Skip to main content
Advertisement
Browse Subject Areas
?

Click through the PLOS taxonomy to find articles in your field.

For more information about PLOS Subject Areas, click here.

  • Loading metrics

Negative feedback loop of bone resorption by NFATc1-dependent induction of Cadm1

  • Shinya Nakamura ,

    Contributed equally to this work with: Shinya Nakamura, Takuma Koyama

    Affiliation Department of Orthopaedic Surgery, Faculty of Medicine, The University of Tokyo, Bunkyo-ku, Tokyo, Japan

  • Takuma Koyama ,

    Contributed equally to this work with: Shinya Nakamura, Takuma Koyama

    Affiliation Department of Orthopaedic Surgery, Faculty of Medicine, The University of Tokyo, Bunkyo-ku, Tokyo, Japan

  • Naohiro Izawa,

    Affiliation Department of Orthopaedic Surgery, Faculty of Medicine, The University of Tokyo, Bunkyo-ku, Tokyo, Japan

  • Seitaro Nomura,

    Affiliation Genome Science Division, Research Center for Advanced Science and Technology (RCAST), The University of Tokyo, Tokyo, Japan

  • Takanori Fujita,

    Affiliation Genome Science Division, Research Center for Advanced Science and Technology (RCAST), The University of Tokyo, Tokyo, Japan

  • Yasunori Omata,

    Affiliation Department of Orthopaedic Surgery, Faculty of Medicine, The University of Tokyo, Bunkyo-ku, Tokyo, Japan

  • Takashi Minami,

    Affiliation Division of Phenotype Disease Analysis, Institute of Resource Development and Analysis, Kumamoto University, Kumamoto, Kumamoto, Japan

  • Morio Matsumoto,

    Affiliation Department of Orthopedic Surgery, Keio University, Tokyo, Japan

  • Masaya Nakamura,

    Affiliation Department of Orthopedic Surgery, Keio University, Tokyo, Japan

  • Eriko Fujita-Jimbo,

    Affiliation Department of Pediatrics, Jichi Medical University School of Medicine, Shimotsuke, Tochigi, Japan

  • Takashi Momoi,

    Affiliation Department of Pathophysiology, Tokyo Medical University, Shinjuku, Tokyo, Japan

  • Takeshi Miyamoto,

    Affiliations Department of Orthopaedic Surgery, Faculty of Medicine, The University of Tokyo, Bunkyo-ku, Tokyo, Japan, Department of Orthopedic Surgery, Keio University, Tokyo, Japan

  • Hiroyuki Aburatani,

    Affiliation Genome Science Division, Research Center for Advanced Science and Technology (RCAST), The University of Tokyo, Tokyo, Japan

  • Sakae Tanaka

    TANAKAS-ORT@h.u-tokyo.ac.jp

    Affiliation Department of Orthopaedic Surgery, Faculty of Medicine, The University of Tokyo, Bunkyo-ku, Tokyo, Japan

Abstract

Trimethylation of histone H3 lysine 4 and lysine 27 (H3K4me3 and H3K27me3) at gene promoter regions critically regulates gene expression. Key developmental genes tend to exhibit changes in histone modification patterns from the H3K4me3/H3K27me3 bivalent pattern to the H3K4me3 monovalent pattern. Using comprehensive chromatin immunoprecipitation followed by sequencing in bone marrow-derived macrophages (BMMs) and mature osteoclasts, we found that cell surface adhesion molecule 1 (Cadm1) is a direct target of nuclear factor of activated T cells 1 (NFATc1) and exhibits a bivalent histone pattern in BMMs and a monovalent pattern in osteoclasts. Cadm1 expression was upregulated in BMMs by receptor activator of nuclear factor kappa B ligand (RANKL), and blocked by a calcineurin/NFATc1 inhibitor, FK506. Cadm1-deficient mice exhibited significantly reduced bone mass compared with wild-type mice, which was due to the increased osteoclast differentiation, survival and bone-resorbing activity in Cadm1-deficient osteoclasts. These results suggest that Cadm1 is a direct target of NFATc1, which is induced by RANKL through epigenetic modification, and regulates osteoclastic bone resorption in a negative feedback manner.

Introduction

Skeletal homeostasis is maintained in the balance between bone resorption and bone formation. Osteoclasts (OCs) are multinucleated cells derived from monocyte/macrophage-lineage precursor cells and are primarily responsible for bone resorption [1]. Receptor activator of nuclear factor kappa B ligand (RANKL), a member of the tumor necrosis factor superfamily of cytokines, is indispensable for osteoclast differentiation [2], and its deficiency abrogates osteoclastogenesis and leads to osteopetrosis [3, 4]. RANKL binds to its receptor RANK and activates a wide range of intracellular signaling cascades that activate nuclear factor of activated T cells 1 (NFATc1), a transcription factor essential for osteoclast differentiation [57].

Epigenetic regulation of gene expression through post-translational modification of histone proteins by methylation and/or acetylation plays an important role in various processes such as cell cycle regulation, embryonic development and cellular differentiation [8]. Trimethylation of histone H3 at lysine 4 (H3K4me3) is mainly localized at gene promoter regions and is correlated with transcriptional activation, while trimethylation of H3 lysine 27 (H3K27me3) is involved in polycomb-mediated gene repression [9, 10]. Bernstein et al. reported that changes in the histone modification status in embryonic stem cells from H3K4me3/H3K27me3 bivalent patterns to H3K4me3 monovalent patterns by H3K27 demethylation are associated with active gene expression [11]. We previously demonstrated the bivalent to monovalent change of histone methylation at the promoter region of the NFATc1 gene during RANKL-induced osteoclastogenesis using chromatin immunoprecipitation followed by sequencing (ChIP-seq) [12]. NFATc1 is induced by RANKL stimulation downstream of calcium/calcineurin signaling [13], and plays key roles in osteoclastogenesis by regulating the expression of various osteoclast-related genes, such as Nfatc1 itself, Cathepsin K (Ctsk) and dendritic cell-specific transmembrane protein (Dc-stamp) [1416].

Cell adhesion molecule 1 (CADM1), also known as IGSF4 (immunoglobulin superfamily 4), NECL2 (Nectin-like molecule 2), SynCAM1 (synaptic cell adhesion molecule 1), sgIGSF (spermatogenic immunoglobulin superfamily) or TSLC1 (tumor suppressor in lung cancer 1), is a member of the immunoglobulin superfamily. A missense mutation in the CADM1 gene was reported in some patients with autism [1723]. CADM1 was originally identified as a tumor suppressor gene that suppressed tumor growth in nude mice and suppressed cancer metastasis by regulating cell adhesion [24]. Cadm1 knockout (KO) mice exhibited oligoasthenoteratozoospermia and impaired social/emotional behaviors [25, 26]. However, no skeletal phenotypes in Cadm1 KO mice have been reported.

In this study, we explored NFATc1 target genes whose histone methylation status at the promoter regions changed from bivalent to monovalent patterns.

Materials and methods

Reagents

Alpha-minimum essential medium (α-MEM) and fetal bovine serum (FBS) were purchased from Life Technologies (Carlsbad, CA, USA). Recombinant human M-CSF was purchased from R&D Systems (Minneapolis, MN, USA), and GST-RANKL was purchased from Oriental Yeast Co., Ltd (Shiga, Japan). Anti-trimethyl-histone H3 lysine 4 and anti-trimethyl-histone H3 lysine 27 were from Active Motif (rabbit polyclonal antibody, 39159, Carlsbad, CA, USA) and Millipore (rabbit polyclonal antibody, 07–449, Billerica, MA, USA), respectively. Anti-NFATc1 antibody detects all DNA binding domain-containing NFATc1 splicing isoforms, but not the closely related NFATc2 [27]. Anti-FAK, anti-Pyk2 and anti-Src antibodies were from Cell Signaling Technology (rabbit polyclonal antibody; Beverly, MA, USA).

Animal models

Wild-type mice on a C57BL/6 background were purchased from Sankyo Labo Service (Tokyo, Japan). Cadm1 KO mice on a C57BL/6 background were generated as described [25, 26]. Mice were housed in individually ventilated cages with paper bedding and provided standard rodent maintenance diet and water throughout the study. At the studies, mice were euthanized by cervical dislocation. Ethics committee Institutional Care and Use Committee (IACUC) in the University of Tokyo specifically approved this study (P14-096).

Cell culture

Murine bone marrow cells were isolated from the femur and tibia of male mice at 7–9 weeks of age. To prepare bone marrow macrophages (BMMs), cells were cultured in α-MEM/10% FBS with 50 ng/ml M-CSF for 2 days. BMMs were further cultured in the presence of 50 ng/ml M-CSF and 25 ng/ml RANKL for 3–5 days to generate osteoclasts.

To detect actin rings, cells were incubated for 30 min with rhodamine-conjugated phalloidin solution (Molecular Probes, Inc., Eugene, OR, USA) and observed under a fluorescence microscope (BZ-8100, Keyence).

ChIP-seq

Cells were fixed with 1% formaldehyde at room temperature and then neutralized with glycine. Samples were then sonicated and incubated with protein A/G beads that had been pre-incubated with 4–10 μg of antibody. Immunoprecipitates were washed and reverse-crosslinked, and samples were then DNA purified using a PCR purification kit (Qiagen GmbH, Hilden, Germany). DNA libraries were prepared for sequencing using the standard Illumina protocol. Purified DNA was applied for cluster generation and sequencing was performed using the cBot Cluster Generation system and Genome Analyzer IIx system (Illumina; San Diego, CA, USA), following the manufacturer’s instructions. Obtained sequences were mapped to the reference mouse genome.

RNA-seq

Total RNA was extracted using ISOGEN (Nippon Gene, Tokyo, Japan) from BMMs or OCs following the manufacturer’s protocol. Libraries were made using TruSeq RNA Sample Preparation Kits (Illumina), according to the manufacturer’s instructions. RNA sequencing was performed using the Illumina Genome Analyzer IIx.

Analysis of bone mineral density (BMD) and skeletal morphology

Eight-week-old male wild-type or Cadm1 KO mice were subcutaneously injected with 10 mg/kg body weight of calcein on 4 days and 1 day before sacrifice. Hindlimbs were removed, fixed with 70% ethanol, and subjected to a dual energy x-ray absorptiometric scan analysis to measure BMD (mg/cm2) using the DCS-600R system (Aloka Co. Ltd, Tokyo, Japan). Hindlimbs were then undecalcified, embedded in glycol methacrylate, cut into 3 μm sections longitudinally in the proximal region of the tibia, and stained with toluidine blue O or tartrate-resistant acid phosphatase (TRAP). For TRAP staining, tissue slices were stained at pH 5.0 in the presence of L (+)-tartaric acid using naphthol AS-MX phosphate (Sigma-Aldrich) in N,N-dimethyl formamide as substrate. Histomorphometric measurement was performed in stained sections from the secondary spongiosa area, 1.05 mm from the growth plate and 0.4 mm from the end of metaphysis, using OsteoMeasure software (OsteoMetrics, Inc., Decatur, GA).

Bone resorption assay

Osteoclasts were cultured on dentine slices. Resorption areas were visualized by staining with 1% toluidine blue and then measured under a microscope (BZ-8100).

Survival assay

Once osteoclasts were generated, RANKL and M-CSF were removed from the culture medium (at a time defined as time 0), and cells were cultured for the time periods indicated in the text. Survival was calculated by dividing the number of morphologically intact TRAP+ multinucleated cells by those present at time 0.

Real-time PCR analysis

Total RNA was extracted with an RNeasy mini kit (Qiagen GmbH, Hilden, Germany) and reverse transcription was performed using a Prime Script RT reagent kit (Takara Bio) according to the manufacturer’s instructions. Realtime PCR was performed with a Thermal Cycler Dice Real-Time System using SYBR Premix Ex Taq (Takara Bio). All reactions were performed in triplicate. Relative mRNA expression levels were normalized to that of mouse β-actin.

Primer sequences were as follows:

β-actin forward, 5′-TGAGAGGGAAATCGTGCGTGAC-3′;

β-actin reverse, 5′-AAGAAGGAAGGCTGGAAAAGAG-3′;

Cadm1 forward, 5′-GAAGGACAGCAGGTTTCAGC-3′;

Cadm1 reverse, 5′-TCAACTGCCGTGTCTTTCTG-3′;

Nfatc1 forward, 5′-CAAGTCTCACCACAGGGCTCACTA-3′;

Nfatc1 reverse, 5′-GCGTGAGAGGTTCATTCTCCAAGT-3′;

Ctsk forward, 5′-GGACCCATCTCTGTGTCCAT-3′;

Ctsk reverse, 5′-CCGAGCCAAGAGAGCATATC-3′;

Dc-stamp forward, 5′-TCCTCCATGAACAAACAGTTCCAA-3′;

Dc-stamp reverse, 5′-AGACGTGGTTTAGGAATGCAGCTC-3′;

Acp5 forward, 5′-TTGCGACCATTGTTAGCCACATA-3′;

Acp5 reverse, 5′-TCAGATCCATAGTGAAACCGCAAG-3′.

Western blotting

Cells were washed with ice-cold PBS and lysed at 4°C with RIPA buffer (1% Tween20, 0.1% SDS, 150 mM NaCl, 10 mM Tris-HCl (pH 7.4), 0.25 mM phenylmethylsulfonyl fluoride, 10 μg/ml aprotinin, 10 μg/ml leupeptin, 1 mM Na3VO4, and 5 mM NaF (Sigma)). Proteins were subjected to SDS-PAGE on 7.5–15% Tris-Glycine gradient gels or 15% Tris-Glycine gels and transferred onto PVDF membranes (Millipore Corp.). After blocking, membranes were incubated with primary antibodies to FAK, Pyk2, and Src (Cell Signaling Technology) or Actin (Sigma-Aldrich), followed by HRP-conjugated goat anti-rabbit IgG (Promega). Immunoreactive bands were visualized by ECL (GE Healthcare) according to the manufacturer’s instructions. Blots were stripped by a 30-min incubation in buffer (2% SDS, 100 mM 2-mercaptoethanol, and 62.5 mM Tris-HCl, pH 6.7) at 50°C and then re-probed with other antibodies.

Statistical analysis

Data are presented as means ± SD. Statistical analyses were performed using a two-tailed unpaired Student’s t-test.

Results

Cadm1 exhibits changes from the H3K4me3/H3K27me3 bivalent to the H3K4me3 monovalent pattern and is an NFATc1 target in osteoclasts

We first identified genes that exhibited H3K4me3/H3K27me3 bivalent patterns at the promoter region in BMMs and H3K4me3 monovalent patterns in mature osteoclasts by comprehensive ChIP-seq analysis using anti-H3K4me3 and anti-H3K27me3 antibodies, as previously described [12, 28, 29]. We selected 49 genes that showed the change of the histone methylation pattern within 5 kb of transcriptional start sites (TSSs), which are putative promoter regions. We also performed ChIP-seq using anti-NFATc1 antibody, and found that 33 genes among the 49 genes exhibited NFATc1 binding at the promoter regions in osteoclasts. These genes included representative osteoclast-specific genes, such as Nfatc1 and carbonic anhydrase 2 (Car2) (Table 1) [14, 30].

thumbnail
Table 1. Genes whose histone modification pattern changes from bivalent to monovalent during osteoclastogenesis and with NFATc1 binding near the transcription start site.

https://doi.org/10.1371/journal.pone.0175632.t001

Among the 33 identified genes, we focused on Cadm1, which encodes a cell adhesion molecule belonging to the immunoglobulin superfamily [24]. Immunoglobulin superfamily proteins played a critical role in RANKL-induced osteoclastogenesis by regulating intracellular calcium signaling [31]. In addition, members such as intercellular adhesion molecule-1 (ICAM1) played critical roles in osteoclast function [31]. Therefore, Cadm1 may regulate osteoclast differentiation and function via its immunoglobulin-like motifs. Histone modifications at the Cadm1 promoter changed from H3K4me3/H3K27me3 bivalent patterns in BMMs to H3K4me3 monovalent patterns in response to RANKL (Fig 1A). NFATc1 binding at the Cadm1 promoter in osteoclasts was detected by ChIP-seq (Fig 1A), confirming our selection criteria. Real-time PCR analysis revealed that Cadm1 expression was greatly increased during RANKL-stimulated osteoclastogenesis and positively correlated with elevated expression of Ctsk, a marker of osteoclast differentiation (Fig 1B). RANKL-dependent Cadm1 upregulation in osteoclasts was blocked by treatment of cells with a calcium/calcineurin signaling inhibitor, FK506, confirming that Cadm1 is an NFATc1 target (Fig 1C).

thumbnail
Fig 1. Epigenetic regulation and expression of Cadm1 during osteoclastogenesis.

(A) ChIP-seq analysis of H3K4me3, H3K27me3 and NFATc1 near the Cadm1 transcription start site (TSS). (B and C) Expression of Cadm1 and Ctsk mRNAs relative to that of β-actin in BMMs cultured in the presence of M-CSF (50 ng/ml) and RANKL (25 ng/ml) with or without FK506 (1 μM) for indicated periods. Data represent mean Cadm1 or Ctsk expression relative to β-actin ± SD (*P < 0.05; n = 3).

https://doi.org/10.1371/journal.pone.0175632.g001

Cadm1 KO mice exhibit decreased bone mass

We next analyzed the skeletal tissue of Cadm1 KO mice. Cadm1 KO mice showed a significant reduction in bone mass compared with wild-type mice (Fig 2A). Bone morphometric analysis revealed that Cadm1 KO mice exhibited reduced BV/TV and Tb.N and elevated Tb.Sp (Fig 2B and 2C). In contrast, osteoclast parameters, such as ES/BS, Oc.S/BS and N.Oc/BS, or osteoblast parameters, such as Ob.S/BS, MS/BS and BFR, were not significantly different between wild-type and KO mice (Fig 2B and 2C).

thumbnail
Fig 2. Cadm1 KO mice exhibit decreased bone mineral density.

(A) Bone mineral density (BMD) of femurs from male wild-type or Cadm1 KO mice (*P < 0.05; **P < 0.005; n = 5). (B) Representative toluidine blue O staining (top) and TRAP staining (bottom) images of proximal tibias from male wild-type (left) or Cadm1 KO (right) mice. Bars, 100 μm. (C) Bone morphometric analysis of 8-week-old wild-type or Cadm1 KO mice. BV/TV, bone volume per total volume; Tb.Th, trabecular thickness; Tb.N, trabecular number; Tb.Sp, trabecular separation; ES/BS, eroded surface per bone surface; Oc.S/BS, osteoclast surface per bone surface; N.Oc/BS, osteoclast number per bone surface; Ob.S/BS, osteoblast surface per bone surface; MS/BS, mineralizing surface per bone surface; BFR/BS, bone formation rate per bone surface. Data represent mean value of the indicated parameter ± SD (#P < 0.1; *P < 0.05; n = 5).

https://doi.org/10.1371/journal.pone.0175632.g002

Cadm1 KO osteoclasts show enhanced differentiation and bone-resorbing function

To analyze the possible function of Cadm1 in osteoclasts, we isolated osteoclast progenitor cells from Cadm1 KO and wild-type mice and cultured them in the presence of M-CSF and RANKL (Fig 3). Osteoclast differentiation as analyzed by multinucleated TRAP-positive cell formation did not differ between Cadm1 KO and wild-type cells (Fig 3A). In addition, nuclear localization of Nfatc1 was not different between cadm1-deficient and wild-type osteoclasts (S1 Fig). However, real-time PCR analysis indicated that expressions of osteoclast-specific genes such as Nfatc1, Ctsk, acid phosphatase 5, tartrate resistant (Acp5) and Dc-stamp were slightly but significantly higher in Cadm1 KO osteoclasts as compared with wild-type cells (Fig 3B). In addition, Cadm1 KO osteoclasts were more resistant to cytokine depletion-induced cell death as compared with wild-type osteoclasts (Fig 3C). Furthermore, bone-resorbing activity, as evaluated by pit formation on dentine slices, was significantly higher in Cadm1 KO osteoclasts than wild-type osteoclasts (Fig 3D). Overall, these results suggest that Cadm1 negatively regulates osteoclast differentiation and function.

thumbnail
Fig 3. Cadm1 loss stimulates osteoclast differentiation and enhances survival and bone-resorption.

(A and B) BMMs from wild-type or Cadm1 KO mouse were cultured in the presence of M-CSF (50 ng/ml) and RANKL (25 ng/ml) for 4 days and then stained for TRAP (A) or evaluated by real-time PCR for expression of the indicated osteoclast markers (B). TRAP+ multinucleated cells (MNCs) containing more than three nuclei were counted as osteoclasts. Data represent mean number of osteoclasts ± SD (n.s., not significant; n = 3). Bar, 100 μm. Data represent mean Nfatc1, Cathepsin K (Ctsk), acid phosphatase 5, tartrate resistant (Acp5) or dendritic cell-specific transmembrane protein (Dc-stamp) expression relative to that of β-actin ± SD (B, **P < 0.01; n = 6). (C) BMMs from wild-type or Cadm1 KO mouse were cultured in the presence of M-CSF (50 ng/ml) and RANKL (25 ng/ml) for 4 days. M-CSF and RANKL were then removed from the medium, and cells were stained with TRAP at indicated times after cytokine withdrawal. Remaining TRAP+ cells were scored as surviving cells. Data represent mean number of surviving cells per well (%) ± SD (*P < 0.05; n = 4). (D) BMMs from wild-type or Cadm1 KO mouse were cultured in the presence of M-CSF (50 ng/ml) and RANKL (25 ng/ml) on dentine slices for 4 days; resorption areas were visualized by toluidine blue staining (left) and the resorption area was scored. Data represent mean resorption area (%) ± SD (**P < 0.01; n = 6).

https://doi.org/10.1371/journal.pone.0175632.g003

Cadm1 KO osteoclasts show enhanced expression of adhesion-related factors

Given that Cadm1 KO osteoclasts exhibited elevated bone-resorbing activity (Fig 3D), we first analyzed the formation of actin rings, which are cytoskeletal structures essential for the bone-resorbing activity of osteoclasts (Fig 4A). However, our results showed that actin rings were similarly formed in Cadm1 KO and wild-type osteoclasts (Fig 4A). We next analyzed the expression of FAK, Pyk2 and c-Src, signaling molecules implicated in cell adhesion and cytoskeletal organization of osteoclasts [32]. c-Src protein levels as assessed by western blot were equivalent between Cadm1 KO and wild-type osteoclasts, while FAK and Pyk2 levels were elevated in Cadm1 KO osteoclasts (Fig 4B).

thumbnail
Fig 4. Cadm1 ablation enhances bone-resorbing activity of osteoclasts.

(A) BMMs from wild-type or Cadm1 KO mouse were cultured in the presence of M-CSF (50 ng/ml) and RANKL (25 ng/ml) for 4 days, stained with rhodamine phalloidin to visualize actin protein, and observed under a fluorescence microscope. Bar, 100 μm. (B) Whole cell lysates from wild-type or Cadm1 KO osteoclasts were subjected to western blotting using antibodies against FAK, Pyk2 and Src. Actin protein expression served as an internal control. Protein levels relative to actin were quantified by densitometry and are shown below.

https://doi.org/10.1371/journal.pone.0175632.g004

Discussion

Epigenetic modifications are critical in the organization of chromatin structures at various levels and therefore regulate gene expression [33]. Among the various epigenetic modifications, histone methylation is associated with both transcriptionally active and repressive chromatin status. The methylated sites in histones H3 or H4 are mainly located in the histone tail (H3K4, H3K9, H3K36 and H4K20) and the center of the nucleosome (H3K79) [34]. In this study, we explored NFATc1 target genes whose histone methylation status at the promoter regions changed from bivalent to monovalent patterns, and we identified Cadm1 gene as a candidate. CADM1 is an immunoglobulin-like cell adhesion molecule that regulates intercellular attachment through homophilic interactions [20, 24]. Recent studies showed that the heterophilic interaction between CADM1 and other molecules also plays a critical role in various types of cells. For example, the attachment and functional interaction between dorsal root ganglia neurons and mast cells are mediated by heterophilic binding between mast cell CADM1 and neuronal nectin-3 [35]. The intracellular domain of CADM1 contains FERM (protein 4.1/ezrin/radixin/moesin)-binding and PDZ-binding motifs, which potentially act as a molecular scaffold to form signaling complexes [36]. Stagi et al. reported that FAK binds to CADM1 independently of PDZ domains, but requires the FERM motif, and is implicated in the morphogenetic activities of growth cones [37]. These studies suggest that CADM1 regulates cell-cell interactions and cytoskeletal organization.

Our findings showed that Cadm1 expression was induced by RANKL in BMMs in an NFATc1-dependent manner, and histone modification at its promoter regions changed from bivalent to monovalent patterns in response to RANKL. Osteoclasts differentiated from Cadm1 KO mouse BMMs showed higher expression of osteoclast differentiation markers such as Nfatc1, Ctsk, Acp5 and Dc-stamp than wild-type osteoclasts. In addition, osteoclasts differentiated from Cadm1 KO mouse BMMs showed enhanced bone-resorbing activity and survival. The increased bone-resorbing activity of Cadm1 KO osteoclasts is at least partly due to the increased expression of FAK and Pyk2, which are intracellular signaling molecules that play critical roles in the cytoskeletal organization in osteoclasts [32, 38, 39]. However, the mechanisms by which Cadm1 regulates FAK and Pyk2 expression in osteoclasts remain elusive. Consistent with these in vitro findings, Cadm1 KO mice exhibited reduced bone mass. However, bone resorption parameters such as ES/BS, Oc.S/BS and N.Oc/BS were not different between wild type and KO mice. The reason for this discrepancy is unclear, but it may be because the marked reduction of cancellous bone may hamper the accurate measurement of histomorphometric markers.

It is well established that immunoreceptor tyrosine-based activation motif (ITAM) signaling plays a pivotal role in osteoclast differentiation. Both DNAX activating protein of 12 kDa (DAP12) and Fc receptor common γ chain (FcRγ) contain cytoplasmic ITAM motif, and mice lacking DAP12 and FcRγ exhibit severe osteopetrosis due to the lack of osteoclasts [40, 41]. BMMs obtained from DAP12 knock out mouse macrophages undergo osteoclastogenesis in response to RANKL but fail to form actin rings, indicating the role of DAP12 in the cytoskeletal organization as well [42]. It was recently reported that sialic acid binding immunoglobulin-like lectin (Siglec)-15 is essential for the functional osteoclast differentiation. Siglec-15 is associated with DAP12 and regulates osteoclast differentiation through ITAM-mediated signaling pathways [43, 44]. Although the interaction between Cadm1 and DAP12 remains unknown, it is possible that Cadm1 negatively regulates osteoclast differentiation by antagonizing Siglec-15 and DAP12 interaction.

Several cytokines and chemokines induced by RANKL promote osteoclast differentiation, thus forming a positive feedback loop in osteoclastogenesis [45]. Conversely, negative regulators of osteoclastogenesis such as interferon regulatory factor 8 (Irf8) and Maf are downregulated by RANKL through B lymphocyte-induced maturation protein-1 (Blimp-1), which is induced by NFATc1 [45]. Here we demonstrated the possibility that the negative regulator Cadm1 is directly regulated by NFATc1 through epigenetic modifications. Our results may provide a novel mechanism for the fine-tuning of osteoclastogenesis through epigenetic and NFATc1-dependent mechanisms.

Conclusions

Cadm1 expression is induced by changes in epigenetic modification and by NFATc1 in osteoclasts following RANKL stimulation and serves to inhibit osteclastogenesis.

Supporting information

S1 Fig. BMMs from wild-type or cadm1 KO mice were cultured in the presence of M-CSF (50 ng/ml) and RANKL (25 ng/ml) for 4 days.

The cells were stained for Nfatc1 using mouse anti-NFATc1 antibody (mouse monoclonal antibody; Santa Cruz Biotechnology, Inc., Dallas, TX, USA) followed by Alexa546-conjugated goat anti-mouse IgG (Invitrogen, Carlsbad, CA, USA). DAPI (Wako Pure Chemicals Industries, Osaka, Japan) was used for a nuclear staining. Nfatc1 was visualized in red, DAPI was double-labeled in blue, and merged images were shown. Bar = 100 μm.

https://doi.org/10.1371/journal.pone.0175632.s001

(TIF)

Acknowledgments

We thank Ms. J. Sugita (Department of Orthopaedic Surgery, The University of Tokyo) for providing expert technical assistance. This work was supported in part by Grants-in-Aid from the Ministry of Education, Culture, Sports, Science and Technology of Japan.

Author Contributions

  1. Conceptualization: S Nakamura TK NI YO T Miyamoto ST.
  2. Data curation: S Nakamura NI S Nomura TF.
  3. Formal analysis: S Nakamura NI S Nomura TF.
  4. Funding acquisition: ST.
  5. Investigation: S Nakamura TK NI YO.
  6. Methodology: S Nakamura TK NI YO T Minami EF T Momoi T Miyamoto ST.
  7. Project administration: S Nakamura T Miyamoto HA ST.
  8. Resources: S Nakamura TK NI S Nomura TF YO T Minami EF T Momoi.
  9. Software: TF.
  10. Supervision: MM MN.
  11. Validation: S Nakamura TK T Miyamoto ST.
  12. Visualization: S Nakamura TK T Miyamoto ST.
  13. Writing – original draft: S Nakamura TK T Miyamoto ST.
  14. Writing – review & editing: S Nakamura TK T Miyamoto ST.

References

  1. 1. Boyle WJ, Simonet WS, Lacey DL. Osteoclast differentiation and activation. Nature. 2003;423(6937):337–42. pmid:12748652
  2. 2. Tanaka S. Signaling axis in osteoclast biology and therapeutic targeting in the RANKL/RANK/OPG system. Am J Nephrol. 2007;27(5):466–78. pmid:17652963
  3. 3. Theill LE, Boyle WJ, Penninger JM. RANK-L and RANK: T cells, bone loss, and mammalian evolution. Annu Rev Immunol. 2002;20:795–823. pmid:11861618
  4. 4. Lazner F, Gowen M, Pavasovic D, Kola I. Osteopetrosis and osteoporosis: two sides of the same coin. Hum Mol Genet. 1999;8(10):1839–46. pmid:10469835
  5. 5. Asagiri M, Sato K, Usami T, Ochi S, Nishina H, Yoshida H, et al. Autoamplification of NFATc1 expression determines its essential role in bone homeostasis. The Journal of experimental medicine. 2005;202(9):1261–9. pmid:16275763
  6. 6. Matsuo K, Galson DL, Zhao C, Peng L, Laplace C, Wang KZ, et al. Nuclear factor of activated T-cells (NFAT) rescues osteoclastogenesis in precursors lacking c-Fos. J Biol Chem. 2004;279(25):26475–80. pmid:15073183
  7. 7. Takayanagi H, Kim S, Koga T, Nishina H, Isshiki M, Yoshida H, et al. Induction and activation of the transcription factor NFATc1 (NFAT2) integrate RANKL signaling in terminal differentiation of osteoclasts. Dev Cell. 2002;3(6):889–901. pmid:12479813
  8. 8. Goldberg AD, Allis CD, Bernstein E. Epigenetics: a landscape takes shape. Cell. 2007;128(4):635–8. pmid:17320500
  9. 9. Strahl BD, Ohba R, Cook RG, Allis CD. Methylation of histone H3 at lysine 4 is highly conserved and correlates with transcriptionally active nuclei in Tetrahymena. Proc Natl Acad Sci U S A. 1999;96(26):14967–72. pmid:10611321
  10. 10. Schwartz YB, Pirrotta V. Polycomb silencing mechanisms and the management of genomic programmes. Nature reviews Genetics. 2007;8(1):9–22. pmid:17173055
  11. 11. Bernstein BE, Mikkelsen TS, Xie X, Kamal M, Huebert DJ, Cuff J, et al. A bivalent chromatin structure marks key developmental genes in embryonic stem cells. Cell. 2006;125(2):315–26. pmid:16630819
  12. 12. Yasui T, Kadono Y, Nakamura M, Oshima Y, Matsumoto T, Masuda H, et al. Regulation of RANKL-induced osteoclastogenesis by TGF-beta through molecular interaction between Smad3 and Traf6. J Bone Miner Res. 2011;26(7):1447–56. pmid:21305609
  13. 13. Hirotani H, Tuohy NA, Woo JT, Stern PH, Clipstone NA. The calcineurin/nuclear factor of activated T cells signaling pathway regulates osteoclastogenesis in RAW264.7 cells. J Biol Chem. 2004;279(14):13984–92. pmid:14722106
  14. 14. Takayanagi H. Osteoimmunology: shared mechanisms and crosstalk between the immune and bone systems. Nat Rev Immunol. 2007;7(4):292–304. pmid:17380158
  15. 15. Kim Y, Sato K, Asagiri M, Morita I, Soma K, Takayanagi H. Contribution of nuclear factor of activated T cells c1 to the transcriptional control of immunoreceptor osteoclast-associated receptor but not triggering receptor expressed by myeloid cells-2 during osteoclastogenesis. J Biol Chem. 2005;280(38):32905–13. pmid:16046394
  16. 16. Matsumoto M, Kogawa M, Wada S, Takayanagi H, Tsujimoto M, Katayama S, et al. Essential role of p38 mitogen-activated protein kinase in cathepsin K gene expression during osteoclastogenesis through association of NFATc1 and PU.1. J Biol Chem. 2004;279(44):45969–79. pmid:15304486
  17. 17. Zhiling Y, Fujita E, Tanabe Y, Yamagata T, Momoi T, Momoi MY. Mutations in the gene encoding CADM1 are associated with autism spectrum disorder. Biochemical and biophysical research communications. 2008;377(3):926–9. pmid:18957284
  18. 18. Biederer T, Stagi M. Signaling by synaptogenic molecules. Curr Opin Neurobiol. 2008;18(3):261–9. pmid:18725297
  19. 19. Gomyo H, Arai Y, Tanigami A, Murakami Y, Hattori M, Hosoda F, et al. A 2-Mb sequence-ready contig map and a novel immunoglobulin superfamily gene IGSF4 in the LOH region of chromosome 11q23.2. Genomics. 1999;62(2):139–46. pmid:10610705
  20. 20. Kuramochi M, Fukuhara H, Nobukuni T, Kanbe T, Maruyama T, Ghosh HP, et al. TSLC1 is a tumor-suppressor gene in human non-small-cell lung cancer. Nat Genet. 2001;27(4):427–30. pmid:11279526
  21. 21. Takai Y, Miyoshi J, Ikeda W, Ogita H. Nectins and nectin-like molecules: roles in contact inhibition of cell movement and proliferation. Nat Rev Mol Cell Biol. 2008;9(8):603–15. pmid:18648374
  22. 22. Urase K, Soyama A, Fujita E, Momoi T. Expression of RA175 mRNA, a new member of the immunoglobulin superfamily, in developing mouse brain. Neuroreport. 2001;12(15):3217–21. pmid:11711859
  23. 23. Wakayama T, Ohashi K, Mizuno K, Iseki S. Cloning and characterization of a novel mouse immunoglobulin superfamily gene expressed in early spermatogenic cells. Mol Reprod Dev. 2001;60(2):158–64. pmid:11553913
  24. 24. Murakami S, Sakurai-Yageta M, Maruyama T, Murakami Y. Trans-homophilic interaction of CADM1 activates PI3K by forming a complex with MAGuK-family proteins MPP3 and Dlg. PLoS One. 2014;9(9):e110062. pmid:25268382
  25. 25. Fujita E, Kouroku Y, Ozeki S, Tanabe Y, Toyama Y, Maekawa M, et al. Oligo-astheno-teratozoospermia in mice lacking RA175/TSLC1/SynCAM/IGSF4A, a cell adhesion molecule in the immunoglobulin superfamily. Mol Cell Biol. 2006;26(2):718–26. pmid:16382161
  26. 26. Takayanagi Y, Fujita E, Yu Z, Yamagata T, Momoi MY, Momoi T, et al. Impairment of social and emotional behaviors in Cadm1-knockout mice. Biochemical and biophysical research communications. 2010;396(3):703–8. pmid:20450890
  27. 27. Suehiro J, Kanki Y, Makihara C, Schadler K, Miura M, Manabe Y, et al. Genome-wide approaches reveal functional vascular endothelial growth factor (VEGF)-inducible nuclear factor of activated T cells (NFAT) c1 binding to angiogenesis-related genes in the endothelium. The Journal of biological chemistry. 2014;289(42):29044–59. pmid:25157100
  28. 28. Nakamura H, Nakashima T, Hayashi M, Izawa N, Yasui T, Aburatani H, et al. Global epigenomic analysis indicates protocadherin-7 activates osteoclastogenesis by promoting cell-cell fusion. Biochemical and biophysical research communications. 2014;455(3–4):305–11. pmid:25446128
  29. 29. Omata Y, Nakamura S, Koyama T, Yasui T, Hirose J, Izawa N, et al. Identification of Nedd9 as a TGF-beta-Smad2/3 Target Gene Involved in RANKL-Induced Osteoclastogenesis by Comprehensive Analysis. PLoS One. 2016;11(6):e0157992. pmid:27336669
  30. 30. Sly WS, Hu PY. Human carbonic anhydrases and carbonic anhydrase deficiencies. Annu Rev Biochem. 1995;64:375–401. pmid:7574487
  31. 31. Negishi-Koga T, Gober HJ, Sumiya E, Komatsu N, Okamoto K, Sawa S, et al. Immune complexes regulate bone metabolism through FcRgamma signalling. Nat Commun. 2015;6:6637. pmid:25824719
  32. 32. Ray BJ, Thomas K, Huang CS, Gutknecht MF, Botchwey EA, Bouton AH. Regulation of osteoclast structure and function by FAK family kinases. J Leukoc Biol. 2012;92(5):1021–8. pmid:22941736
  33. 33. Margueron R, Trojer P, Reinberg D. The key to development: interpreting the histone code? Curr Opin Genet Dev. 2005;15(2):163–76. pmid:15797199
  34. 34. Mersfelder EL, Parthun MR. The tale beyond the tail: histone core domain modifications and the regulation of chromatin structure. Nucleic Acids Res. 2006;34(9):2653–62. pmid:16714444
  35. 35. Furuno T, Hagiyama M, Sekimura M, Okamoto K, Suzuki R, Ito A, et al. Cell adhesion molecule 1 (CADM1) on mast cells promotes interaction with dorsal root ganglion neurites by heterophilic binding to nectin-3. J Neuroimmunol. 2012;250(1–2):50–8. pmid:22703826
  36. 36. Yageta M, Kuramochi M, Masuda M, Fukami T, Fukuhara H, Maruyama T, et al. Direct association of TSLC1 and DAL-1, two distinct tumor suppressor proteins in lung cancer. Cancer Res. 2002;62(18):5129–33. pmid:12234973
  37. 37. Stagi M, Fogel AI, Biederer T. SynCAM 1 participates in axo-dendritic contact assembly and shapes neuronal growth cones. Proc Natl Acad Sci U S A. 2010;107(16):7568–73. pmid:20368431
  38. 38. Tanaka S, Takahashi N, Udagawa N, Murakami H, Nakamura I, Kurokawa T, et al. Possible involvement of focal adhesion kinase, p125FAK, in osteoclastic bone resorption. J Cell Biochem. 1995;58(4):424–35. pmid:7593264
  39. 39. Xiong WC, Feng X. PYK2 and FAK in osteoclasts. Front Biosci. 2003;8:d1219–26. pmid:12957821
  40. 40. Koga T, Inui M, Inoue K, Kim S, Suematsu A, Kobayashi E, et al. Costimulatory signals mediated by the ITAM motif cooperate with RANKL for bone homeostasis. Nature. 2004;428(6984):758–63. pmid:15085135
  41. 41. Mocsai A, Humphrey MB, Van Ziffle JA, Hu Y, Burghardt A, Spusta SC, et al. The immunomodulatory adapter proteins DAP12 and Fc receptor gamma-chain (FcRgamma) regulate development of functional osteoclasts through the Syk tyrosine kinase. Proc Natl Acad Sci U S A. 2004;101(16):6158–63. pmid:15073337
  42. 42. Zou W, Zhu T, Craft CS, Broekelmann TJ, Mecham RP, Teitelbaum SL. Cytoskeletal dysfunction dominates in DAP12-deficient osteoclasts. J Cell Sci. 2010;123(Pt 17):2955–63. pmid:20720152
  43. 43. Ishida-Kitagawa N, Tanaka K, Bao X, Kimura T, Miura T, Kitaoka Y, et al. Siglec-15 protein regulates formation of functional osteoclasts in concert with DNAX-activating protein of 12 kDa (DAP12). J Biol Chem. 2012;287(21):17493–502. pmid:22451653
  44. 44. Kameda Y, Takahata M, Komatsu M, Mikuni S, Hatakeyama S, Shimizu T, et al. Siglec-15 regulates osteoclast differentiation by modulating RANKL-induced phosphatidylinositol 3-kinase/Akt and Erk pathways in association with signaling Adaptor DAP12. J Bone Miner Res. 2013;28(12):2463–75. pmid:23677868
  45. 45. Miyamoto K, Ninomiya K, Sonoda KH, Miyauchi Y, Hoshi H, Iwasaki R, et al. MCP-1 expressed by osteoclasts stimulates osteoclastogenesis in an autocrine/paracrine manner. Biochemical and biophysical research communications. 2009;383(3):373–7. pmid:19364494