:239–246.
129.
Kong YY, Yoshida H, Sarosi I, Tan HL, Timms E, Capparelli C, Morony S, Oliveira-dos-Santos AJ, Van G, Itie A, Khoo W, Wakeham A, Dunstan CR, Lacey DL, Mak TW, Boyle WJ, Penninger JM: OPGL is a key regulator of osteoclastogenesis, lymphocyte development and lymph-node organogenesis. Nature 1999;397:315–323.
130.
Khosla S: Minireview: the OPG/RANKL/RANK system. Endocrinology 2001;142:5050–5055.
131.
Nakagawa N, Kinosaki M, Yamaguchi K, Shima N, Yasuda H, Yano K, Morinaga T, Higashio K: RANK is the essential signaling receptor for osteoclast differentiation factor in osteoclastogenesis. Biochem Biophys Res Commun 1998;253:395–400.
132.
Dougall WC, Glaccum M, Charrier K, Rohrbach K, Brasel K, De Smedt T, Daro E, Smith J, Tometsko ME, Maliszewski CR, Armstrong A, Shen V, Bain S, Cosman D, Anderson D, Morrissey PJ, Peschon JJ, Schuh J: RANK is essential for osteoclast and lymph node development. Genes Dev 1999;13:2412–2424.
133.
Kim HH, Lee DE, Shin JN, Lee YS, Jeon YM, Chung CH, Ni J, Kwon BS, Lee ZH: Receptor activator of NF-kappaB recruits multiple TRAF family adaptors and activates c-Jun N-terminal kinase. FEBS Lett 1999;443:297–302.
134.
Armstrong AP, Tometsko ME, Glaccum M, Sutherland CL, Cosman D, Dougall WC: A RANK/TRAF6-dependent signal transduction pathway is essential for osteoclast cytoskeletal organization and resorptive function. J Biol Chem 2002;277:44347–44355.
135.
Naito A, Yoshida H, Nishioka E, Satoh M, Azuma S, Yamamoto T, Nishikawa S, Inoue J: TRAF6-deficient mice display hypohydrotic ectodermal dysplasia. Proc Natl Acad Sci USA 2002;99:8766–8771.
136.
Lomaga MA, Yeh WC, Sarosi I, Duncan GS, Furlonger C, Ho A, Morony S, Capparelli C, Van G, Kaufman S, van der Heiden A, Itie A, Wakeham A, Khoo W, Sasaki T, Cao Z, Penninger JM, Paige CJ, Lacey DL, Dunstan CR, Boyle WJ, Goeddel DV, Mak TW: TRAF6 deficiency results in osteopetrosis and defective interleukin-1, CD40, and LPS signaling. Genes Dev 1999;13:1015–1024.
137.
Kobayashi N, Kadono Y, Naito A, Matsumoto K, Yamamoto T, Tanaka S, Inoue J: Segregation of TRAF6-mediated signaling pathways clarifies its role in osteoclastogenesis. EMBO J 2001;20:1271–1280.
138.
Ye H, Arron JR, Lamothe B, Cirilli M, Kobayashi T, Shevde NK, Segal D, Dzivenu OK, Vologodskaia M, Yim M, Du K, Singh S, Pike JW, Darnay BG, Choi Y, Wu H: Distinct molecular mechanism for initiating TRAF6 signalling. Nature 2002;418:443–447.
139.
Franzoso G, Carlson L, Xing L, Poljak L, Shores EW, Brown KD, Leonardi A, Tran T, Boyce BF, Siebenlist U: Requirement for NF-kappaB in osteoclast and B-cell development. Genes Dev 1997;11:3482–3496.
140.
Xing L, Bushnell TP, Carlson L, Tai Z, Tondravi M, Siebenlist U, Young F, Boyce BF: NF-kappaB p50 and p52 expression is not required for RANK-expressing osteoclast progenitor formation but is essential for RANK- and cytokine-mediated osteoclastogenesis. J Bone Miner Res 2002;17:1200–1210.
141.
Grigoriadis AE, Wang ZQ, Cecchini MG, Hofstetter W, Felix R, Fleisch HA, Wagner EF: c-Fos: a key regulator of osteoclast-macrophage lineage determination and bone remodeling. Science 1994;266:443–448.
142.
Ruocco MG, Maeda S, Park JM, Lawrence T, Hsu LC, Cao Y, Schett G, Wagner EF, Karin M: I{kappa}B kinase (IKK){beta}, but not IKK{alpha}, is a critical mediator of osteoclast survival and is required for inflammation-induced bone loss. J Exp Med 2005;201:1677–1687.
143.
Chaisson ML, Branstetter DG, Derry JM, Armstrong AP, Tometsko ME, Takeda K, Akira S, Dougall WC: Osteoclast differentiation is impaired in the absence of inhibitor of kappa B kinase alpha. J Biol Chem 2004;279:54841–54848.
144.
David JP, Sabapathy K, Hoffmann O, Idarraga MH, Wagner EF: JNK1 modulates osteoclastogenesis through both c-Jun phosphorylation-dependent and -independent mechanisms. J Cell Sci 2002;115:4317–4325.
145.
Wei S, Teitelbaum SL, Wang MW, Ross FP: Receptor activator of nuclear factor-kappa b ligand activates nuclear factor-kappa B in osteoclast precursors. Endocrinology 2001;142:1290–1295.
146.
Iotsova V, Caamano J, Loy J, Yang Y, Lewin A, Bravo R: Osteopetrosis in mice lacking NF-kappaB1 and NF-kappaB2. Nat Med 1997;3:1285–1289.
147.
Lee SW, Han SI, Kim HH, Lee ZH: TAK1-dependent activation of AP-1 and c-Jun N-terminal kinase by receptor activator of NF-kappaB. J Biochem Mol Biol 2002;35:371–376.
148.
Shim JH, Xiao C, Paschal AE, Bailey ST, Rao P, Hayden MS, Lee KY, Bussey C, Steckel M, Tanaka N, Yamada G, Akira S, Matsumoto K, Ghosh S: TAK1, but not TAB1 or TAB2, plays an essential role in multiple signaling pathways in vivo. Genes Dev 2005;19:2668–2681.
149.
Zhou G, Lee SC, Yao Z, Tan TH: Hematopoietic progenitor kinase 1 is a component of transforming growth factor beta-induced c-Jun N-terminal kinase signaling cascade. J Biol Chem 1999;274:13133–13138.
150.
Li X, Udagawa N, Takami M, Sato N, Kobayashi Y, Takahashi N: p38 Mitogen-activated protein kinase is crucially involved in osteoclast differentiation but not in cytokine production, phagocytosis, or dendritic cell differentiation of bone marrow macrophages. Endocrinology 2003;144:4999–5005.
151.
Matsumoto M, Sudo T, Saito T, Osada H, Tsujimoto M: Involvement of p38 mitogen-activated protein kinase signaling pathway in osteoclastogenesis mediated by receptor activator of NF-kappa B ligand (RANKL). J Biol Chem 2000;275:31155–31161.
152.
Mansky KC, Sankar U, Han J, Ostrowski MC: Microophthalmia transcription factor is a target of the p38 MAPK pathway in response to receptor activator of NF-kappa B ligand signaling. J Biol Chem 2002;277:11077–11083.
153.
Hotokezaka H, Sakai E, Kanaoka K, Saito K, Matsuo K, Kitaura H, Yoshida N, Nakayama K: U0126 and PD98059, specific inhibitors of MEK, accelerate differentiation of RAW264.7 cells into osteoclast-like cells. J Biol Chem 2002;277:47366–47372.
154.
Gingery A, Bradley E, Shaw A, Oursler MJ: Phosphatidylinositol 3-kinase coordinately activates the MEK/ERK and AKT/NFkappaB pathways to maintain osteoclast survival. J Cell Biochem 2003;89:165–179.
155.
Liu W, Wang S, Sun SW, Feng X: Receptor activator of NF-κB (RANK) cytoplasmic motif, 369PFQEP373, plays a predominant role in osteoclast survival in part by activating Akt/PKB and its downstream effector AFX/FOXO4. J Biol Chem 2005;280:43064.
156.
Seales EC, Micoli KJ, McDonald JM: Calmodulin is a critical regulator of osteoclastic differentiation, function, and survival. J Cell Biochem 2006;97:45–55.
157.
Ikeda F, Nishimura R, Matsubara T, Tanaka S, Inoue J, Reddy SV, Hata K, Yamashita K, Hiraga T, Watanabe T, Kukita T, Yoshioka K, Rao A, Yoneda T: Critical roles of c-Jun signaling in regulation of NFAT family and RANKL-regulated osteoclast differentiation. J Clin Invest 2004;114:475–484.
158.
Takayanagi H, Kim S, Koga T, Nishina H, Isshiki M, Yoshida H, Saiura A, Isobe M, Yokochi T, Inoue J, Wagner EF, Mak TW, Kodama T, Taniguchi T: Induction and activation of the transcription factor NFATc1 (NFAT2) integrate RANKL signaling in terminal differentiation of osteoclasts. Dev Cell 2002;3:889–901.
159.
Kudo O, Fujikawa Y, Itonaga I, Sabokbar A, Torisu T, Athanasou NA: Proinflammatory cytokine (TNF/IL-1) induction of human osteoclast formation. J Pathol 2002;198:220–227.
160.
Palmqvist P, Persson E, Conaway HH, Lerner UH: IL-6, leukemia inhibitory factor, and oncostatin M stimulate bone resorption and regulate the expression of receptor activator of NFκB ligand, osteoprotegerin, and receptor activator of NFκB in mouse calvariae. J Immunol 2002;169:3353–3362.
161.
Sabokbar A, Kudo O, Athanasou NA: Two distinct cellular mechanisms of osteoclast formation and bone resorption in periprosthetic osteolysis. J Orthopaed Res 2003;21:73–80.
162.
Toraldo G, Roggia C, Qian WP, Pacifici R, Weitzmann MN: IL-7 induces bone loss in vivo by induction of receptor activator of nuclear factor κ B ligand and tumor necrosis factor α from T cells. Proc Natl Acad Sci USA 2003;100:125–130.
163.
Roodman GD: Cell biology of the osteoclast. Exp Hematol 1999;27:1229–1241.
164.
Bendixen AC, Shevde NK, Dienger KM, Willson TM, Funk CD, Pike JW: IL-4 inhibits osteoclast formation through a direct action on osteoclast precursors via peroxisome proliferator-activated receptor γ 1. Proc Natl Acad Sci USA 2001;98:2443–2448.
165.
Fox SW, Haque SJ, Lovibond AC, Chambers TJ: The possible role of TGF-β-induced suppressors of cytokine signaling expression in osteoclast/macrophage lineage commitment in vitro. J Immunol 2003;170:3679–3687.
166.
Horwood NJ, Elliott J, Martin TJ, Gillespie MT: IL-12 alone and in synergy with IL-18 inhibits osteoclast formation in vitro. J Immunol 2001;166:4915–4921.
167.
Huang W, O’Keefe RJ, Schwarz EM: Exposure to receptor-activator of NFκB ligand renders pre-osteoclasts resistant to IFN-γ by inducing terminal differentiation. Arthritis Res Ther 2003;5:49–59.
168.
Khapli SM, Mangashetti LS, Yogesha SD, Wani MR: IL-3 acts directly on osteoclast precursors and irreversibly inhibits receptor activator of NFκB ligand-induced osteoclast differentiation by diverting the cells to macrophage lineage. J Immunol 2003;171:142–151.
169.
Quinn JM, Itoh K, Udagawa N, Hausler K, Yasuda H, Shima N, Mizuno A, Higashio K, Takahashi N, Suda T, Martin TJ, Gillespie MT: Transforming growth factor β affects osteoclast differentiation via direct and indirect actions. J Bone Miner Res 2001;16:1787–1794.
170.
Takayanagi H, Sato K, Takaoka A, Taniguchi T: Interplay between interferon and other cytokine systems in bone metabolism. Immunol Rev 2005;208:181–193.
171.
Bordin L, Priante G, Musacchio E, Giunco S, Tibaldi E, Clari G, Baggio B: Arachidonic acid-induced IL-6 expression is mediated by PKC α activation in osteoblastic cells. Biochemistry 2003;42:4485–4491.
172.
Thomson BM, Bennett J, Dean V, Triffitt J, Meikle MC, Loveridge N: Preliminary characterization of porcine bone marrow stromal cells: skeletogenic potential, colony-forming activity, and response to dexamethasone, transforming growth factor β, and basic fibroblast growth factor. J Bone Miner Res 1993;8:1173–1183.
173.
Mackie EJ: Cells in focus – Osteoblasts: novel roles in orchestration of skeletal architecture. Int J Biochem Cell Biol 2003;35:1301–1305.
174.
Jang JH, Kim JH: Improved cellular response of osteoblast cells using recombinant human osteopontin protein produced by Escherichia coli. Biotechnol Lett 2005;27:1767–1770.
175.
Gruber R, Kandler B, Fürst G, Fischer MB, Watzek G: Porcine sinus mucosa holds cells that respond to bone morphogenetic protein (BMP)-6 and BMP-7 with increased osteogenic differentiation in vitro. Clin Oral Implants Res 2004;15:575–580.
176.
Yates KE, Troulis MJ, Kaban LB, Glowacki J: IGF-1, TGF-β, and BMP-4 are expressed during distraction osteogenesis of the pig mandible. Int J Oral Maxillofac Surg 2002;31:173–178.
177.
Friedman MS, Long MW, Hankenson KD: Osteogenic differentiation of human mesenchymal stem cells is regulated by bone morphogenetic protein-6. J Cell Biochem 2006;98:538–554.
178.
Compston J: Osteoporosis in inflammatory bowel disease. Gut 2003;52:63–64.
179.
Grčeviç D, Sun-Kyeong L, Marušiç A, Lorenzo JA: Depletion of CD4 and CD8 T lymphocytes in mice in vivo enhances 1,25-dihydroxyvitamin D3-stimulated osteoclast-like cell formation in vitro by a mechanism that is dependent on prostaglandin synthesis. J Immunol 2000;165:4231–4238.
180.
Shinoda K, Sugiyama E, Taki H, Harada S, Mino T, Maruyama M, Kobayashi M: Resting T cells negatively regulate osteoclast generation from peripheral blood monocytes. Bone 2003;33:711–720.
181.
John V, Hock JM, Short LL, Glasebrook AL, Galvin RJS: A role for CD8+ T lymphocytes in osteoclast differentiation in vitro. Endocrinology 1996;137:2457–2463.
182.
Wyzga N, Varghese S, Wikel S, Canalis E, Sylvester FA: Effects of activated T cells on osteoclastogenesis depend on how they are activated. Bone 2004;35:614–620.
183.
Plows D, Kontogeorgos G, Kollias G: Mice lacking mature T and B lymphocytes develop arthritic lesions after immunization with type II collagen. J Immunol 1999;162:1018–1023.
184.
Firestein GS, Zvaifler NJ: How important are T cells in chronic rheumatoid synovitis? Arthritis Rheum 1990;33:768–773.
185.
Kinne RW, Palombo-Kinne E, Emmrich F: T-cells in the pathogenesis of rheumatoid arthritis: villains or accomplices? Biochem Biophys Acta 1997;1360:109–141.
186.
Marriott I, Rati D, McCall S, Tranguch S: Induction of Nod1 and Nod2 intracellular pattern recognition receptors in murine osteoblasts following bacterial challenge. Infect Immun 2005;73:2967–2973.
187.
Stanley KT, VanDort Ch, Motyl Ch, Endres J, Fox DA: Immunocompetent properties of human osteoblasts: interactions with T lymphocytes. J Bone Miner Res 2006;1:29–36.
188.
Kronenberg HM: PTH signaling and hematopoiesis. 1st Int Conf Osteoimmunology, 2006, A15.
189.
Marie PJ: Role of N-cadherin in bone formation. J Cell Physiol 2006;190:297–305.
190.
Aerssens J, Boonen S, Lowet G, Dequeker J: Interspecies differences in bone composition, density, and quality: potential implications for in vivo bone research. Endocrinology 1998;139:663–670.
191.
Arai F, Hirao A, Ohmura M, Sato H, Matsuoka S, Takubo K, Ito K, Koh GY, Suda T: Tie2/Angiopoietin-1 signaling regulates hematopoietic stem cell quiescence in the bone marrow niche. Cell 2004;118:149–161.
192.
Sipos W, Duvigneau JC, Hofbauer G, Schmoll F, Baravalle G, Exel B, Hartl R, Dobretsberger M, Pietschmann P: Characterization of the cytokine pattern of porcine bone marrow-derived cells treated with 1α,25(OH)2D3. J Vet Med [A] 2005;52:382–387.
193.
Rude RK, Gruber HE, Wie L, Frausto A, Mills BG: Magnesium deficiency: effect on bone and mineral metabolism in the mouse. Calcified Tissue Int 2003;72:32–41.
194.
Miura M, Tanaka K, Komatsu J, Suda M, Yasoda A, Sakuma Y, Ozasa A, Nakao K: A novel interaction between thyroid hormones and 1,25(OH)2D3 in osteoclast formation. Biochem Biophys Res Commun 2002;291:987–994.
195.
Pufe T, Scholz-Ahrens KE, Franke ATM, Petersen W, Mentlein R, Varoga D, Tillmann B, Schrezenmeir J, Glüer CC: The role of vascular endothelial growth factor in glucocorticoid-induced bone loss: evaluation in a minipig model. Bone 2003;33:869–876.
196.
Sipos W, Krenbek D, Rauner M, Pietschmann P: Cytokine expression profiles of porcine osteoblasts as determined by flow cytometry. Proc 1st Int Congr Osteoimmunology. Aegean Conf Ser 2006;22:73.
197.
Anonymous: Osteoporosis prevention, diagnosis and therapy. JAMA 2001;285:785–795.
198.
Melton LJ, Chrischilles EA, Lane AW, Riggs BL: Perspective. How many women have osteoporosis? J Bone Miner Res 1992;7:1005–1010.
199.
Obermayer-Pietsch B: Genetics of Osteoporosis. Wien Med Wochenschr 2006;156:162–167.
200.
Pietschmann P, Kerschan-Schindl K: Osteoporosis: gender-specific aspects. Wien Med Wochenschr 2004;154:411–415.
201.
Schett G: Rheumatoid arthritis: inflammation and bone loss. Wien Med Wochenschr 2006;156:34–41.
202.
Tanaka Y, Nakayamada S, Okoada Y: Osteoblasts and osteoclasts in bone remodeling and inflammation. Curr Drug Targets Inflamm Allergy 2005;4:325–328.
203.
Riggs BL, Khosla S, Melton LJ III: A unitary model for involutional osteoporosis: estrogen deficiency causes both type I and type II osteoporosis in postmenopausal women and contributes to bone loss in aging men. J Bone Miner Res 1998;13:763–773.
204.
Riggs BL, Melton LJ III: Evidence for two distinct syndromes of involutional osteoporosis. Am J Med 1983;75:899–901.
205.
Pacifici R, Rifas L, Teitelbaum S, Slatopolsky E, McCracken R, Bergfeld M, Lee W, Avioli LV, Peck WA: Spontaneous release of interleukin 1 from human blood monocytes reflects bone formation in idiopathic osteoporosis. Proc Natl Acad Sci 1987;84:4616–4620.
206.
Pacifici R, Rifas L, McCracken R, Vered I, McMurtry C, Avioli LV, Peck WA: Ovarian steroid treatment blocks a postmenopausal increase in blood monocyte interleukin 1 release. Proc Natl Acad Sci 1989;86:2398–2402.
207.
Pacifici R, Brown C, Puscheck E, Friedrich E, Slatopolsky E, Maggio D, McCracken R, Avioli LV: Effect of surgical menopause and estrogen replacement on cytokine release from human blood mononuclear cells. Proc Natl Acad Sci USA 1991;88:5134–5138.
208.
Zarrabeitia MT, Riancho JA, Amado JA, Napal J, Gonzalez-Macias J: Cytokine production by peripheral blood cells in postmenopausal osteoporosis. Bone Mineral 1991;14:161–167.
209.
Zheng SX, Vrindts Y, Lopez M, De Groote D, Zangerle PF, Collette J, Franchimont N, Geenen V, Albert A, Reginster JY: Increase in cytokine production (IL-1β, IL-6, TNF-α but not IFN-γ, GM-CSF or LIF) by stimulated whole blood cells in postmenopausal osteoporosis. Maturitas 1997;26:63–71.
210.
Franceschi C, Bonafe M, Valensin S, Oliviera F, De Luca M, Ottaviani E, De Benedictis G: Inflamm-Aging. An evolutionary perspective on immunosenescence. Ann NY Acad Sci 2000;908:244–254.
211.
Fujita T, Matsui T, Nakao Y, Watanabe S: T lymphocyte subsets in osteoporosis. Effect of 1-alpha hydroxyvitamin D. Min Electrolyte Metab 1984;10:375–378.
212.
Imai Y, Tsunenari T, Fukase M, Fujita T: Quantitative bone histomorphometry and circulating T lymphocyte subsets in postmenopausal osteoporosis. J Bone Miner Res 1990;4:393–399.
213.
Rosen CJ, Usiskin K, Owens M, Barlascini CO, Belsky M, Adler RA: T lymphocyte surface antigen markers in osteoporosis. J Bone Miner Res 1990;5:851–855.
214.
Hustmyer FG, Walker E, Yu X-P, Girasole G, Sakagami Y, Peacock M, Manolagas SC: Cytokine production and surface antigen expression by peripheral blood mononuclear cells in postmenipausal osteoporosis. J Bone Miner Res 1993;8:51–59.
215.
Pietschmann P, Grisar J, Thien R, Willheim M, Kerschan-Schindl K, Preisinger E, Peterlik M: Immune phenotype and intracellular cytokine production of peripheral blood mononuclear cells from postmenopausal patients with osteoporotic fractures. Experimental Gerontol 2001;36:1749–1759.
216.
Eghbali-Fatourechi G, Khosla S, Sanyal A, Boyle WJ, Lacey DL, Riggs BL: Role of RANK ligand in mediating increased bone resorption in early postmenopausal women. J Clin Invest 2003;111:1221–1230.
217.
Abdallah BM, Stilgren LS, Nissen N, Kassem M, Jorgensen HR, Abrahamsen B: Increased RANKL/OPG mRNA ratio in iliac bone biopsies from women with hip fractures. Calcif Tissue Int 2005;76:90–97.
218.
Yano K, Tsuda E, Washida N, Kobayashi F, Goto M, Harada A, Ikeda K, Higashio K, Yamada Y: Immunological characterization of circulation osteoprotegerin/osteoclastogenesis in inhibitory factor: increased serum concentrations in postmenopausal women with osteoprosis. J Bone Miner Res 1999;14:518–527.
219.
Riggs BL, Khosla S, Atkinson EJ, Dunstan CR, Melton LJ III: Evidence that type I osteoporosis results from enhanced responsiveness of bone to estrogen deficiency. Osteoporos Int 2003;14:728–733.
220.
Grigrie D, Neacsu E, Marinescu M, Popa O: Circulating osteoprotegerin and leptin levels in postmenopausal women with and without osteoporosis. Rom J Intern Med 2003;41:409–415.
221.
Mezquita-Raya P, de la Higuera M, Garcia DF, Alonso G, Ruiz-Requena ME, de Dios Luna J, Escobar-Jimenez F, Monoz-Tprres M: The contribution of serum osteoprotegerin to bone mass and vertebral fractures in postmenopausal women. Osteoporos Int 2005;16:1368–1374.
222.
Fahrleitner-Pammer A, Dobnig H, Piswanger-Soelkner C, Bonelli C, Dimai HP, Leb G, Obermayer-Pietsch B: Osteoprotegerin serum levels in women: correlation with age, bone mass, bone turnover and fracture status. Wien Klin Wochenschr 2003;115:291–297.
223.
Liu JM, Zhao HY, Ning G, Zhao YJ, Chen Y, Zhang LZH, Sun LH, Xu M-Y, Chen JL: Relationships between the changes of serum levels of OPG and RANKL with age, menopause, bone biochemical markers and bone mineral density in Chinese women aged 20–75. Calcif Tissue Int 2005;76:1–6.
224.
Schett G, Kiechl S, Redlich K, Oberhollenzer F, Weger S, Egger G, Mayr A, Jocher J, Xu Q, Pietschmann P, Teitelbaum S, Smolen J, Willeit J: Soluble RANKL and risk of nontraumatic fracture. JAMA 2004;291:1108–1113.
225.
Rogers A, Eastell R: Review: Circulating osteoprotegerin and receptor activator for nuclear factor κB ligand: clinical utility in metabolic bone disease assessment. J Clin Endocrinol Metab 2005;90:6323–6331.
226.
McClung MR, Lewiecki EM, Cohen SB, Bolognese MA, Woodson GC, Moffett AH, Peacock M, Miller PD, Lederman SN, Chesnut CH, Lain D, Kivitz AJ, Holloway DL, Zhang C, Peterson MC, Bekker PJ: AMG 162 Bone Loss Study Group: denosumab in postmenopausal women with low bone mineral density. N Engl J Med 2006;354:860–863.
227.
The Women’s Health Initiative (WHI) Steering Committee: Effects of conjugated equine estrogen in postmenopausal women with hysterectomy. JAMA 2004;291:1701.
228.
Writing Group of the WHI Investigators: Risks and benefits of estrogen plus progestin in healthy postmenopausal women. JAMA 2002;288:321.
229.
Riggs BL, Hartmann LC: Selective estrogen-receptor modulators: mechanisms of action and application to clinical practice. N Engl J Med 2003;348:618–629.
Copyright / Drug Dosage / Disclaimer
Copyright: All rights reserved. No part of this publication may be translated into other languages, reproduced or utilized in any form or by any means, electronic or mechanical, including photocopying, recording, microcopying, or by any information storage and retrieval system, without permission in writing from the publisher.
Drug Dosage: The authors and the publisher have exerted every effort to ensure that drug selection and dosage set forth in this text are in accord with current recommendations and practice at the time of publication. However, in view of ongoing research, changes in government regulations, and the constant flow of information relating to drug therapy and drug reactions, the reader is urged to check the package insert for each drug for any changes in indications and dosage and for added warnings and precautions. This is particularly important when the recommended agent is a new and/or infrequently employed drug.
Disclaimer: The statements, opinions and data contained in this publication are solely those of the individual authors and contributors and not of the publishers and the editor(s). The appearance of advertisements or/and product references in the publication is not a warranty, endorsement, or approval of the products or services advertised or of their effectiveness, quality or safety. The publisher and the editor(s) disclaim responsibility for any injury to persons or property resulting from any ideas, methods, instructions or products referred to in the content or advertisements.