DOI 10.1515/cclm-2013-0770      Clin Chem Lab Med 2014; 52(5): 589–608

Review Peter Vrtačnik, Janja Marc and Barbara Ostanek*

Epigenetic mechanisms in bone Abstract: Epigenetics refers to the study of mechanisms able to influence gene expression in a stable and potentially heritable manner without altering the DNA sequence. These mechanisms include posttranslational histone modifications, miRNA-mediated post-transcriptional regulation and DNA methylation. The accumulation of molecular errors over time resulting, at least partly, in the alteration of normal epigenetic patterns is being widely associated with aging. Epigenetic processes are also considered important mechanisms through which environmental and stochastic stressors promote numerous pathologies in humans. It is, therefore, reasonable to expect that several complex multi-factorial late-onset disorders, like osteoporosis and osteoarthritis, could have a strong epigenetic component. The focal point of all skeletal pathologies is the deregulation of bone remodeling, mediated by bone-forming osteoblasts and boneresorbing osteoclasts. In order to keep both processes in balance, the activity, differentiation and apoptosis of both cell types have to be tightly regulated. In particular, the differentiation of osteoblasts and osteoclasts is accompanied by profound changes in gene expression. It has been shown that histone deacetylation and DNA methylation negatively regulate the expression of several genes associated with different stages of osteoblast differentiation; however, several miRNAs promote osteoblastogenesis. Furthermore, inactivating mutations in the miRNA coding regions could be associated with the pathogenesis of osteoporosis. The aim of this review is to highlight the role of epigenetic mechanisms in bone remodeling and bone homeostasis, so as to implicate their diagnostic and therapeutic potential in skeletal diseases. Keywords: bone remodeling; DNA methylation; miRNA; osteoarthritis; osteoporosis; posttranslational histone modifications. *Corresponding author: Barbara Ostanek, Faculty of Pharmacy, Department of Clinical Biochemistry, University of Ljubljana, Aškerčeva 7, 1000 Ljubljana, Slovenia, Phone: +386 1 4769689, Fax: +386 1 4258031, E-mail: [email protected] Peter Vrtačnik and Janja Marc: Faculty of Pharmacy, Department of Clinical Biochemistry, University of Ljubljana, Ljubljana, Slovenia

Abbreviations: ACVR1B/2A, activin A receptor, type IB/IIA; ALPL, alkaline phosphatase; APC, adenomatous polyposis coli; ATF4, activating transcription factor 4; BAMBI, BMP and activin membrane-bound inhibitor; BMD, bone mineral density; BMP2, bone morphogenetic protein 2; C-FOS, activation of FBJ murine osteosarcoma viral oncogene homolog; CK1, casein kinase 1; COL1A1/4A2/5A3, collagen, type I, alpha 1/type IV, alpha 2/type V, alpha 3; CRIM1, cysteine rich transmembrane BMP regulator 1 (chordin-like); CSF1R, colony stimulating factor 1 receptor; CTNNBIP1, catenin, beta interacting protein 1; CX43, connexin 43; CXCL11, chemokine (C-X-C motif) ligand 11; CXCR3, chemokine (C-X-C motif) receptor 3; DKK1/2, Dickkopf WNT sig­naling pathway inhibitor 1/2; DLX5, distal-less homeo­ box 5; DNMT1/3a/3b, DNA methyltransferase 1/3a/3b; DUSP2, dual specificity phosphatase 2; E2, 17β-estradiol; ETS1, v-ets avian erythroblastosis virus E26 oncogene homolog  1; FAK, focal adhesion kinase; FASLG, Fas ligand; FGF2, fibroblast growth factor 2; FOXO1, forkhead box O1; FZD1/4, frizzled family receptor 1/4; GALNT7, UDP-N-acetyl-alphaD-galactosamine:polypeptide N-acetylgalactosaminyl transferase 7 (GalNAc-T7); GSK3β, glycogen synthase kinase 3β; GTP, guanosine triphosphate; HAT, histone acetyltransferase; HDAC, histone deacetyl transferase; HOXA2/10, homeobox A2/10; HSC, hematopoietic stem cell; IFNβ/γ, interferon β/γ; IL1/4/6/7/8/10/11/13/17/18/23, interleukin 1/4/6/7/8/10/11/13/17/18/23; JNK, c-Jun N-terminal kinase; KREMEN2, kringle containing transmembrane protein 2; LEF1, lymphoid enhancer-binding factor 1; LEP, leptin; LRP5/6, low density lipoprotein receptor-related protein 5/6; MAFB, v-maf musculoaponeurotic fibrosarcoma oncogene homolog B; MAPK, mitogen-activated protein kinase; MECP2, methyl CpG binding protein 2 (Rett syndrome); MITF, microphthalmia-associated transcription factor; MSC, mesenchymal stem cell; MYF5, myogenic factor 5; MYOD, myogenic differentiation; NFATC1, nuclear factor of activated T cells, cytoplasmic, calcineurin-dependent 1; NFI-A, nuclear factor I-A; NFκB, nuclear factor κB; OPG, osteoprotegerin; OPN, osteopontin; OSCAR, osteoclast associated, immunoglobulin-like receptor; OSX, osterix; PBMC, peripheral blood mononuclear cell; PCP, planar cell planarity; PDCD4, programmed

Brought to you by | Tokyo Daigaku Authenticated Download Date | 6/7/15 3:27 AM

590      Vrtačnik et al.: Epigenetic mechanisms in bone cell death 4; poly-miRTs, polymorphisms in miRNA target sites; PPARγ, peroxisome proliferator-activated receptor γ; PU.1, spleen focus forming virus (SFFV) proviral integration oncogene; RANK, receptor activator of nuclear factor κB; RANKL, receptor activator of nuclear factor κB ligand; RHO, ras homolog family; RISC, RNA-induced silencing complex; ROS, reactive oxygen species; RUNX2, runt-related transcription factor 2; SAHA, suberoylanilide hydroxamic acid; SATB2, SATB homeobox 2; SFRP1/2, secreted frizzled-related protein 1/2; SIRT1-7, sirtuin 1-7; SLC39A1, solute carrier family 39 (zinc transporter), member 1; SMAD1/5, SMAD family member 1/5; SNP, single nucleotide polymorphism; SOCS1, suppressor of cytokine signaling 1; SOST, sclerostin; SOX9, SRY (sex determining region Y)-box 9; SPARC, secreted protein, acidic, cysteine-rich (osteonectin); SPRY1, sprouty homolog 1, antagonist of FGF signaling (Drosophila); STUB1, STIP1 homology and U-box containing protein 1, E3 ubiquitin protein ligase; TBP, TATA-box binding protein; TCF7, transcription factor 7 (T cell-specific, HMG-box); TGFβ, transforming growth factor β; TNFα, tumor necrosis factor α; TRAF6, tumor necrosis factor receptor activated factor 6; TSA, trichostatin A; WIF1, WNT inhibitory factor 1; WNT, wingless-type MMTV integration site family; WTX, Wilm’s tumor genes on chromosome X.

Introduction to bone biology Before epigenetic mechanisms, just a short overview of bone remodeling, bone cell differentiation and their regulation will be given. Bone remodeling is tightly coordinated and requires the synchronized action of osteoclasts, osteoblasts, bone-lining cells and osteocytes, in order to ensure that there are no major net changes in bone mass or mechanical properties after each remodeling cycle [1]. It takes place in a microanatomical structure that is separated from the bone marrow cavity by a canopy of cells but accessible through microcapillaries [2]. The remodeling process starts with the retraction of bone-lining cells covering the bone surface and the recruitment of osteoclast precursors to this remodeling site. Mature osteoclasts – large, multinucleated, short-lived, highly active cells attached to the bone surface – are responsible for the dissolution of the minerals and enzymatic degradation of the remaining organic matrix. After osteoclastmediated resorption is complete, collagen remnants are removed and the resorption lacunae is prepared for subsequent osteoblast-mediated bone formation in a process

that is still poorly understood. Bone formation starts with the differentiation of osteoblasts and laying down of the organic osteoid, consisting mainly of collagen type I. It is completed after osteoblast-mediated mineralization of the organic matrix. Finally, the resting bone surface covered by bone-lining cells belonging to the osteoblast lineage is re-established [3]. Osteocytes – terminally differentiated osteoblasts entombed within the bone matrix that account for almost 95% of all cells in the mature bone tissue – form a network of canaliculi within the mineralized bone. These mechanosensing cells are thought to detect mechanical strain and associated bone microdamage and to respond by initiating bone resorption and the regulation of bone remodeling. Data show that hormonal stimuli also trigger bone remodeling via osteocytes [4]. The differentiation of osteoblasts and osteoclasts from their precursors is an inherent part of bone remodel­ ing. The transcriptional regulator RUNX2 is indispensible in all stages of osteoblastogenesis and bone formation. It also represents an intersection of several signaling pathways regulating osteoblast differentiation, including BMP, WNT, Notch, Hedgehog and FGF [5]. By interacting with many transcriptional activators, repressors and other coregulatory proteins, RUNX2 can regulate the expression of several osteoblast-specific genes either positively or negatively. A further transcription factor necessary for osteoblast-specific gene expression, osteoblastogenesis and bone formation is OSX [6]. The differentiation factors responsible for osteoclast maturation are, however, much more diverse and include PU.1 and MITF, as well as CSF1R, C-FOS and RANK. Osteo­ clast bone-resorptive activity is mediated by RANKL through the activation of NFκB and NFATC1, two principal transcriptional mediators [7]. The multiple stages, together with key transcription factors and signaling pathways in osteoblast and osteoclast differentiation, are summarized in Figure 1. Controlled and balanced differentiation of osteoblasts and osteoclasts is a prerequisite of normal bone remodeling and skeletal health. Deregulation of these processes can lead to several metabolic bone diseases, including renal osteodystrophy, Paget’s disease, osteo­ petrosis, rickets and osteoporosis, among others. Osteoporosis, by far the most prevalent among them, is characterized by a decrease in bone mineral density (BMD) and deterioration in bone microarchitecture [1]. Osteoarthritis, however, is associated with degraded articular cartilage and subchondral bone sclerosis. Like osteoporosis, it is a highly prevalent disorder with a considerable amount of controversy surrounding the tissue of disease onset, some supporting the idea that

Brought to you by | Tokyo Daigaku Authenticated Download Date | 6/7/15 3:27 AM

Vrtačnik et al.: Epigenetic mechanisms in bone      591

Figure 1 Osteoblast and osteoclast differentiation. Left: The osteoblast lineage derives from MSCs under the control of the transcriptional regulator RUNX2 [5]. The multipotent differential capacity of MSCs can also give rise to chondrocyte, adipocyte, myocyte and other cell lineages, utilizing lineage-specific transcription factors SOX9, PPARγ2 and MYOD/MYF5, respectively [8, 9]. RUNX2 is indispensible in all stages of osteoblast differentiation. After reaching maturity, three different potential fates await osteoblasts. Cells that become entombed within the bone matrix are called osteocytes, bone-lining cells cover all bone surfaces while the remainder undergo apoptosis [4]. Right: Osteoclasts, however, originate from the HSC lineage. Differentiation factors responsible for osteoclast maturation include PU.1 and MITF, as well as CSF1R, C-FOS and RANK. Mature osteoclasts capable of resorptive activity are formed only after the fusion of several precursor cells and attachment to bone through a process that is primary stimulated by RANKL [7].

it is a bone disease that also affects cartilage, and vice versa [10]. Due to the high prevalence of osteoporosis and osteo­arthritis as well as the lack of studies in other metabolic bone disorders, the present review is focused on these two pathologies. Similarly, since the WNT/β-catenin signaling pathway and RANK/RANKL/OPG system are, respectively, the most well described and perhaps the most significant stimulators of bone formation and bone resorption, only these two pathways are discussed, briefly, below.

 ANK/RANKL/OPG system in osteoclast R regulation and bone resorption The RANK/RANKL/OPG system is one of the most important regulators of bone resorption and remodeling and, therefore, an obvious candidate for the studies of epigenetic modifications related to bone pathologies. RANK, located on the surface of osteoclasts and their precursors, and its ligand RANKL are essential for the formation,

differentiation, activity and survival of osteoclasts. RANKL is produced by cells of the osteoblast lineage as well as other cell types in both soluble and membranebound forms. The binding of RANKL to RANK, results in the activation of transcription factors NFκB and NFATC1 and the expression of osteoclastogenic genes. In contrast, OPG, secreted by osteoblasts and a few other cell types, functions as a decoy receptor by binding to RANKL, thereby preventing the activation of RANK. This inhibition of RANKL, which leads to the rapid arrest of osteoclast formation, activation and survival, is crucial for the suppression of bone resorption and maintenance of bone mass [11]. Pro-inflammatory cytokines secreted by different immune cells also notably influence osteoclast activity and bone resorption by either targeting osteoclasts directly or modulating the RANK/RANKL/ OPG system [12]. The importance of the RANK/RANKL/ OPG system in maintaining bone homeostasis is further supported by the therapeutic success of denosumab, a monoclonal antibody directed against RANKL, used in patients with osteoporosis and cancer patients with treatment-induced bone loss or metastases [11, 13]. The RANK/RANKL/OPG system and its cytokine modulators are depicted in Figure 2.

 NT/β-catenin signaling pathway in W ­osteoblast regulation and bone formation The very precise regulation of the intensity, amplitude and duration of WNT/β-catenin signaling is necessary for proper skeletal development and bone remodeling, indicating a likely connection with the epigenetic control of gene and protein expression. There are several WNT signaling pathways, including the non-β-catenin PCP pathway, calcium ion signaling, RHO family GTPase pathways, and the JNK pathways; however, it is the WNT/βcatenin signaling pathway that is the most crucial for multiple stages of osteoblast differentiation, their proliferation and survival. In the latter, WNTs cause an increase in β-catenin levels and the activation of transcription factors TCF7 and LEF1. Their transcriptional targets include the aforementioned OPG and RUNX2, linking the WNT/β-catenin signaling pathway with the inhibition of osteoblast-mediated osteoclastogenesis and the stimulation of osteoblast proliferation and survival, respectively. The presence of several neutralizing extracellular factors that directly bind WNTs (e.g., SFRP1, WIF1) or their coreceptors (e.g., SOST, DKK1/2) increases the complexity of the WNT/β-catenin signaling pathway but, at the same time, represents an attractive opportunity for therapeutic

Brought to you by | Tokyo Daigaku Authenticated Download Date | 6/7/15 3:27 AM

592      Vrtačnik et al.: Epigenetic mechanisms in bone

Figure 2 RANK/RANKL/OPG system. The RANK/RANKL/OPG system is essential for the formation and differentiation of osteoclasts, their resorptive activity and survival. The binding of RANKL to RANK results in the recruitment of TRAF6, which activates various protein kinase pathways and transcription factors like NFκB. The activated NFκB up-regulates the expression of C-FOS, which subsequently interacts with NFATC1 to induce the expression of osteoclastogenic genes. Conversely, OPG prevents the activation of RANK by binding RANKL [11]. Pro-inflammatory cytokines secreted by different immune cells – including activated T cells, B cells, macrophages, mast cells and natural killer cells – also modulate the RANK/ RANKL/OPG system. TNFα, IL1, IL6, IL8, IL11 and IL17 are osteoclastogenic cytokines promoting RANKL-mediated osteoclast differentiation and activity, while the anti-osteoclastogenic cytokines IFNβ, IFNγ, IL4, IL10, IL13 and IL18 inhibit osteoclasts through the RANK/RANKL/OPG system. Certain cytokines can exert opposite effects on osteoclasts (e.g., IL7 and IL23) [12].

intervention [14]. In particular, SOST has been identified as a key negative regulator of bone mass and formation. Its inhibition using a monoclonal antibody has a profound anabolic effect on the skeleton. Since it is secreted primarily in bone, specifically by osteocytes, and facilitates a novel anabolic approach, it represents a promising therapeutic target in the case of bone-related disorders such as osteoporosis [15, 16]. Due to the importance of WNT signaling pathways in osteoblast regulation and bone formation, it is not surprising that certain known bone-influencing stimuli exert their effect at least in part through WNT signaling pathways. These include estrogen and parathyroid hormone, as well as mechanical loading and pro-inflammatory cytokines, among others [17–19]. The WNT/β-catenin signaling pathway is schematically shown in Figure 3. All of the aforementioned signaling pathways play an important role in the regulation of bone metabolism by fine tuning the expression of thousands of target genes. Since posttranslational histone modifications, microRNAs (miRNAs) and DNA methylation act as important regulators of gene expression, they could also interfere significantly with bone homeostasis. Furthermore, these epigenetic mechanisms are increasingly linked to aging, providing

additional support for their possible association with lateonset disorders like osteoporosis and osteoarthritis. Following a brief overview of epigenetic mechanisms, the complex intertwinement of epigenetic processes and bone remodeling will be discussed, revealing potential new diagnostic and therapeutic targets in skeletal diseases. The articles included in the review have been critically selected from the results of a PubMed database search using combinations of the keywords osteoblast differentiation, osteoclast differentiation, histone acetylation, HDAC, SIRT1, miRNA, DNA methylation, osteoporosis and osteoarthritis.

Epigenetic mechanisms ­controlling gene expression in bone cells  osttranslational histone modifications in P regulation of gene expression Of the three epigenetic mechanisms, posttranslational histone modifications and accompanying histone-modifying enzymes form probably the biggest and most complex

Brought to you by | Tokyo Daigaku Authenticated Download Date | 6/7/15 3:27 AM

Vrtačnik et al.: Epigenetic mechanisms in bone      593

add, while histone deacetyl transferases (HDACs) remove, acetylation marks. HATs are classified as belonging to one of two classes, HAT A and HAT B, and several subclasses, depending on the mechanism of their action and cellular localization. HDACs are divided into four groups based on their structural and functional similarities; class I includes HDAC 1, 2, 3 and 8, class II HDAC 4–7, 9 and 10, class III SIRT 1–7 and class IV solely HDAC11. Due to the fact that many transcriptional coactivators and corepressors possess HAT or HDAC activity, respectively, or associate with such enzymes, the balance between acetylation and deacetylation of histones that harbor target genes significantly influences their expression [20, 22]. Figure 3 WNT/β-catenin signaling pathway. The WNT/β-catenin signaling pathway is the most important among WNT signaling pathways for multiple stages of osteoblast differentiation, their proliferation and survival. WNTs bind to cell-surface FZD receptors and LRP5/6 coreceptors, which causes the release of β-catenin from its destruction complex containing AXIN1/2, APC, CK1, GSK3β and WTX. Increased levels of β-catenin facilitate its translocation from cytoplasm to the nucleus where it activates transcription factors TCF7 and LEF1 and regulates the expression of target genes including OPG and RUNX2 [14]. Numerous bonestimulatory factors exert their effects, at least in part, through WNT signaling pathways, including estrogen and parathyroid hormone [17, 18]. Mechanical loading, a well-known bone-forming stimulus, also causes the activation of the WNT/β-catenin signaling pathway by reducing the secretion of WNT inhibitors, most prominently SOST, from osteocytes [18]. Conversely, the pro-inflammatory cytokine TNFα induces the expression of the WNT inhibitor DKK1, resulting in impaired bone formation in a mouse model of rheumatoid arthritis [19].

regulatory entity. As far as is presently known, an octamer of four different canonical histone molecules (H2A, H2B, H3 and H4) wrapped by a 147-bp segment of DNA forms a nucleosome, the basic unit of chromatin. Flexible N-terminal tails of core histones that protrude from the nucleosome are subjected to various posttranslational modifications, including acetylation, methylation, phosphorylation, ubiquitination, sumoylation, ADP ribosylation, deimination and a noncovalent proline isomerization [20]. It has been shown that euchromatin, a more relaxed, actively transcribed state of DNA, is characterized by high levels of acetylation and trimethylated H3K4, H3K36 and H3K79, while low levels of acetylation and high levels of H3K9, H3K27 and H4K20 methylation are indicative of a more condensed, transcriptionally inactive heterochromatin [21]. Most histone posttranslational modifications are dynamic and regulated by families of enzymes that promote or reverse specific modifications. For example, histone acetyltransferases (HATs)

MicroRNAs in regulation of gene expression miRNAs are small – approximately 21 nucleotides long – non-coding RNA molecules. Mature miRNAs are incorporated into the RISC in order to facilitate binding to the target mRNA, usually within its 3′ UTR region. A perfect match between miRNA and the target results in the degradation of the targeted transcript. More commonly, miRNAs bind imperfectly to their targets, causing translational repression without the cleavage of mRNA [23]. In general, miRNAs are considered predominantly as negative regulators of gene expression involved in physiological and pathological processes, even though some of them have been shown to activate translation under certain conditions [24].

 NA methylation and DNMTs in r­ egulation of D gene expression DNA methylation is a reversible, covalent modification of the 5′-carbon of a cytosine residue, which results in the production of 5-methyl-cytosine. In mammals, cytosine methylation is restricted to residues located 5′ to a guanosine (commonly referred to as CpGs), which occur more frequently in regions known as CpG islands. The methylation of DNA can either inhibit or facilitate the binding of proteins to the DNA molecule. In general, DNA methylation is associated with gene repression. As DNA methylation patterns can be preserved during DNA replication and mitosis, the associated repressed state can be inherited. Enzymes responsible for the transfer of a methyl group from S-adenosyl methionine to DNA include DNMT1, DNMT3a and DNMT3b. DNMT1 is the so-called maintenance DNMT, copying the DNA methylation pattern of the parent DNA strand onto the daughter

Brought to you by | Tokyo Daigaku Authenticated Download Date | 6/7/15 3:27 AM

594      Vrtačnik et al.: Epigenetic mechanisms in bone strand during replication, although it also possesses some de novo DNMT activity. DNMT3a and DNMT3b are the main enzymes responsible for de novo DNA methylation of unmethylated CpGs and for the establishment of DNA methylation marks during embryonic development. In addition, they, in all probability, assist DNMT1 in maintaining the already established DNA methylation patterns. Demethylation, however, can occur either passively, by reducing maintenance methylation following DNA replication, or actively, although the exact molecular mechanism of the latter has still not been fully elucidated [20, 21].

 he role of histone-modifying T enzymes in bone remodeling As the large number of histone marks and their combinations increase the complexity of transcriptional regulation through histone-modifying enzymes, we will concentrate only on the importance of histone acetylation in bone remodeling.

 istone acetylation in osteoblast H differentiation Several studies have used HDAC inhibitors to elucidate the influence of general histone hyperacetylation on osteoblast differentiation and gene expression. There are many different HDAC inhibitors, including TSA, SAHA, MS-275, sodium butyrate, and valproic acid; some of these, in addition to being in use with regard to certain neurological conditions, also have great potential as anti-cancer drugs and therapeutics for inflammatory and cardiac diseases. TSA and SAHA are potent inhibitors of class I and II HDACs, MS-275 inhibits class I HDACs and HDAC9 – with the exception of HDAC8 and a preference for HDAC1 – while sodium butyrate and valproic acid also inhibit class I HDACs, albeit with somewhat lower potency, especially towards HDAC8 [25, 26]. In vitro studies have revealed that the inhibition of either class I and II HDACs or class I HDACs alone accelerates osteoblast maturation, matrix mineralization and the expression of genes associated with osteoblast differentiation (e.g., type I collagen, bone sialoprotein, osteopontin, osteocalcin, ALPL, OSX and RUNX2). This was achieved, at least in part, through the up-regulation of RUNX2 transcriptional activity. Early changes in gene expression also included two WNT receptor genes, FZD1 and FZD4. TSA-induced minimal

cytotoxicity in osteoblasts at low concentrations still sufficient to cause histone hyperacetylation, while MS-275 and sodium butyrate enhanced osteoblast viability and proliferation. Valproic acid was also shown to enhance the viability and proliferation of osteoblasts while inhibiting the proliferation of mesenchymal stem cells (MSCs) without affecting their survival [27–29]. Interestingly, RANKL expression was also enhanced by the TSA-mediated increase in the acetylation of histones H4 and H3 at the RANKL promoter [30]. In vivo studies in animal models of bone loss revealed stimulatory effects on bone regeneration after MS-275 treatment, while studies using healthy animals resulted in bone loss during treatment with SAHA or valproate [31–33]. Furthermore, human epidemiological studies in patients with different neurological conditions treated with valproate tend to provide support for there being negative effects of HDAC inhibitors on bone, as evidenced by decreased BMD and increased fracture risk in children and adults [34]. It has been suggested that this discrepancy could be due to various HDAC inhibitors targeting individual HDACs with different specificity and affinity as well as their broad range of action, which can include the inhibition of other enzymes besides HDACs. In the case of studies involving animals, the selection of animal models could also play an important part, as certain mice strains were shown to be resistant to bone effects caused by HDAC-inhibitor treatment [33]. Human studies also featured many confounding factors, since they involved very different patient cohorts, including patients with decreased physical activity [34]. Another group of studies focused on the expression of osteocalcin in order to elucidate the role of histone acetylation in osteoblastogenesis. Osteocalcin is a very abundant bone-specific protein that is able to bind calcium and, consequently, hydroxyapatite, and is essential for proper bone mineralization. Serum osteocalcin levels are used as a bone formation marker, while its expression in cell cultures enables osteoblast differentiation status and activity to be determined. These studies showed that the osteocalcin promoter is associated with the acetylated histones H3 and H4 when transcriptionally active, while very low levels of acetylated H3 and H4 accompany an inactive osteocalcin gene [35, 36]. It was also shown that the transcriptional coactivator and HAT p300 interacted with RUNX2 at the osteocalcin promoter in such a way as to stimulate its expression, although it was not responsible for the increased levels of acetylation. It appears that other proteins that contain HAT activity and associate with p300, e.g., PCAF, could potentially increase H3 and H4 acetylation [37]. Given that the active expression of other bone-related genes was also reflected in

Brought to you by | Tokyo Daigaku Authenticated Download Date | 6/7/15 3:27 AM

Vrtačnik et al.: Epigenetic mechanisms in bone      595

the presence of acetylated H3 and H4 at their respective loci, it is plausible that the acetylation of histones associated with bone-related gene promoters is functionally coupled to the chromatin remodeling events that mediate the regulation of gene expression during osteoblast differentiation [38]. Accordingly, decreased histone acetylation was associated with the negative regulation of osteoblast differentiation and concomitant repression of osteocalcin expression. Elevated levels of NFATC1, e.g., interacted with HDAC3 at the osteocalcin promoter and prevented the acetylation of both H3 and H4 [39]. In addition, HDAC3 was also shown to repress activation of the osteocalcin promoter by interacting with RUNX2, thereby antagonizing its transcriptional activity [40, 41]. Lamour et al. confirmed the suppressive role of HDAC3 by showing it is also highly likely to be responsible for the deacetylation of H3 at the bone sialoprotein promoter, thus preventing its gene expression [42]. Interestingly, the conditional knockout of HDAC3 in mice had no effect on osteocalcin levels. Furthermore, it reduced trabecular and cortical bone mass due to impaired osteoblast function associated with increased DNA damage [43, 44]. This large discrepancy between the results of in vitro and in vivo studies of HDAC3 action is probably due to differences in duration and the extent of HDAC3 regulation as well as the maturity of the osteoblasts used. However, TGFβ, a well-known negative regulator of bone formation, inhibited osteoblast differentiation and osteocalcin expression by recruiting HDAC4 and 5 to RUNX2, resulting in the deacetylation of H4 at the osteocalcin promoter [45]. Interestingly, HDAC4 and 5 could also directly decrease and p300 directly increase RUNX2 acetylation, thereby reducing or enhancing, respectively, its protein stability and transcriptional activity [46]. Besides HDAC3, 4 and 5, HDAC1 was also identified as an important regulator of osteoblast differentiation. Lee et al. discovered that the hyperacetylation of H3 and H4 at promoters of OSX and osteocalcin in differentiated osteoblasts was due to a decrease in the recruitment of HDAC1 and the enhanced association of p300 at those promoters. Furthermore, the knockdown of HDAC1 alone was enough to promote the expression of osteogenic genes and induce osteogenesis [47].

 irtuin 1 as an important regulator of bone S homeostasis SIRT1 is probably one of the most studied HDACs due to its proposed role in linking calorie restriction and life­ span extension. It also turned out to be a major regulator of bone mass. Cohen-Kfir et al. demonstrated that SIRT1

repressed the expression of SOST by deacetylating H3K9 at the SOST promoter. Accordingly, SIRT1 haplo-insufficient female mice exhibited substantially increased SOST levels and reduced bone formation and mass. In addition, differentiated MSCs derived from SIRT1 haplo-insufficient mice showed reduced osteoblast activity, while the same cells displayed increased adipogenesis and PPARγ expression when adipogenesis was induced [48]. A reciprocal relationship between the level of SIRT1, on one hand, and adipocyte differentiation and PPARγ expression, on the other, has also been observed by others in vitro and in vivo [49–51]. It is important to note that SIRT1 action is not limited to epigenetic mechanisms as it exerts its multiple activities by interacting not only with histones but also with numerous transcription factors, enzymes and other protein species [52]. Additional studies, for example, have shown that SIRT1 promotes the activation of RUNX2 and the suppression of NFκB signaling, thus stimulating osteoblastogenesis while also inhibiting osteoclastogenesis; this indicates that it plays an important role in the coupling of bone formation and resorption with a possible contribution to bone quality and a reduction in bone frailty (Figure 4) [51, 53–55].

 istone acetylation in osteoclast H differentiation In contrast to osteoblasts, the inhibition of either class I and II HDACs or class I HDACs alone suppressed osteoclast differentiation in vitro [56–59]. Class I HDAC inhibitors targeting predominantly HDAC1 and HDAC2 also decreased bone destruction in animal models of bone loss in vivo [57, 59]. Osteoclasts exposed to TSA, SAHA or sodium butyrate treatment exhibited enhanced apoptosis and reduced activation of NFκB and MAPK signaling [56, 58]. In addition, a reduction in the expression of transcription factors C-FOS and NFATC1, osteoclast-specific genes cathepsin K, calcitonin receptor and OSCAR, and even the induction of IFNβ, an inhibitor of osteoclastogenesis, were observed after the inhibition of class I and II HDACs or class I HDACs alone [56, 57, 59, 60]. In support of this, Pham et al. were able to inhibit osteoclast differentiation in vitro by suppressing the expression of HDAC3 using appropriate short hairpin RNAs. Conversely, the suppression of HDAC7 resulted in enhanced osteoclast formation. This was due to the fact that HDAC7 was able to inhibit osteoclast differentiation by repressing the MITF transcription factor, in all likelihood by means of a deacetylation-independent mechanism [61]. Furthermore, RANKL was able to increase the transcriptional activity of NFATC1 by stimulating the

Brought to you by | Tokyo Daigaku Authenticated Download Date | 6/7/15 3:27 AM

596      Vrtačnik et al.: Epigenetic mechanisms in bone discussed hereinafter, placed HDAC5 downstream from a miRNA associated with primary osteoporosis in adolescents [63, 64]. Nevertheless, little is known about the involvement of HATs and HDACs and their acetylation marks in the pathogenesis of bone disorders. When taking into account all the possible posttranslational modifications potentially present at individual nucleosomes, there is great potential for the discovery of biomarkers associated with bone pathologies.

 he role of miRNA in bone T remodeling Figure 4 Multiple roles of SIRT1 in bone. SIRT1 has many roles in cellular physiology; it is involved in nutrient sensing, stress response, survival/death decisions, the regulation of energy metabolism, autophagy, inflammation and senescence. SIRT1 is also believed to interfere with fundamental pathogenic processes that underlie aging, e.g., cellular degeneration and neoplasia [52]. In the bone environment, SIRT1 causes the deacetylation of histones at the SOST promoter, thus repressing SOST expression and preventing its negative regulation of osteoblast activity. In addition, SIRT1 promotes the activation of RUNX2 and suppression of NFκB signaling, thereby both stimulating osteoblastogenesis and inhibiting osteoclastogenesis [51, 53–55]. It is also positively linked to estrogen signaling, an essential regulator of bone mass, both upstream and downstream of the estrogen receptor. Conversely, SIRT1 was shown to be a negative regulator of adipocyte differentiation and PPARγ expression [49–51].

acetylation of the NFATC1 protein via the HATs PCAF and p300 during osteoclastogenesis. Conversely, HDACs 5 and 6 showed the ability to deacetylate NFATC1, thereby down-regulating its transcriptional activity and diminishing osteoclast differentiation [62]. In conclusion, there is no doubt that both HATs and HDACs play substantial roles in regulating osteoblast and osteoclast differentiation, transcriptional activity and survival, but the exact mechanisms are still unclear. It is important that future studies concentrate on the function of individual enzymes instead of using broad acting inhibitors as well as on the balance between acetylating and deacetylating enzymes since it is likely to be essential for maintaining intra- and inter-cell homeostasis. Furthermore, HATs and HDACs target several important transcription factors and other non-histone proteins and can, in certain cases, exert their biological function without enzymatic activity. A genetic study uncovered that HDAC5 could be associated with hip BMD, while another study,

Various cancers often exhibit unique miRNA-expression profiles of up- and down-regulated miRNAs, opening up the possibility of using them as biomarkers for cancer diagnosis, prognosis, and response to therapy [65, 66]. The most promising epigenetic biomarkers in lung, colorectal and prostate cancers have recently been reviewed by Sandoval et  al. [67]. Increasing evidence indicates a similar role for miRNA in metabolic bone diseases.

miRNAs in osteoblast differentiation In recent years, more than 30 papers have investigated the role of miRNA in osteoblast differentiation. They have identified several miRNAs – namely, miR-23a, miR30a–d, miR-34c, miR-133a, miR-135a, miR-137, miR-204, miR-205, miR-211, miR-217, miR-335, miR-338, miR-433 and miR-3077-5p – that impede osteoblast differentiation by directly targeting the master osteogenic transcription factor RUNX2 with differences in efficacy and quantity depending on the stage of osteoblast lineage progression [68–76]. Conversely, miR-2861 and miR-3960 have been shown to be able to indirectly stimulate RUNX2 expression by suppressing the translation of its inhibitors HDAC5 and HOXA2, respectively, thus promoting osteoblast ­differentiation [64, 77]. Furthermore, RUNX2 was shown to positively and negatively regulate the expression of particular miRNAs [74, 77]. To a certain extent, OSX is also under miRNA control, forming a unique autoregulatory feedback loop with miR-93. It has been shown that miR-93 inhibited osteoblast mineralization by directly targeting OSX [78]. The WNT/β-catenin signaling pathway in human osteoblasts regulates, while also being under the control of, specific miRNAs. By directly targeting known WNT antagonists (e.g., DKK1, DKK2, SFRP2 and SOST), miR-29a,

Brought to you by | Tokyo Daigaku Authenticated Download Date | 6/7/15 3:27 AM

Vrtačnik et al.: Epigenetic mechanisms in bone      597

miR-218 and miR-335-5p were able to enhance the WNT/βcatenin signaling pathway and potentiate osteoblast differentiation [79–81]. Activation of the WNT/β-catenin signaling pathway was also observed following the suppression of the expression of APC, an integral part of the β-catenin destruction complex, by miR-27 and miR142-3p [82, 83]. Similarly, miR-29b was indicated to target CTNNBIP1, an inhibitor of β-catenin-mediated transcription, resulting in enhanced osteogenesis [84]. Together with those previously described, additional positive and negative miRNA regulators of osteoblast differentiation and their respective targets are presented in Table 1. In the described differentiation experiments, researchers used both mesenchymal and more committed preosteoblast cell lines; these were predominantly of mouse origin, often supported by primary mouse- or human bone marrow-derived MSCs or mouse calvarial osteoblasts. When also taking into account the variability of the differentiation media used, the diversity of RNA isolation methods and differences in miRNA screening procedures, it is not surprising that there is little overlap between the results of the aforementioned studies. However, each miRNA targets several different transcripts and most mRNAs are regulated by a number of miRNAs. In this setting, it is expected that a vast network of miRNAs takes part in the precise regulation of the complex process that is osteoblast differentiation.

miRNAs in osteoclast differentiation In comparison to osteoblast differentiation, the involvement of miRNAs in osteoclast differentiation is markedly less researched (Table 2). Sugatani et al. showed that both the overexpression of pre-miR-223 and antisense inhibitors of miR-223 suppressed osteoclastogenesis, suggesting that expression levels of miR-223 that were either too high or too low were not appropriate for efficient osteoclast differentiation. They proposed a mechanism by which miR-223 suppressed NFI-A, an indirect inhibitor of CSF1R, thus stimulating osteoclast differentiation and the expression of transcription factors such as MITF, C-FOS and PU.1 [103, 105]. In addition, miR-21 was shown to promote osteoclast differentiation and protect osteoclasts from apoptosis. By targeting PDCD4, miR-21 derepressed C-FOS, resulting in increased osteoclastogenesis and miR-21 expression via a positive feedback mechanism [99]. Protection against estrogen-induced apoptosis during osteoclastogenesis was, however, achieved by down-regulating the inducer of apoptosis, Fas ligand, another miR-21 target; this suggests that the anti-osteoclastic activity of estrogen

could be counteracted by miR-21 overexpression [100]. Conversely, miR-155 efficiently blocked the activation of the osteoclast transcriptional program and repressed osteoclastogenesis by directly targeting MITF [104]. The stimulatory effects of miR-127, miR-136 [An et al., Unpublished manuscript], miR-133a [101] and miR-148a [102], as well as the inhibitory effects of miR-503 [Chen et al., Unpublished manuscript], on osteoclastogenesis were also observed and are described in the following section.

miRNAs associated with osteoporosis Most studies analyzing the associations between metabolic bone diseases and epigenetic changes have concentrated on differences in miRNA expression and function. Among these, studies using bone tissue as their main study material have focused almost entirely on osteoporosis, with the exception of one on osteoarthritis. The latter identified 30 miRNAs differentially expressed in osteoarthritic bone when compared to normal tissue, with miR27a, miR-34b, miR-98 and miR-330 showing the greatest fold change. Interestingly, miR-9 and miR-98 were up-regulated in both osteoarthritic bone and cartilage tissue and could be functionally linked to inflammation [106]. Profiling of bone tissue and bone marrow-derived MSCs from ovariectomized mice, however, revealed increased expression of miR-127, miR-133a, miR-133b, miR-136, miR-206, and miR-378a, together with decreased expression of miR-204 in bone tissue, while MSCs showed overexpression of miR-705 and miR-3077-5p and the suppression of miR-21 after ovariectomy [76, 86, An et al., Unpublished manuscript]. The high expression of miR133a, miR-133b, miR-206 and miR-378a is in line with the already described studies; however, this is not the case for miR-204, since it has previously been shown to inhibit osteoblast differentiation. The discrepancy in this instance could be due to differences in the starting material, with one study using osteoblastic cell lines, the other bone tissue. Based on the predicted targets of identified miRNAs and the associated mRNA expression profile, it was suggested that the PPARγ and CREB pathways were important mediators of change in bone tissue after ovariectomy. In addition, miR-127 and miR-136 were shown to suppress osteoblast differentiation and osteocyte function and survival while at the same time promoting osteoclast differentiation [An et al., Unpublished manuscript]. Similarly, miR-705 and miR-3077-5p were shown to inhibit osteogenic differentiation and promote the adipogenic differentiation of MSCs by directly targeting HOXA10 and RUNX2, respectively. Both target genes are involved in the

Brought to you by | Tokyo Daigaku Authenticated Download Date | 6/7/15 3:27 AM

598      Vrtačnik et al.: Epigenetic mechanisms in bone Table 1 miRNAs involved in the regulation of osteoblast differentiation. Positive miRNA regulators of osteoblast differentiation miRNA



Target mRNA symbol 

Target mRNA name



References

miR-20a



PPARγ; BAMBI; CRIM1



Peroxisome proliferator-activated receptor γ; BMP and activin   membrane-bound inhibitor; cysteine-rich transmembrane BMP regulator 1 (chordin-like)

[85]

miR-21



SPRY1



Sprouty homolog 1, antagonist of FGF signaling (Drosophila)  

[86]

miR-27; miR-142-3p



APC



Adenomatous polyposis coli



[82, 83]

miR-29a



DKK1; KREMEN2; SFRP2; SPARC



Dickkopf WNT signaling pathway inhibitor 1; kringle   containing transmembrane protein 2; secreted frizzled-related protein 2; secreted protein, acidic, cysteine-rich (osteonectin)

[79, 87]

miR-29b



COL1A1; COL5A3; COL4A2; HDAC4; TGFβ3; ACVR2A; CTNNBIP1; DUSP2



Collagen, type I, alpha 1; collagen, type V, alpha 3; collagen,   type IV, alpha 2; histone deacetyl transferase 4; transforming growth factor β3; activin A receptor, type IIA; catenin, beta interacting protein 1; dual specificity phosphatase 2

[84]

miR-29c



SPARC



Secreted protein, acidic, cysteine-rich (osteonectin)



[87]

miR-210



ACVR1B



Activin A receptor, type IB



[88]

miR-218



SOST; DKK2; SFRP2  

Sclerostin; Dickkopf WNT signaling pathway inhibitor 2; secreted frizzled-related protein 2



[80]

miR-335-5p



DKK1



Dickkopf WNT signaling pathway inhibitor 1



[81]

miR-764-5p



STUB1



STIP1 homology and U-box containing protein 1, E3 ubiquitin   protein ligase

[89]

miR-2861



HDAC5



Histone deacetyl transferase 5



[64]

miR-3960



HOXA2



Homeobox A2



[77]

Negative miRNA regulators of osteoblast differentiation miRNA



Target mRNA symbol 

Target mRNA name



References

miR-23a



RUNX2; SATB2



Runt-related transcription factor 2; SATB homeobox 2



[68, 70, 74]

miR-24-2; miR-27a



SATB2



SATB homeobox 2



[74]

miR-30a–d



RUNX2; SMAD1



Runt-related transcription factor 2; SMAD family member 1



[68, 70, 72]

 iR-34c; miR-133a; miR-137;   m miR-204; miR-205; miR-211; miR-217; miR-335; miR-338; miR-433; miR-3077-5p

RUNX2



Runt-related transcription factor 2



[68–71, 73, 75, 76]

miR-93

OSX



Osterix



[78]



miR-135a



RUNX2; SMAD5



Runt-related transcription factor 2; SMAD family member 5



[68–70]

miR-138



FAK



Focal adhesion kinase



[90]

miR-141; miR-200a



DLX5



Distal-less homeobox 5



[91]

miR-155



SOCS1



Suppressor of cytokine signaling 1



[92]

miR-182



FOXO1



Forkhead box O1



[93]

miR-206



CX43



Connexin 43



[94]

miR-208



ETS1



v-ets avian erythroblastosis virus E26 oncogene homolog 1



[95]

miR-214



ATF4



Activating transcription factor 4



[96]

miR-370



BMP2; ETS1



Bone morphogenetic protein 2; v-ets avian erythroblastosis virus E26 oncogene homolog 1



[97]

miR-378



GALNT7



UDP-N-acetyl-alpha-D-galactosamine:polypeptide N-acetylgalactosaminyltransferase 7 (GalNAc-T7)



[98]

miR-705



HOXA10



Homeobox A10



[76]

Brought to you by | Tokyo Daigaku Authenticated Download Date | 6/7/15 3:27 AM

Vrtačnik et al.: Epigenetic mechanisms in bone      599 Table 2 miRNAs involved in the regulation of osteoclast differentiation. Positive miRNA regulators of osteoclast differentiation miRNA



Target mRNA symbol



Target mRNA name



References

miR-21   miR-127; miR-136  

PDCD4; FASLG Not reported

   

Programmed cell death 4; Fas ligand Not reported

   

miR-133a



miR-148a miR-223

   

CXCL11; CXCR3; SLC39A1   (in silico predicted) MAFB   NFI-A  

[99, 100] [An et al., Unpublished manuscript] [101]

Chemokine (C-X-C motif) ligand 11; chemokine (C-X-C motif)   receptor 3; solute carrier family 39 (zinc transporter), member 1 v-maf musculoaponeurotic fibrosarcoma oncogene homolog B   Nuclear factor I-A  

[102] [103]

Negative miRNA regulators of osteoclast differentiation miRNA



Target mRNA symbol



Target mRNA name



References

miR-155 miR-223 miR-503

     

MITF Not reported RANK

     

Microphthalmia-associated transcription factor Not reported Receptor activator of nuclear factor κB

     

[104] [105] [Chen et al., Unpublished manuscript]

BMP signaling pathway, potentially providing a synergistic effect on MSC differentiation. Furthermore, increased levels of TNFα and reactive oxygen species (ROS), associated with an estrogen deficiency, significantly increased the expression of miR-705 and miR-3077-5p through the activation of NFκB pathway. The shift of cell lineage commitment in MSCs from osteoblasts towards adipocytes was suggested as being an important event in the pathogenesis of osteo­porosis [76]. Conversely, miR-21 was shown to directly target SPRY1, a key negative regulator of FGF and MAPK signaling pathways, resulting in enhanced osteoblast differentiation and bone formation. Decreased miR-21 in MSCs from osteoporosis patients and ovariectomized mice was also most likely to be due to high levels of TNFα, associated with estrogen deficiency (Figure 5) [86]. Another recent study showed that high miR-214 expression in bone tissue correlated with reduced bone formation in human subjects and mouse models. The researchers showed that miR-214 suppressed osteoblast activity and matrix mineralization by directly targeting ATF4, which is responsible for promoting osteoblast-specific gene expression, amino acid uptake and type I collagen synthesis. Moreover, the osteoblast-specific inhibition of miR-214 significantly improved bone formation and bone mass in ovariectomy-induced osteoporosis and hind limb-unloading mouse models. In contrast, bone resorption was not affected by miR-214 [96]. A different study identified the repressed expression of miR-2861 due to a mutation in pre-miR-2861 as a contributing factor in primary osteoporosis in two related adolescents. It was shown that miR-2861 promoted

osteoblast differentiation by directly targeting HDAC5, which is responsible for reducing RUNX2 protein stability and transcriptional activity. This was confirmed in vivo in a mouse model where the absence of miR-2861 significantly reduced BMD, bone formation, osteoblast number and RUNX2 protein levels while increasing HDAC5 protein expression. Bone resorption and osteoclast activity were unaffected. According to the study’s results, the discovered mutation in pre-miR-2861 is, in all likelihood, a rare variant and thus unlikely to be associated with postmenopausal osteoporosis [64]. A different DNA-based study conducted an association analysis between selected polymorphisms in miRNA target sites (poly-miRTSs) and osteoporosis, identifying three poly-miRTSs in the FGF2 gene significantly associated with femoral neck BMD. These three single nucleotide polymorphisms (SNPs) in the 3′-UTR of the FGF2 gene reside within predicted binding sites for nine miRNAs (miR-146a, miR-146b, miR-545, miR-25, miR-32, miR-92, miR-363, miR-367 and miR-92b) and possibly alter the binding affinity between FGF2 transcripts and the identified miRNAs. Theoretically, an unfavorable allele would reduce the efficiency of miRNA binding to its target site, resulting in higher levels of target-protein expression. In the case of FGF2, such levels would stimulate osteoclastogenesis, enhance bone resorption and reduce BMD [107]. Even though this study lacks extensive functional confirmation, it raises an interesting point that encourages re-analysis of genome-wide association data, with the emphasis on miRNA target sites and pri-miRNA coding sequences.

Brought to you by | Tokyo Daigaku Authenticated Download Date | 6/7/15 3:27 AM

600      Vrtačnik et al.: Epigenetic mechanisms in bone

Figure 5 miRNAs possibly associated with low bone mineral density and osteoporosis. miR-705, miR-3077-5p and miR-214 were shown to suppress osteogenic differentiation by directly inhibiting their respective targets, thus, in all likelihood, contributing to bone loss associated with osteoporosis. Conversely, miR-2861 and miR-21 directly target negative regulators of osteoblastogenesis resulting in enhanced osteoblast differentiation and bone formation. The increased expression of miR-705 and miR-3077-5p and decreased levels of miR-21 were linked to high levels of TNFα and ROS, associated with estrogen deficiency. However, it was suggested that miR-503 and miRNAs targeting the FGF2 gene reduce osteoclastogenesis, suppress bone resorption and increase bone mass. Conversely, miR-148a and miR-133a promoted osteoclastogenesis and bone resorption by directly targeting negative regulators of osteoclast differentiation. Full lines represent established or experimentally proven associations while dotted lines represent hypothetical or in silico predicted connections.

Two studies have profiled the expression of miRNAs in peripheral blood mononuclear cells (PBMCs) from postmenopausal women with low and high BMD, while a third study focused on premenopausal women with systemic lupus erythematosus and low BMD. They found that miR-503 was markedly reduced and miR-133a significantly increased in osteoporosis patients [101, Chen et al., Unpublished manuscript]. Substantially raised levels of miR-148a were associated with lower BMD in lupus patients [102]. RANK was identified as a direct target of miR-503 and in vivo experiments in an ovariectomized mouse model showed that silencing of miR-503 increased RANK protein levels and osteoclastogenesis, promoted bone resorption

and decreased bone mass [Chen et al., Unpublished manuscript]. Both in vitro and in vivo experiments also confirmed that miR-148a promoted osteoclastogenesis and bone resorption by directly targeting MAFB, a negative regulator of NFATC1, C-FOS and MITF [102]. Functional testing of miR133a was not carried out, but in silico testing predicted that osteoclast-related target genes include CXCL11, CXCR3 and SLC39A1 (Figure 5) [101]. Since all three miRNAs are present in PBMCs, they represent easily accessible potential biomarkers associated with postmenopausal osteoporosis and increased bone resorption. Of these, miR-503 and miR-148a are especially promising as it has been experimentally proven that they play a significant role in regulating bone

Brought to you by | Tokyo Daigaku Authenticated Download Date | 6/7/15 3:27 AM

Vrtačnik et al.: Epigenetic mechanisms in bone      601

resorption and turnover. However, their status as putative therapeutic targets is more questionable, as it has been shown that the stimulation of miR-503 and inhibition of miR-148a expression suppressed both bone resorption and formation [102, Chen et al., Unpublished manuscript]. Bone is a highly vascularized tissue with a close relationship between osteogenesis and angiogenesis as well as bone remodeling and vascularization [108]. Furthermore, bone remodeling takes place within the bone-remodeling unit tightly linked to bone marrow capillaries; this is in order to provide necessary precursor cells and nutrients to this otherwise isolated environment [2]. Since bone tissue samples are rarely obtainable, the fact that miRNAs are present in serum, plasma, saliva, urine and other body fluids gives hope that miRNAs could also be used as potential non-invasive biomarkers for skeletal disorders [109].

 he role of DNA methylation in bone T remodeling All three types of epigenetic marks represent novel diagnostic and therapeutic opportunities; however, DNA methylation and DNMTs are currently showing the most promising results as adjuvant treatment for certain malignancies. Given that DNA methylation is also relevant in considerations of aging and age-related chronic diseases, future applications in osteoporosis and other bone-related pathologies are expected. An aberrant DNA methylation pattern that includes both global DNA hypomethylation and site-specific DNA hypermethylation seems to be one of the hallmarks of cancer. Several articles discussing DNA methylation in the context of clinical laboratory cancer management and neurodegenerative disorder research have been published recently in Clinical Chemistry and Laboratory Medicine [110–113].

region of the ALPL gene has been shown to be inversely associated with ALPL transcriptional levels. Accordingly, ALPL-expressing osteoblasts displayed the least methylated ALPL gene, while it was hypermethylated in osteocytes, which do not express ALPL. Bone lining cells, which represent a differentiation stage somewhere between osteoblasts and osteocytes, exhibited an intermediate methylation status; this suggests that ALPL methylation may increase progressively during osteoblastic differentiation, thus contributing notably to the phenotypic change associated with this process [117]. Interestingly, the transition from osteoblasts towards osteocytes was accompanied by a progressive decrease in levels of the SOST promoter methylation, enabling SOST to be expressed in an osteocyte-specific manner [118]. Genes shown to be coregulated by DNA methylation in osteoblasts also include transcription factors OSX and DLX5 [119], estrogen receptor α [120] and osteopontin [121]. Interestingly, mechanical stimulation, essential for maintaining bone homeostasis, was able to reduce DNA methylation at the osteopontin promoter, thus increasing its expression. This indicates that epigenetic mechanisms could represent the means through which mechanical signals are able to regulate target gene expression and cell differentiation [121]. Both the expression of RANKL in osteoblasts and the concomitant stimulation of osteoclastogenesis were also shown to be regulated, at least in part, by DNA methylation. Methylation of the RANKL gene promoter suppressed its transcriptional activity and responsiveness to vitamin D3 in osteoblasts, this probably being due to the binding of MECP2 to the methylated RANKL gene promoter, which, in turn, prevented occupancy by TBP [122, 123]. Interestingly, unlike in the cases of ALPL and SOST, the DNA methylation pattern of RANKL appeared to be established early in osteoblast precursors and did not undergo significant Table 3 Genes coregulated by DNA methylation in osteoblasts.

 NA methylation in osteoblast D differentiation Like posttranslational histone modifications and miRNAs, DNA methylation also takes part in regulating gene expression in bone tissue cells (Table 3). Several studies have shown that methylation of the gene encoding ALPL is a major controller of its expression, since DNA demethylating agents repeatedly increased ALPL expression and activity in osteoblasts under osteogenic and non-osteogenic conditions [114–117]. More specifically, the degree of methylation in the CpG island located in the proximal

Gene symbol  

Gene name



References

ALPL SOST OSX DLX5 ESR1

         

[114–117] [118] [119] [119] [120]

OPN RANKL

   

OPG SFRP1 LEP

     

Alkaline phosphatase   Sclerostin   Osterix   Distal-less homeobox 5   Estrogen receptor 1 (estrogen   receptor α) Osteopontin   Receptor activator of nuclear factor   κB ligand Osteoprotegerin   Secreted frizzled-related protein 1   Leptin  

Brought to you by | Tokyo Daigaku Authenticated Download Date | 6/7/15 3:27 AM

[121] [122–124] [124] [125] [125]

602      Vrtačnik et al.: Epigenetic mechanisms in bone changes during osteoblast differentiation. In addition to RANKL, a similar epigenetic regulation was also suggested for the expression of OPG. Even though DNA methylationbased regulation of RANKL and OPG expression was clearly shown at the cellular level, methylation analysis of RANKL and OPG genes in osteoporotic and osteoarthritic bone tissue samples revealed hypomethylated RANKL and OPG CpG islands without differences between the two pathologies, even though the expression of RANKL and the RANKL:OPG transcript ratio differed significantly. This could be due to the heterogeneity of bone tissue composed of cells of several different osteoblast and osteoclast differentiation stages compared to the homogeneity of individual cell lines. However, there is also the possibility that other mechanisms independent of DNA methylation are responsible for the increased RANKL:OPG ratio in osteo­ porotic patients [124]. In addition to analyses of individual genes, a genome-wide methylation profiling of bone tissue samples from patients with osteoporosis and osteoarthritis was also conducted, which revealed differences in the methylation of genes generally associated with the fate of less differentiated cells and with cell-matrix interactions that tend to be involved in skeletal development. The homeobox group of genes was particularly overrepresented among various differently methylated genes that, with the exceptions of SFRP1, a WNT-related inhibitor, and leptin, did not include the classic bone candidate genes discussed throughout this review. Even though this study has identified some differently methylated regions in osteoporosis and osteoarthritis and suggested a develop­ mental component associated with these disorders, it is necessary to keep in mind that DNA methylation is only one of the many factors influencing gene expression that can change during aging or under environmental stimuli. The involvement of aberrant DNA methylation in the pathogenesis of osteoporosis and osteoarthritis is therefore still inconclusive [125]. DNA methylation plays an undeniable regulatory role in osteoblasts that, through the control of RANKL gene expression, extends to osteoclasts and bone remodeling. Nevertheless, little is known about osteoclast-specific methylation patterns and about DNA-methylation upstream regulatory elements in general. Revealing the nature of the regulatory role DNA methylation plays in gene expression is difficult because methylation patterns are cell specific and because only some of the numerous CpGs present in the human genome take part in the process of gene transcription. Even though there are bioinformatic prediction tools available, it is still necessary to confirm experimentally the involvement of individual CpGs. Unlike in the case of SNPs, the analysis of single CpGs is not usually

enough, which necessitates the use of sequencing techniques. Similarly to that which is the case with miRNAs, DNA methylation biomarkers can be detected in several different non-invasive biological samples like serum, plasma, blood, PBMCs, saliva, sputum, oral rinses, bronchial washings and even exhaled breath. This is especially important in the context of metabolic bone diseases where tissue samples are seldom used. In addition, DNA is the most stable biological macromolecule and its methylation represents a covalent modification that is chemically inert and resistant to sample processing perturbations [111]. Therefore, DNA methylation biomarkers possess great potential that is yet to be fully explored and utilized. It is likely that many aberrantly methylated genes will never reach wider clinical significance as biomarkers in their own right; however, when assembled into sets of several genes there is a much greater probability of them becoming useful clinical tools. We have shown that DNA methylation is an important coregulator of bone homeostasis and it is reasonable to expect that deregulated methylation patterns could be associated with certain bone disorders. Unlike in oncology, where methylation tests have already reached clinical application, we are still waiting for a bone-associated DNA methylation biomarker. Analysis of posttranslational histone modifications still presents many challenges, ranging from the isolation of histones with intact posttranslational modifications to the simultaneous identification of numerous histones, histone variants and posttranslational histone modifications in a single sample (the latter only possible using mass spectroscopy-based proteomics). The analysis of miRNA expression and DNA methylation is perhaps less challenging, utilizing predominantly various PCR-, microarray- or sequencing-based methodologies, although their isolation from body fluids has its difficulties [67, 126].

Conclusions The precise control of gene expression is essential for proper development, differentiation, function and homeostasis. Epigenetic mechanisms are important regulators of gene expression and are capable of establishing stable, long-term expression patterns passed on during mitosis [20]. DNA methylation, miRNAs and posttranslational histone modifications thus represent additional layers of information-focusing and, in a way, help tailor genetic information to suit each cell in accordance with its location and purpose. It therefore makes sense to look at all the complementary layers of information, genetic and

Brought to you by | Tokyo Daigaku Authenticated Download Date | 6/7/15 3:27 AM

Vrtačnik et al.: Epigenetic mechanisms in bone      603

epigenetic, in order to understand both physiological and pathological processes at the molecular level. This is even more obvious when we look at how these layers interact with each other; e.g., SNPs in miRNA coding and target regions can change the efficiency of miRNA signaling, miRNAs can down-regulate the expression of many genes, including those coding for DNA-methylating and histonemodifying enzymes, while DNA methylation can suppress miRNA expression and so on. As epigenetic mechanisms are so essential to cell biology, they are associated with all cells of the body, all biological processes and, probably, all pathologies. This last is true even though there is usually no known causal relationship. Various environmental stimuli, such as physical and mental stresses, an imbalance of nutrients, infections of bacterial or viral origin, or contact with pollutants and chemical entities, can influence the organism through epigenetic changes [113]. Epigenetic mechanisms can also be targeted by different pharmacological agents, an approach successfully adopted in the treatment of certain cancers and neurological disorders. As mentioned earlier, broad-acting HDAC inhibitors showed great osteogenic potential in vitro, but in vivo observations have indicated negative skeletal consequences [34]. However, resveratrol, a SIRT1-activating plant polyphenol, promotes osteogenesis in vitro and prevents bone loss in vivo [49, 51, Zhao et al., Unpublished manuscript]. There are also a number of miRNAs that may represent appealing pharmacological targets. The in vivo inhibition of bone resorption by overexpressing miR-503 with a specific agomir is one such example [Chen et al., Unpublished manuscript]. Hypomethylating agents are also available, although their skeletal effects are yet to be well characterized and their potential utility is limited by their widespread mechanism of action. Manipulation of epigenetic mechanisms holds great promise for the

future treatment of metabolic bone diseases but a better understanding of epigenetic regulation and increased specificity of pharmacological agents are needed first. The presence of epigenetic marks in various body fluids also makes them very attractive biomarkers, especially in pathologies associated with tissues and organs normally impossible to sample. This is also the case in bone disorders, where imaging techniques still represent the main diagnostic tools, while biochemical markers have failed to reach wider significance and tissue biopsies are only seldom obtained. Epigenetic marks represent a new generation of potential biochemical markers that will hopefully enable patients at risk of bone loss to be screened before proceeding with imaging techniques, be able to identify small changes in the progress of a given disease and response to therapy beyond the capabilities of current methods, and will facilitate the personalized therapy of skeletal diseases. Conflict of interest statement Authors’ conflict of interest disclosure: The authors stated that there are no conflicts of interest regarding the publication of this article. Research support played no role in the study design; in the collection, analysis, and interpretation of data; in the writing of the report; or in the decision to submit the report for publication. Research funding: This work was supported by the Slovenian Research Agency (Grants ARMR19, P3-298 and J3-551). Employment or leadership: None declared. Honorarium: None declared. Received September 13, 2013; accepted November 19, 2013; previously published online December 18, 2013

References 1. Feng X, McDonald JM. Disorders of bone remodeling. Annu Rev Pathol 2011;6:121–45. 2. Andersen TL, Sondergaard TE, Skorzynska KE, Dagnaes-Hansen F, Plesner TL, Hauge EM, et al. A physical mechanism for coupling bone resorption and formation in adult human bone. Am J Pathol 2009;174:239–47. 3. Raggatt LJ, Partridge NC. Cellular and molecular mechanisms of bone remodeling. J Biol Chem 2010;285:25103–8. 4. Long F. Building strong bones: molecular regulation of the osteoblast lineage. Nat Rev Mol Cell Biol 2012;13:27–38. 5. Lin GL, Hankenson KD. Integration of BMP, Wnt, and notch signaling pathways in osteoblast differentiation. J Cell Biochem 2011;112:3491–501.

6. Krause C, de Gorter DJ, Karperien M, ten Dijke P. Chapter 2. Signal transduction cascades controlling osteoblast differentiation. In: Rosen CJ, editor. Primer on the metabolic bone diseases and disorders of mineral metabolism, 7th ed. Washington: ASBMR, 2009:10–6. 7. Edwards JR, Mundy GR. Advances in osteoclast biology: old findings and new insights from mouse models. Nat Rev Rheumatol 2011;7:235–43. 8. Abdallah BM, Kassem M. Human mesenchymal stem cells: from basic biology to clinical applications. Gene Ther 2008;15: 109–16. 9. Jang YN, Baik EJ. JAK-STAT pathway and myogenic differentiation. JAKSTAT 2013;2:e23282.

Brought to you by | Tokyo Daigaku Authenticated Download Date | 6/7/15 3:27 AM

604      Vrtačnik et al.: Epigenetic mechanisms in bone 10. Abramson SB, Attur M. Developments in the scientific understanding of osteoarthritis. Arthritis Res Ther 2009;11:227. 11. Kearns AE, Khosla S, Kostenuik PJ. Receptor activator of nuclear factor kappaB ligand and osteoprotegerin regulation of bone remodeling in health and disease. Endocr Rev 2008;29:155–92. 12. Zupan J, Jeras M, Marc J. Osteoimmunology and the influence of pro-inflammatory cytokines on osteoclasts. Biochem Med (Zagreb) 2013;23:43–63. 13. Lacey DL, Boyle WJ, Simonet WS, Kostenuik PJ, Dougall WC, Sullivan JK, et al. Bench to bedside: elucidation of the OPG-RANK-RANKL pathway and the development of denosumab. Nat Rev Drug Discov 2012;11:401–19. 14. Monroe DG, McGee-Lawrence ME, Oursler MJ, Westendorf JJ. Update on Wnt signaling in bone cell biology and bone disease. Gene 2012;492:1–18. 15. Li X, Ominsky MS, Warmington KS, Morony S, Gong J, Cao J, et al. Sclerostin antibody treatment increases bone formation, bone mass, and bone strength in a rat model of postmeno­pausal osteoporosis. J Bone Miner Res 2009;24:578–88. 16. Padhi D, Jang G, Stouch B, Fang L, Posvar E. Single-dose, placebo-controlled, randomized study of AMG 785, a sclerostin monoclonal antibody. J Bone Miner Res 2011;26:19–26. 17. Centrella M, McCarthy TL. Estrogen receptor dependent gene expression by osteoblasts – direct, indirect, circumspect, and speculative effects. Steroids 2012;77:174–84. 18. Burgers TA, Williams BO. Regulation of Wnt/β-catenin signaling within and from osteocytes. Bone 2013;54:244–9. 19. Diarra D, Stolina M, Polzer K, Zwerina J, Ominsky MS, Dwyer D, et al. Dickkopf-1 is a master regulator of joint remodeling. Nat Med 2007;13:156–63. 20. Gibney ER, Nolan CM. Epigenetics and gene expression. Heredity (Edinb) 2010;105:4–13. 21. Portela A, Esteller M. Epigenetic modifications and human disease. Nat Biotechnol 2010;28:1057–68. 22. Peserico A, Simone C. Physical and functional HAT/HDAC interplay regulates protein acetylation balance. J Biomed Biotechnol 2011;2011:371832. 23. Kapinas K, Delany AM. MicroRNA biogenesis and regulation of bone remodeling. Arthritis Res Ther 2011;13:220. 24. Vasudevan S, Tong Y, Steitz JA. Switching from repression to activation: microRNAs can up-regulate translation. Science 2007;318:1931–4. 25. Khan N, Jeffers M, Kumar S, Hackett C, Boldog F, Khramtsov N, et al. Determination of the class and isoform selectivity of small-molecule histone deacetylase inhibitors. Biochem J 2008;409:581–9. 26. Balasubramanian S, Verner E, Buggy JJ. Isoform-specific histone deacetylase inhibitors: the next step? Cancer Lett 2009;280:211–21. 27. Schroeder TM, Nair AK, Staggs R, Lamblin AF, Westendorf JJ. Gene profile analysis of osteoblast genes differentially regulated by histone deacetylase inhibitors. BMC Genomics 2007;8:362. 28. Schroeder TM, Westendorf JJ. Histone deacetylase inhibitors promote osteoblast maturation. J Bone Miner Res 2005;20:2254–63. 29. Cho HH, Park HT, Kim YJ, Bae YC, Suh KT, Jung JS. Induction of osteogenic differentiation of human mesenchymal stem

cells by histone deacetylase inhibitors. J Cell Biochem 2005;96:533–42. 30. Fan X, Roy EM, Murphy TC, Nanes MS, Kim S, Pike JW, et al. Regulation of RANKL promoter activity is associated with histone remodeling in murine bone stromal cells. J Cell Biochem 2004;93:807–18. 31. Kim HN, Lee JH, Bae SC, Ryoo HM, Kim HH, Ha H, et al. Histone deacetylase inhibitor MS-275 stimulates bone formation in part by enhancing Dhx36-mediated TNAP transcription. J Bone Miner Res 2011;26:2161–73. 32. McGee-Lawrence ME, McCleary-Wheeler AL, Secreto FJ, Razidlo DF, Zhang M, Stensgard BA, et al. Suberoylanilide hydroxamic acid (SAHA; vorinostat) causes bone loss by inhibiting immature osteoblasts. Bone 2011;48:1117–26. 33. Senn SM, Kantor S, Poulton IJ, Morris MJ, Sims NA, O′Brien TJ, et al. Adverse effects of valproate on bone: defining a model to investigate the pathophysiology. Epilepsia 2010;51:984–93. 34. McGee-Lawrence ME, Westendorf JJ. Histone deacetylases in skeletal development and bone mass maintenance. Gene 2011;474:1–11. 35. Shen J, Montecino M, Lian JB, Stein GS, Van Wijnen AJ, Stein JL. Histone acetylation in vivo at the osteocalcin locus is functionally linked to vitamin D-dependent, bone tissue-specific transcription. J Biol Chem 2002;277:20284–92. 36. Montecino M, Frenkel B, van Wijnen AJ, Lian JB, Stein GS, Stein JL. Chromatin hyperacetylation abrogates vitamin D-mediated transcriptional upregulation of the tissue-specific osteocalcin gene in vivo. Biochemistry 1999;38:1338–45. 37. Sierra J, Villagra A, Paredes R, Cruzat F, Gutierrez S, Javed A, et al. Regulation of the bone-specific osteocalcin gene by p300 requires Runx2/Cbfa1 and the vitamin D3 receptor but not p300 intrinsic histone acetyltransferase activity. Mol Cell Biol 2003;23:3339–51. 38. Shen J, Hovhannisyan H, Lian JB, Montecino MA, Stein GS, Stein JL, et al. Transcriptional induction of the osteocalcin gene during osteoblast differentiation involves acetylation of histones h3 and h4. Mol Endocrinol 2003;17:743–56. 39. Choo MK, Yeo H, Zayzafoon M. NFATc1 mediates HDAC-dependent transcriptional repression of osteocalcin expression during osteoblast differentiation. Bone 2009;45:579–89. 40. Schroeder TM, Kahler RA, Li X, Westendorf JJ. Histone deacetylase 3 interacts with runx2 to repress the osteocalcin promoter and regulate osteoblast differentiation. J Biol Chem 2004;279:41998–2007. 41. Hesse E, Saito H, Kiviranta R, Correa D, Yamana K, Neff L, et al. Zfp521 controls bone mass by HDAC3-dependent attenuation of Runx2 activity. J Cell Biol 2010;191:1271–83. 42. Lamour V, Detry C, Sanchez C, Henrotin Y, Castronovo V, Bellahcène A. Runx2- and histone deacetylase 3-mediated repression is relieved in differentiating human osteoblast cells to allow high bone sialoprotein expression. J Biol Chem 2007;282:36240–9. 43. Razidlo DF, Whitney TJ, Casper ME, McGee-Lawrence ME, Stensgard BA, Li X, et al. Histone deacetylase 3 depletion in osteo/chondroprogenitor cells decreases bone density and increases marrow fat. PLoS One 2010;5:e11492. 44. McGee-Lawrence ME, Bradley EW, Dudakovic A, Carlson SW, Ryan ZC, Kumar R, et al. Histone deacetylase 3 is required for maintenance of bone mass during aging. Bone 2013;52: 296–307.

Brought to you by | Tokyo Daigaku Authenticated Download Date | 6/7/15 3:27 AM

Vrtačnik et al.: Epigenetic mechanisms in bone      605 45. Kang JS, Alliston T, Delston R, Derynck R. Repression of Runx2 function by TGF-beta through recruitment of class II histone deacetylases by Smad3. EMBO J 2005;24:2543–55. 46. Jeon EJ, Lee KY, Choi NS, Lee MH, Kim HN, Jin YH, et al. Bone morphogenetic protein-2 stimulates Runx2 acetylation. J Biol Chem 2006;281:16502–11. 47. Lee HW, Suh JH, Kim AY, Lee YS, Park SY, Kim JB. Histone deacetylase 1-mediated histone modification regulates osteoblast differentiation. Mol Endocrinol 2006;20:2432–43. 48. Cohen-Kfir E, Artsi H, Levin A, Abramowitz E, Bajayo A, Gurt I, et al. Sirt1 is a regulator of bone mass and a repressor of Sost encoding for sclerostin, a bone formation inhibitor. Endocrinology 2011;152:4514–24. 49. Bäckesjö CM, Li Y, Lindgren U, Haldosén LA. Activation of Sirt1 decreases adipocyte formation during osteoblast differentiation of mesenchymal stem cells. J Bone Miner Res 2006;21:993–1002. 50. Elbaz A, Rivas D, Duque G. Effect of estrogens on bone marrow adipogenesis and Sirt1 in aging C57BL/6J mice. Biogerontology 2009;10:747–55. 51. Tseng PC, Hou SM, Chen RJ, Peng HW, Hsieh CF, Kuo ML, et al. Resveratrol promotes osteogenesis of human mesenchymal stem cells by upregulating RUNX2 gene expression via the SIRT1/FOXO3A axis. J Bone Miner Res 2011;26:2552–63. 52. Fusco S, Maulucci G, Pani G. Sirt1: def-eating senescence? Cell Cycle 2012;11:4135–46. 53. Shakibaei M, Buhrmann C, Mobasheri A. Resveratrol-mediated SIRT-1 interactions with p300 modulate receptor activator of NF-kappaB ligand (RANKL) activation of NF-kappaB signaling and inhibit osteoclastogenesis in bone-derived cells. J Biol Chem 2011;286:11492–505. 54. Edwards JR, Perrien DS, Fleming N, Nyman JS, Ono K, Connelly L, et al. Silent information regulator (Sir)T1 inhibits NF-κB signaling to maintain normal skeletal remodeling. J Bone Miner Res 2013;28:960–9. 55. Huang W, Shang WL, Wang HD, Wu WW, Hou SX. Sirt1 overexpression protects murine osteoblasts against TNF-α-induced injury in vitro by suppressing the NF-κB signaling pathway. Acta Pharmacol Sin 2012;33:668–74. 56. Rahman MM, Kukita A, Kukita T, Shobuike T, Nakamura T, Kohashi O. Two histone deacetylase inhibitors, trichostatin A and sodium butyrate, suppress differentiation into osteoclasts but not into macrophages. Blood 2003;101:3451–9. 57. Nakamura T, Kukita T, Shobuike T, Nagata K, Wu Z, Ogawa K, et al. Inhibition of histone deacetylase suppresses osteoclastogenesis and bone destruction by inducing IFN-beta production. J Immunol 2005;175:5809–16. 58. Takada Y, Gillenwater A, Ichikawa H, Aggarwal BB. Suberoylanilide hydroxamic acid potentiates apoptosis, inhibits invasion, and abolishes osteoclastogenesis by suppressing nuclear factor-kappaB activation. J Biol Chem 2006;281: 5612–22. 59. Kim HN, Lee JH, Jin WJ, Ko S, Jung K, Ha H, et al. MS-275, a benzamide histone deacetylase inhibitor, prevents osteoclastogenesis by down-regulating c-Fos expression and suppresses bone loss in mice. Eur J Pharmacol 2012;691:69–76. 60. Cantley MD, Fairlie DP, Bartold PM, Rainsford KD, Le GT, Lucke AJ, et al. Inhibitors of histone deacetylases in class I and class II suppress human osteoclasts in vitro. J Cell Physiol 2011;226:3233–41.

61. Pham L, Kaiser B, Romsa A, Schwarz T, Gopalakrishnan R, Jensen ED, et al. HDAC3 and HDAC7 have opposite effects on osteoclast differentiation. J Biol Chem 2011;286:12056–65. 62. Kim JH, Kim K, Youn BU, Jin HM, Kim JY, Moon JB, et al. RANKL induces NFATc1 acetylation and stability via histone acetyltransferases during osteoclast differentiation. Biochem J 2011;436:253–62. 63. Rivadeneira F, Styrkársdottir U, Estrada K, Halldórsson BV, Hsu YH, Richards JB, et al. Twenty bone-mineral-density loci identified by large-scale meta-analysis of genome-wide association studies. Nat Genet 2009;41:1199–206. 64. Li H, Xie H, Liu W, Hu R, Huang B, Tan YF, et al. A novel microRNA targeting HDAC5 regulates osteoblast differentiation in mice and contributes to primary osteoporosis in humans. J Clin Invest 2009;119:3666–77. 65. Cortez MA, Bueso-Ramos C, Ferdin J, Lopez-Berestein G, Sood AK, Calin GA. MicroRNAs in body fluids – the mix of hormones and biomarkers. Nat Rev Clin Oncol 2011;8: 467–77. 66. Sun Y, Zhang K, Fan G, Li J. Identification of circulating microRNAs as biomarkers in cancers: what have we got? Clin Chem Lab Med 2012;50:2121–6. 67. Sandoval J, Peiró-Chova L, Pallardó FV, García-Giménez JL. Epigenetic biomarkers in laboratory diagnostics: emerging approaches and opportunities. Expert Rev Mol Diagn 2013;13:457–71. 68. Zhang Y, Xie RL, Croce CM, Stein JL, Lian JB, van Wijnen AJ, et al. A program of microRNAs controls osteogenic lineage progression by targeting transcription factor Runx2. Proc Natl Acad Sci USA 2011;108:9863–8. 69. Li Z, Hassan MQ, Volinia S, van Wijnen AJ, Stein JL, Croce CM, et al. A microRNA signature for a BMP2-induced osteoblast lineage commitment program. Proc Natl Acad Sci USA 2008;105:13906–11. 70. Zhang Y, Xie RL, Gordon J, LeBlanc K, Stein JL, Lian JB, et al. Control of mesenchymal lineage progression by microRNAs targeting skeletal gene regulators Trps1 and Runx2. J Biol Chem 2012;287:21926–35. 71. Huang J, Zhao L, Xing L, Chen D. MicroRNA-204 regulates Runx2 protein expression and mesenchymal progenitor cell differentiation. Stem Cells 2010;28:357–64. 72. Wu T, Zhou H, Hong Y, Li J, Jiang X, Huang H. miR-30 family members negatively regulate osteoblast differentiation. J Biol Chem 2012;287:7503–11. 73. Kim EJ, Kang IH, Lee JW, Jang WG, Koh JT. MiR-433 mediates ERRγ-suppressed osteoblast differentiation via direct targeting to Runx2 mRNA in C3H10T1/2 cells. Life Sci 2013;92:562–8. 74. Hassan MQ, Gordon JA, Beloti MM, Croce CM, van Wijnen AJ, Stein JL, et al. A network connecting Runx2, SATB2, and the miR-23a~27a~24-2 cluster regulates the osteoblast d ­ ifferentiation program. Proc Natl Acad Sci USA 2010;107:19879–84. 75. Tomé M, López-Romero P, Albo C, Sepúlveda JC, FernándezGutiérrez B, Dopazo A, et al. miR-335 orchestrates cell proliferation, migration and differentiation in human mesenchymal stem cells. Cell Death Differ 2011;18: 985–95. 76. Liao L, Yang X, Su X, Hu C, Zhu X, Yang N, et al. Redundant miR-3077-5p and miR-705 mediate the shift of mesenchymal

Brought to you by | Tokyo Daigaku Authenticated Download Date | 6/7/15 3:27 AM

606      Vrtačnik et al.: Epigenetic mechanisms in bone stem cell lineage commitment to adipocyte in osteoporosis bone marrow. Cell Death Dis 2013;4:e600. 77. Hu R, Liu W, Li H, Yang L, Chen C, Xia ZY, et al. A Runx2/ miR-3960/miR-2861 regulatory feedback loop during mouse osteoblast differentiation. J Biol Chem 2011;286:12328–39. 78. Yang L, Cheng P, Chen C, He HB, Xie GQ, Zhou HD, et al. miR-93/ Sp7 function loop mediates osteoblast mineralization. J Bone Miner Res 2012;27:1598–606. 79. Kapinas K, Kessler C, Ricks T, Gronowicz G, Delany AM. miR-29 modulates Wnt signaling in human osteoblasts through a positive feedback loop. J Biol Chem 2010;285:25221–31. 80. Hassan MQ, Maeda Y, Taipaleenmaki H, Zhang W, Jafferji M, Gordon JA, et al. miR-218 directs a Wnt signaling circuit to promote differentiation of osteoblasts and osteomimicry of metastatic cancer cells. J Biol Chem 2012;287:42084–92. 81. Zhang J, Tu Q, Bonewald LF, He X, Stein G, Lian J, et al. Effects of miR-335-5p in modulating osteogenic differentiation by specifically downregulating Wnt antagonist DKK1. J Bone Miner Res 2011;26:1953–63. 82. Wang T, Xu Z. miR-27 promotes osteoblast differentiation by modulating Wnt signaling. Biochem Biophys Res Commun 2010;402:186–9. 83. Hu W, Ye Y, Zhang W, Wang J, Chen A, Guo F. miR-142-3p promotes osteoblast differentiation by modulating Wnt signaling. Mol Med Rep 2013;7:689–93. 84. Li Z, Hassan MQ, Jafferji M, Aqeilan RI, Garzon R, Croce CM, et al. Biological functions of miR-29b contribute to positive regulation of osteoblast differentiation. J Biol Chem 2009;284: 15676–84. 85. Zhang JF, Fu WM, He ML, Xie WD, Lv Q, Wan G, et al. MiRNA-20a promotes osteogenic differentiation of human mesenchymal stem cells by co-regulating BMP signaling. RNA Biol 2011;8:829–38. 86. Yang N, Wang G, Hu C, Shi Y, Liao L, Shi S, et al. Tumor necrosis factor α suppresses the mesenchymal stem cell osteogenesis promoter miR-21 in estrogen deficiency-induced osteoporosis. J Bone Miner Res 2013;28:559–73. 87. Kapinas K, Kessler CB, Delany AM. miR-29 suppression of osteonectin in osteoblasts: regulation during differentiation and by canonical Wnt signaling. J Cell Biochem 2009;108: 216–24. 88. Mizuno Y, Tokuzawa Y, Ninomiya Y, Yagi K, Yatsuka-Kanesaki Y, Suda T, et al. miR-210 promotes osteoblastic differentiation through inhibition of AcvR1b. FEBS Lett 2009;583:2263–8. 89. Guo J, Ren F, Wang Y, Li S, Gao Z, Wang X, et al. miR-764-5p promotes osteoblast differentiation through inhibition of CHIP/ STUB1 expression. J Bone Miner Res 2012;27:1607–18. 90. Eskildsen T, Taipaleenmäki H, Stenvang J, Abdallah BM, Ditzel N, Nossent AY, et al. MicroRNA-138 regulates osteogenic differentiation of human stromal (mesenchymal) stem cells in vivo. Proc Natl Acad Sci USA 2011;108:6139–44. 91. Itoh T, Nozawa Y, Akao Y. MicroRNA-141 and -200a are involved in bone morphogenetic protein-2-induced mouse pre-osteoblast differentiation by targeting distal-less homeobox 5. J Biol Chem 2009;284:19272–9. 92. Wu T, Xie M, Wang X, Jiang X, Li J, Huang H. miR-155 modulates TNF-α-inhibited osteogenic differentiation by targeting SOCS1 expression. Bone 2012;51:498–505. 93. Kim KM, Park SJ, Jung SH, Kim EJ, Jogeswar G, Ajita J, et al. miR-182 is a negative regulator of osteoblast proliferation,



differentiation, and skeletogenesis through targeting FoxO1. J Bone Miner Res 2012;27:1669–79. 94. Inose H, Ochi H, Kimura A, Fujita K, Xu R, Sato S, et al. A microRNA regulatory mechanism of osteoblast differentiation. Proc Natl Acad Sci USA 2009;106:20794–9. 95. Itoh T, Takeda S, Akao Y. MicroRNA-208 modulates BMP-2-stimulated mouse preosteoblast differentiation by directly targeting V-ets erythroblastosis virus E26 oncogene homolog 1. J Biol Chem 2010;285:27745–52. 96. Wang X, Guo B, Li Q, Peng J, Yang Z, Wang A, et al. miR-214 targets ATF4 to inhibit bone formation. Nat Med 2013;19: 93–100. 97. Itoh T, Ando M, Tsukamasa Y, Akao Y. Expression of BMP-2 and Ets1 in BMP-2-stimulated mouse pre-osteoblast differentiation is regulated by microRNA-370. FEBS Lett 2012;586:1693–701. 98. Kahai S, Lee SC, Lee DY, Yang J, Li M, Wang CH, et al. MicroRNA miR-378 regulates nephronectin expression modulating osteoblast differentiation by targeting GalNT-7. PLoS One 2009;4:e7535. 99. Sugatani T, Vacher J, Hruska KA. A microRNA expression signature of osteoclastogenesis. Blood 2011;117:3648–57. 100. Sugatani T, Hruska KA. Down-regulation of miR-21 biogenesis by estrogen action contributes to osteoclastic apoptosis. J Cell Biochem 2013;114:1217–22. 101. Wang Y, Li L, Moore BT, Peng XH, Fang X, Lappe JM, et al. MiR-133a in human circulating monocytes: a potential biomarker associated with postmenopausal osteoporosis. PLoS One 2012;7:e34641. 102. Cheng P, Chen C, He HB, Hu R, Zhou HD, Xie H, et al. miR-148a regulates osteoclastogenesis by targeting V-maf musculoaponeurotic fibrosarcoma oncogene homolog B. J Bone Miner Res 2013;28:1180–90. 103. Sugatani T, Hruska KA. Impaired micro-RNA pathways diminish osteoclast differentiation and function. J Biol Chem 2009;284:4667–78. 104. Mann M, Barad O, Agami R, Geiger B, Hornstein E. miRNA-based mechanism for the commitment of multipotent progenitors to a single cellular fate. Proc Natl Acad Sci USA 2010;107:15804–9. 105. Sugatani T, Hruska KA. MicroRNA-223 is a key factor in osteoclast differentiation. J Cell Biochem 2007;101:996–9. 106. Jones SW, Watkins G, Le Good N, Roberts S, Murphy CL, Brockbank SM, et al. The identification of differentially expressed microRNA in osteoarthritic tissue that modulate the production of TNF-alpha and MMP13. Osteoarthritis Cartilage 2009;17:464–72. 107. Lei SF, Papasian CJ, Deng HW. Polymorphisms in predicted miRNA binding sites and osteoporosis. J Bone Miner Res 2011;26:72–8. 108. Schneider P, Krucker T, Meyer E, Ulmann-Schuler A, Weber B, Stampanoni M, et al. Simultaneous 3D visualization and quantification of murine bone and bone vasculature using micro-computed tomography and vascular replica. Microsc Res Tech 2009;72:690–701. 109. Weber JA, Baxter DH, Zhang S, Huang DY, Huang KH, Lee MJ, et al. The microRNA spectrum in 12 body fluids. Clin Chem 2010;56:1733–41. 110. Heichman KA, Warren JD. DNA methylation biomarkers and their utility for solid cancer diagnostics. Clin Chem Lab Med 2012;50:1707–21.

Brought to you by | Tokyo Daigaku Authenticated Download Date | 6/7/15 3:27 AM

Vrtačnik et al.: Epigenetic mechanisms in bone      607 111. Markopoulou S, Nikolaidis G, Liloglou T. DNA methylation biomarkers in biological fluids for early detection of respiratory tract cancer. Clin Chem Lab Med 2012;50: 1723–31. 112. Toraño EG, Petrus S, Fernandez AF, Fraga MF. Global DNA hypomethylation in cancer: review of validated methods and clinical significance. Clin Chem Lab Med 2012;50:1733–42. 113. Fuso A. The ‘golden age’ of DNA methylation in neurodegenerative diseases. Clin Chem Lab Med 2013;51:523–34. 114. Locklin RM, Oreffo RO, Triffitt JT. Modulation of osteogenic differentiation in human skeletal cells in Vitro by 5-azacytidine. Cell Biol Int 1998;22:207–15. 115. Vaes BL, Lute C, van der Woning SP, Piek E, Vermeer J, Blom HJ, et al. Inhibition of methylation decreases osteoblast differentiation via a non-DNA-dependent methylation mechanism. Bone 2010;46:514–23. 116. El-Serafi AT, Oreffo RO, Roach HI. Epigenetic modifiers influence lineage commitment of human bone marrow stromal cells: differential effects of 5-aza-deoxycytidine and trichostatin A. Differentiation 2011;81:35–41. 117. Delgado-Calle J, Sañudo C, Sánchez-Verde L, García-Renedo RJ, Arozamena J, Riancho JA. Epigenetic regulation of alkaline phosphatase in human cells of the osteoblastic lineage. Bone 2011;49:830–8. 118. Delgado-Calle J, Sañudo C, Bolado A, Fernández AF, Arozamena J, Pascual-Carra MA, et al. DNA methylation contributes to the regulation of sclerostin expression in human osteocytes. J Bone Miner Res 2012;27:926–37. 119. Lee JY, Lee YM, Kim MJ, Choi JY, Park EK, Kim SY, et al. Methylation of the mouse DIx5 and Osx gene promoters

Peter Vrtačnik is currently a young researcher and a PhD student of Biomedicine, the study program of Clinical Biochemistry and Laboratory Biomedicine, at University of Ljubljana, Faculty of Pharmacy. He received his master’s in Pharmacy from University of Ljubljana, Faculty of Pharmacy in 2009. His work focuses on the interaction between different epigenetic mechanisms and known stressors of bone homeostasis at cellular and tissue levels.

regulates cell type-specific gene expression. Mol Cells 2006;22:182–8. 120. Penolazzi L, Lambertini E, Giordano S, Sollazzo V, Traina G, del Senno L, et al. Methylation analysis of the promoter F of estrogen receptor alpha gene: effects on the level of transcription on human osteoblastic cells. J Steroid Biochem Mol Biol 2004;91:1–9. 121. Arnsdorf EJ, Tummala P, Castillo AB, Zhang F, Jacobs CR. The epigenetic mechanism of mechanically induced osteogenic differentiation. J Biomech 2010;43:2881–6. 122. Kitazawa S, Kitazawa R. Epigenetic control of mouse receptor activator of NF-kappa B ligand gene expression. Biochem Biophys Res Commun 2002;293:126–31. 123. Kitazawa R, Kitazawa S. Methylation status of a single CpG locus 3 bases upstream of TATA-box of receptor activator of nuclear factor-kappaB ligand (RANKL) gene promoter modulates cell- and tissue-specific RANKL expression and osteoclastogenesis. Mol Endocrinol 2007;21:148–58. 124. Delgado-Calle J, Sañudo C, Fernández AF, García-Renedo R, Fraga MF, Riancho JA. Role of DNA methylation in the regulation of the RANKL-OPG system in human bone. Epigenetics 2012;7:83–91. 125. Delgado-Calle J, Fernández AF, Sainz J, Zarrabeitia MT, Sañudo C, García-Renedo R, et al. Genome-wide profiling of bone reveals differentially methylated regions in osteoporosis and osteoarthritis. Arthritis Rheum 2013;65:197–205. 126. García-Giménez JL, Sanchis-Gomar F, Lippi G, Mena S, Ivars D, Gomez-Cabrera MC, et al. Epigenetic biomarkers: a new perspective in laboratory diagnostics. Clin Chim Acta 2012;413:1576–82.

Janja Marc, EurClinChem is a full Professor of Clinical Biochemistry and the head of Laboratory for Molecular Diagnostics at the University of Ljubljana, Faculty of Pharmacy. Her research is focused on genetic and epigenetic background of complex diseases like osteoporosis, osteoarthritis, insulin resistance, obesity and lung cancer. Using the genotype-phenotype association studies and genome-wide expression profiling of bone cells her group succeeded to identify some of important molecules in these diseases. She is also the co-author of over 80 scientific papers with nearly 700 citations. She is a member of the Laboratory Diagnostics Advisory Board at Ministry of Health RS. Since 2012 she is also serving as the General Secretary of European Society of Pharmacogenomics and Theranostics

Brought to you by | Tokyo Daigaku Authenticated Download Date | 6/7/15 3:27 AM

608      Vrtačnik et al.: Epigenetic mechanisms in bone

Barbara Ostanek is an Assistant Professor of Clinical Biochemistry at the University of Ljubljana, Faculty of Pharmacy and works also in the Laboratory of Molecular Diagnostics at the Faculty of Pharmacy. Her research is focused on molecular basis of diseases, identification of new diagnostic markers and new drug targets. So far, this included the identification of relevant DNA polymorphisms and changes in gene expression, mainly in osteoporosis, hyperbilirubinemia and diabetes. She is now complementing these approaches with the study of epigenetic changes. Another topic is also the development and optimization of new genotyping tests. She is interested in pharmacogenomics of antiosteoporotic, anticoagulant and anticancer drugs.

Brought to you by | Tokyo Daigaku Authenticated Download Date | 6/7/15 3:27 AM

Epigenetic mechanisms in bone.

Epigenetics refers to the study of mechanisms able to influence gene expression in a stable and potentially heritable manner without altering the DNA ...
2MB Sizes 0 Downloads 0 Views