AI Chat Paper
Note: Please note that the following content is generated by AMiner AI. SciOpen does not take any responsibility related to this content.
{{lang === 'zh_CN' ? '文章概述' : 'Summary'}}
{{lang === 'en_US' ? '中' : 'Eng'}}
Chat more with AI
PDF (2.4 MB)
Collect
Submit Manuscript AI Chat Paper
Show Outline
Outline
Show full outline
Hide outline
Outline
Show full outline
Hide outline
Review | Open Access

Bone remodeling: an operational process ensuring survival and bone mechanical competence

Simona Bolamperti1Isabella Villa1Alessandro Rubinacci1 ( )
Osteoporosis and Bone and Mineral Metabolism Unit, IRCCS San Raffaele Hospital, Via Olgettina 60, 20132 Milano, Italy
Show Author Information

Abstract

Bone remodeling replaces old and damaged bone with new bone through a sequence of cellular events occurring on the same surface without any change in bone shape. It was initially thought that the basic multicellular unit (BMU) responsible for bone remodeling consists of osteoclasts and osteoblasts functioning through a hierarchical sequence of events organized into distinct stages. However, recent discoveries have indicated that all bone cells participate in BMU formation by interacting both simultaneously and at different differentiation stages with their progenitors, other cells, and bone matrix constituents. Therefore, bone remodeling is currently considered a physiological outcome of continuous cellular operational processes optimized to confer a survival advantage. Bone remodeling defines the primary activities that BMUs need to perform to renew successfully bone structural units. Hence, this review summarizes the current understanding of bone remodeling and future research directions with the aim of providing a clinically relevant biological background with which to identify targets for therapeutic strategies in osteoporosis.

References

1
Frost, H. M. Bone remodeling dynamics (Charles C Thomas Company, 1963).
2

Delaisse, J. M. et al. Re-thinking the bone remodeling cycle mechanism and the origin of bone loss. Bone 141, 115628 (2020).

3

Maggiano, C. M., Maggiano, I. S., Tiesler, V. G., Chi-Keb, J. R. & Stout, S. D. Methods and theory in bone modeling drift: comparing spatial analyses of primary bone distributions in the human humerus. J. Anat. 228, 190–202 (2016).

4

Parfitt, A. M. Osteonal and hemi-osteonal remodeling: the spatial and temporal framework for signal traffic in adult human bone. J. Cell. Biochem. 55, 273–286 (1994).

5

Frost, H. M. A synchronous group of mammalian cells whose in vivo behavior can be studied. Henry Ford. Hosp. Med. Bull. 13, 161–172 (1965).

6
Feynman, R., Leighton, R. & Sands, M. The Feynman lectures on physics (Addison-Wesley Pub. Co., 1964).
7

Parfitt, A. M. The cellular basis of bone remodeling: the quantum concept reexamined in light of recent advances in the cell biology of bone. Calcif. Tissue Int. 36, S37–S45 (1984).

8

Eriksen, E. F. Normal and pathological remodeling of human trabecular bone: three dimensional reconstruction of the remodeling sequence in normals and in metabolic bone disease. Endocr. Rev. 7, 379–408 (1986).

9

Parfitt, A. M. The quantum concept of bone remodeling and turnover: implications for the pathogenesis of osteoporosis. Calcif. Tissue Int. 28, 1–5 (1979).

10

Marais, A. et al. The future of quantum biology. J. R. Soc. Interface 15, 20180640 (2018).

11

Manolagas, S. C. Birth and death of bone cells: basic regulatory mechanisms and implications for the pathogenesis and treatment of osteoporosis. Endocr. Rev. 21, 115–137 (2000).

12

Jilka, R. L. Biology of the basic multicellular unit and the pathophysiology of osteoporosis. Med. Pediatr. Oncol. 41, 182–185 (2003).

13

Parfitt, A. M. Misconceptions (2): turnover is always higher in cancellous than in cortical bone. Bone 30, 807–809 (2002).

14
Bonewald, L. F. & Marcus, D. F. R. In Osteoporosis 3rd edn (eds Nelson, D. & Rosen, C.) 170–189 (Elsevier, 2008).
15

Metz, L. N., Martin, R. B. & Turner, A. S. Histomorphometric analysis of the effects of osteocyte density on osteonal morphology and remodeling. Bone 33, 753–759 (2003).

16

Buenzli, P. R. & Sims, N. A. Quantifying the osteocyte network in the human skeleton. Bone 75, 144–150 (2015).

17

Marotti, G. & Palumbo, C. The mechanism of transduction of mechanical strains into biological signals at the bone cellular level. Eur. J. Histochem. 51, 15–19 (2007).

18

Tiede-Lewis, L. M. et al. Degeneration of the osteocyte network in the C57BL/6 mouse model of aging. Aging 9, 2190–2208 (2017).

19

Lecanda, F. et al. Connexin43 deficiency causes delayed ossification, craniofacial abnormalities, and osteoblast dysfunction. J. Cell Biol. 151, 931–944 (2000).

20

Chung, D. J. et al. Low peak bone mass and attenuated anabolic response to parathyroid hormone in mice with an osteoblast-specific deletion of connexin43. J. Cell Sci. 119, 4187–4198 (2006).

21

Zhang, Y. et al. Enhanced osteoclastic resorption and responsiveness to mechanical load in gap junction deficient bone. PLoS One 6, e23516 (2011).

22

Green, J. Role of bone in regulation of systemic acid-base balance. Miner. Electrolyte Metab. 20, 7–30 (1994).

23

Bushinsky, D. A., Chabala, J. M. & Levi-Setti, R. Ion microprobe analysis of mouse calvariae in vitro: evidence for a “bone membrane”. Am. J. Physiol. Endocrinol. Metab. 256, E152–E158 (1989).

24

Bushinsky, D. A., Gavrilov, K., Chabala, J. M., Featherstone, J. D. & Levi-Setti, R. Effect of metabolic acidosis on the potassium content of bone. J. Bone Miner. Res. 12, 1664–1671 (1997).

25

Rubinacci, A., Benelli, F. D., Borgo, E. & Villa, I. Bone as an ion exchange system: evidence for a pump-leak mechanism devoted to the maintenance of high bone K+. Am. J. Physiol. Endocrinol. Metab. 278, E15–E24 (2000).

26

Dedic, C. et al. Calcium fluxes at the bone/plasma interface: acute effects of parathyroid hormone (PTH) and targeted deletion of PTH/PTH-related peptide (PTHrP) receptor in the osteocytes. Bone 116, 135–143 (2018).

27

Rubinacci, A. et al. Bone as an ion exchange system: evidence for a link between mechanotransduction and metabolic needs. Am. J. Physiol. Endocrinol. Metab. 282, E851–E864 (2002).

28

Qin, L., Liu, W., Cao, H. & Xiao, G. Molecular mechanosensors in osteocytes. Bone Res. 8, 23 (2020).

29

McNamara, L. M., Majeska, R. J., Weinbaum, S., Friedrich, V. & Schaffler, M. B. Attachment of osteocyte cell processes to the bone matrix. Anat. Rec. 292, 355–363 (2009).

30

Cabahug-Zuckerman, P. et al. Potential role for a specialized β(3) integrin-based structure on osteocyte processes in bone mechanosensation. J. Orthop. Res. 36, 642–652 (2018).

31

Malone, A. M. D. et al. Primary cilia mediate mechanosensing in bone cells by a calcium-independent mechanism. Proc. Natl. Acad. Sci. USA. 104, 13325–13330 (2007).

32

Plotkin, L. I., Speacht, T. L. & Donahue, H. J. Cx43 and mechanotransduction in bone. Curr. Osteoporos. Rep. 13, 67–72 (2015).

33

Wang, L. et al. Mechanical sensing protein PIEZO1 regulates bone homeostasis via osteoblast-osteoclast crosstalk. Nat. Commun. 11, 282 (2020).

34

Lyons, J. S. et al. Microtubules tune mechanotransduction through NOX2 and TRPV4 to decrease sclerostin abundance in osteocytes. Sci. Signal. 10, eaan5748 (2017).

35

Morrell, A. E. et al. Mechanically induced Ca(2+) oscillations in osteocytes release extracellular vesicles and enhance bone formation. Bone Res. 6, 6 (2018).

36

Colombo, M., Raposo, G. & Théry, C. Biogenesis, secretion, and intercellular interactions of exosomes and other extracellular vesicles. Annu. Rev. Cell Dev. Biol. 30, 255–289 (2014).

37

Kamel-ElSayed, S. A., Tiede-Lewis, L. M., Lu, Y., Veno, P. A. & Dallas, S. L. Novel approaches for two and three dimensional multiplexed imaging of osteocytes. Bone 76, 129–140 (2015).

38

Baron, R. & Kneissel, M. Wnt signaling in bone homeostasis and disease: from human mutations to treatments. Nat. Med. 19, 179–192 (2013).

39

Karner, C. M. & Long, F. Wnt signaling and cellular metabolism in osteoblasts. Cell. Mol. Life Sci. 74, 1649–1657 (2017).

40

Day, T. F., Guo, X., Garrett-Beal, L. & Yang, Y. Wnt/β-catenin signaling in mesenchymal progenitors controls osteoblast and chondrocyte differentiation during vertebrate skeletogenesis. Dev. Cell 8, 739–750 (2005).

41

Hu, H. et al. Sequential roles of Hedgehog and Wnt signaling in osteoblast development. Development 132, 49–60 (2005).

42

Song, L. et al. Loss of Wnt/β-catenin signaling causes cell fate shift of preosteoblasts from osteoblasts to adipocytes. J. Bone Miner. Res. 27, 2344–2358 (2012).

43

Gaur, T. et al. Canonical Wnt signaling promotes osteogenesis by directly stimulating Runx2 gene expression. J. Biol. Chem. 280, 33132–33140 (2005).

44

Sato, T. et al. A FAK/HDAC5 signaling axis controls osteocyte mechanotransduction. Nat. Commun. 11, 3282 (2020).

45

Wein, M. N. et al. HDAC5 controls MEF2C-driven sclerostin expression in osteocytes. J. Bone Miner. Res. 30, 400–411 (2015).

46

Wein, M. N. et al. SIKs control osteocyte responses to parathyroid hormone. Nat. Commun. 7, 13176 (2016).

47

Stegen, S. et al. Osteocytic oxygen sensing controls bone mass through epigenetic regulation of sclerostin. Nat. Commun. 9, 2557 (2018).

48

Verborgt, O., Gibson, G. J. & Schaffler, M. B. Loss of osteocyte integrity in association with microdamage and bone remodeling after fatigue in vivo. J. Bone Miner. Res. 15, 60–67 (2000).

49

Boyce, B. F. & Xing, L. Functions of RANKL/RANK/OPG in bone modeling and remodeling. Arch. Biochem. Biophys. 473, 139–146 (2008).

50

Holliday, L. S., Patel, S. S. & Rody, W. J. Jr. RANKL and RANK in extracellular vesicles: surprising new players in bone remodeling. Extracell. Vesicles Circ. Nucleic Acids 2, 18–28 (2021).

51

Nakashima, T. et al. Evidence for osteocyte regulation of bone homeostasis through RANKL expression. Nat. Med. 17, 1231–1234 (2011).

52

Xiong, J. et al. Matrix-embedded cells control osteoclast formation. Nat. Med. 17, 1235–1241 (2011).

53

Lim, J., Burclaff, J., He, G., Mills, J. C. & Long, F. Unintended targeting of Dmp1-Cre reveals a critical role for Bmpr1a signaling in the gastrointestinal mesenchyme of adult mice. Bone Res. 5, 16049 (2017).

54

Xiong, J. et al. Osteocytes, not osteoblasts or lining cells, are the main source of the RANKL required for osteoclast formation in remodeling bone. PLoS One 10, e0138189 (2015).

55

Zhao, S. et al. MLO-Y4 osteocyte-like cells support osteoclast formation and activation. J. Bone Miner. Res. 17, 2068–2079 (2002).

56

Al-Dujaili, S. A. et al. Apoptotic osteocytes regulate osteoclast precursor recruitment and differentiation in vitro. J. Cell. Biochem. 112, 2412–2423 (2011).

57

Honma, M. et al. RANKL subcellular trafficking and regulatory mechanisms in osteocytes. J. Bone Miner. Res. 28, 1936–1949 (2013).

58

Honma, M., Ikebuchi, Y. & Suzuki, H. Mechanisms of RANKL delivery to the osteoclast precursor cell surface. J. Bone Miner. Metab. 39, 27–33 (2021).

59

Xiong, J. et al. Soluble RANKL contributes to osteoclast formation in adult mice but not ovariectomy-induced bone loss. Nat. Commun. 9, 2909 (2018).

60

Mizuno, A. et al. Transgenic mice overexpressing soluble osteoclast differentiation factor (sODF) exhibit severe osteoporosis. J. Bone Miner. Metab. 20, 337–344 (2002).

61

Komori, T. Cell death in chondrocytes, osteoblasts, and osteocytes. Int. J. Mol. Sci. 17, 2045 (2016).

62

Emerton, K. B. et al. Osteocyte apoptosis and control of bone resorption following ovariectomy in mice. Bone 46, 577–583 (2010).

63

Kennedy, O. D., Laudier, D. M., Majeska, R. J., Sun, H. B. & Schaffler, M. B. Osteocyte apoptosis is required for production of osteoclastogenic signals following bone fatigue in vivo. Bone 64, 132–137 (2014).

64

Aguirre, J. I. et al. Osteocyte apoptosis is induced by weightlessness in mice and precedes osteoclast recruitment and bone loss. J. Bone Miner. Res. 21, 605–615 (2006).

65

Kennedy, O. D. et al. Activation of resorption in fatigue-loaded bone involves both apoptosis and active pro-osteoclastogenic signaling by distinct osteocyte populations. Bone 50, 1115–1122 (2012).

66

Kringelbach, T. M., Aslan, D., Novak, I., Schwarz, P. & Jørgensen, N. R. UTP-induced ATP release is a fine-tuned signalling pathway in osteocytes. Purinergic Signal 10, 337–347 (2014).

67

Cheung, W. Y. et al. Pannexin-1 and P2X7-receptor are required for apoptotic osteocytes in fatigued bone to trigger RANKL production in neighboring bystander osteocytes. J. Bone Miner. Res. 31, 890–899 (2016).

68

Luckprom, P., Wongkhantee, S., Yongchaitrakul, T. & Pavasant, P. Adenosine triphosphate stimulates RANKL expression through P2Y1receptor-cyclo-oxygenase-dependent pathway in human periodontal ligament cells. J. Periodontal. Res. 45, 404–411 (2010).

69

Elliott, M. R. et al. Nucleotides released by apoptotic cells act as a find-me signal to promote phagocytic clearance. Nature 461, 282–286 (2009).

70

Verborgt, O., Tatton, N. A., Majeska, R. J. & Schaffler, M. B. Spatial distribution of bax and Bcl-2 in osteocytes after bone fatigue: complementary roles in bone remodeling regulation? J. Bone Miner. Res. 17, 907–914 (2002).

71

Takeuchi, O. & Akira, S. Pattern recognition receptors and inflammation. Cell 140, 805–820 (2010).

72

Messmer, D. et al. High mobility group box protein 1: an endogenous signal for dendritic cell maturation and Th1 polarization. J. Immunol. 173, 307–313 (2004).

73

Mori, T. et al. IL-1β and TNFα-initiated IL-6-STAT3 pathway is critical in mediating inflammatory cytokines and RANKL expression in inflammatory arthritis. Int. Immunol. 23, 701–712 (2011).

74

Andreev, D. et al. Osteocyte necrosis triggers osteoclast-mediated bone loss through macrophage-inducible C-type lectin. J. Clin. Investig. 130, 4811–4830 (2020).

75

Glass, D. A. et al. Canonical Wnt signaling in differentiated osteoblasts controls osteoclast differentiation. Dev. Cell 8, 751–764 (2005).

76

Walsh, M. C. & Choi, Y. Biology of the RANKL-RANK-OPG system in immunity, bone, and beyond. Front. Immunol. 5, 511 (2014).

77

Cawley, K. M. et al. Local production of osteoprotegerin by osteoblasts suppresses bone resorption. Cell Rep. 32, 108052 (2020).

78

Delgado-Calle, J. et al. Control of bone anabolism in response to mechanical loading and PTH by distinct mechanisms downstream of the PTH receptor. J. Bone Miner. Res. 32, 522–535 (2017).

79

Ben-Awadh, A. N. et al. Parathyroid hormone receptor signaling induces bone resorption in the adult skeleton by directly regulating the RANKL gene in osteocytes. Endocrinology 155, 2797–2809 (2014).

80

Rhee, Y. et al. PTH receptor signaling in osteocytes governs periosteal bone formation and intracortical remodeling. J. Bone Miner. Res. 26, 1035–1046 (2011).

81

Saini, V. et al. Parathyroid hormone (PTH)/PTH-related peptide type 1 receptor (PPR) signaling in osteocytes regulates anabolic and catabolic skeletal responses to PTH. J. Biol. Chem. 288, 20122–20134 (2013).

82

Hongo, H. et al. Osteocytic osteolysis in PTH-treated wild-type and rankl(-/-) mice examined by transmission electron microscopy, atomic force microscopy, and isotope microscopy. J. Histochem. Cytochem. 68, 651–668 (2020).

83

Tazawa, K. et al. Osteocytic osteolysis observed in rats to which parathyroid hormone was continuously administered. J. Bone Miner. Metab. 22, 524–529 (2004).

84

Misof, B. M. et al. Bone matrix mineralization and osteocyte lacunae characteristics in patients with chronic kidney disease—mineral bone disorder(CKD-MBD). J. Musculoskelet. Neuronal Interact. 19, 196–206 (2019).

85

Gardinier, J. D., Al-Omaishi, S., Morris, M. D. & Kohn, D. H. PTH signaling mediates perilacunar remodeling during exercise. Matrix Biol. 52-54, 162–175 (2016).

86

Qing, H. et al. Demonstration of osteocytic perilacunar/canalicular remodeling in mice during lactation. J. Bone Miner. Res. 27, 1018–1029 (2012).

87

Misof, B. M. et al. No role of osteocytic osteolysis in the development and recovery of the bone phenotype induced by severe secondary hyperparathyroidism in vitamin D receptor deficient mice. Int. J. Mol. Sci. 21, 7989 (2020).

88

Davesne, D. et al. The phylogenetic origin and evolution of acellular bone in teleost fishes: insights into osteocyte function in bone metabolism. Biol. Rev. Camb. Philos. Soc. 94, 1338–1363 (2019).

89

Hung, J. T. et al. Assessing the ability of zebrafish scales to contribute to the short-term homeostatic regulation of [Ca2+] in the extracellular fluid during calcemic challenges. Fish. Sci. 85, 943–959 (2019).

90

Marenzana, M., Shipley, A. M., Squitiero, P., Kunkel, J. G. & Rubinacci, A. Bone as an ion exchange organ: evidence for instantaneous cell-dependent calcium efflux from bone not due to resorption. Bone 37, 545–554 (2005).

91

Farr, J. N., Kaur, J., Doolittle, M. L. & Khosla, S. Osteocyte cellular senescence. Curr. Osteoporos. Rep. 18, 559–567 (2020).

92

Farr, J. N. et al. Identification of senescent cells in the bone microenvironment. J. Bone Miner. Res. 31, 1920–1929 (2016).

93

Uda, Y. et al. Parathyroid hormone signaling in mature osteoblasts/osteocytes protects mice from age-related bone loss. Aging 13, 25607–25642 (2021).

94

Hayashi, M. et al. Autoregulation of osteocyte Sema3A orchestrates estrogen action and counteracts bone aging. Cell Metab. 29, 627–637.e5 (2019).

95

Sharma, D. et al. The effects of estrogen deficiency on cortical bone microporosity and mineralization. Bone 110, 1–10 (2018).

96

Jackson, E. et al. Osteocyte Wnt/β-catenin pathway activation upon mechanical loading is altered in ovariectomized mice. Bone Rep. 15, 101129 (2021).

97

Ma, L. et al. Connexin 43 hemichannels protect bone loss during estrogen deficiency. Bone Res. 7, 11 (2019).

98

Dole, N. S. et al. Osteocyte-intrinsic TGF-β signaling regulates bone quality through perilacunar/canalicular remodeling. Cell Rep. 21, 2585–2596 (2017).

99

Turner, A. G., Hanrath, M. A., Morris, H. A., Atkins, G. J. & Anderson, P. H. The local production of 1,25(OH)2D3 promotes osteoblast and osteocyte maturation. J. Steroid Biochem. Mol. Biol. 144, 114–118 (2014).

100

Yashiro, M. et al. Active vitamin D and vitamin D analogs stimulate fibroblast growth factor 23 production in osteocyte-like cells via the vitamin D receptor. J. Pharm. Biomed. Anal. 182, 113139 (2020).

101

Nociti, F. H. Jr. et al. Vitamin D represses dentin matrix protein 1 in cementoblasts and osteocytes. J. Dent. Res. 93, 148–154 (2014).

102

Rolvien, T. et al. Vitamin D regulates osteocyte survival and perilacunar remodeling in human and murine bone. Bone 103, 78–87 (2017).

103

Yuan, Y. et al. Impaired 1,25 dihydroxyvitamin D3 action and hypophosphatemia underlie the altered lacuno-canalicular remodeling observed in the Hyp mouse model of XLH. PLoS One 16, e0252348 (2021).

104

Ito, N., Findlay, D. M., Anderson, P. H., Bonewald, L. F. & Atkins, G. J. Extracellular phosphate modulates the effect of 1α,25-dihydroxy vitamin D3 (1,25D) on osteocyte like cells. J. Steroid Biochem. Mol. Biol. 136, 183–186 (2013).

105

Takashi, Y. & Fukumoto, S. Phosphate-sensing and regulatory mechanism of FGF23 production. J. Endocrinol. Investig. 43, 877–883 (2020).

106

Gooi, J. H. et al. Decline in calcitonin receptor expression in osteocytes with age. J. Endocrinol. 221, 181–191 (2014).

107

Davey, R. A. & Findlay, D. M. Calcitonin: physiology or fantasy? J. Bone Miner. Res. 28, 973–979 (2013).

108

He, Z. et al. Irisin inhibits osteocyte apoptosis by activating the Erk signaling pathway in vitro and attenuates ALCT-induced osteoarthritis in mice. Bone 141, 115573 (2020).

109

Wood, C. L., Pajevic, P. D. & Gooi, J. H. Osteocyte secreted factors inhibit skeletal muscle differentiation. Bone Rep. 6, 74–80 (2017).

110

Du, J. H. et al. The function of Wnt ligands on osteocyte and bone remodeling. J. Dent. Res. 98, 930–938 (2019).

111

Miller, S. C., Bowman, B. M., Smith, J. M. & Jee, W. S. S. Characterization of endosteal bone-lining cells from fatty marrow bone sites in adult beagles. Anat. Rec. 198, 163–173 (1980).

112

Miller, S. C., de Saint-Georges, L., Bowman, B. M. & Jee, W. S. Bone lining cells: structure and function. Scanning Microsc. 3, 953–960 (1989).

113

Everts, V. et al. The bone lining cell: its role in cleaning howship’s lacunae and initiating bone formation. J. Bone Miner. Res. 17, 77–90 (2002).

114

Bloemen, V., de Vries, T. J., Schoenmaker, T. & Everts, V. Intercellular adhesion molecule-1 clusters during osteoclastogenesis. Biochem. Biophys. Res. Commun. 385, 640–645 (2009).

115

Talmage, R. V. Morphological and physiological considerations in a new concept of calcium transport in bone. Am. J. Anat. 129, 467–476 (1970).

116

Wongdee, K., Riengrojpitak, S., Krishnamra, N. & Charoenphandhu, N. Claudin expression in the bone-lining cells of female rats exposed to long-standing acidemia. Exp. Mol. Pathol. 88, 305–310 (2010).

117

Andersen, T. L. et al. A physical mechanism for coupling bone resorption and formation in adult human bone. Am. J. Pathol. 174, 239–247 (2009).

118

Fields, G. B. Interstitial collagen catabolism. J. Biol. Chem. 288, 8785–8793 (2013).

119

Sprangers, S. & Everts, V. Molecular pathways of cell-mediated degradation of fibrillar collagen. Matrix Biol. 75-76, 190–200 (2019).

120

Delaissé, J. M. et al. Matrix metalloproteinases (MMP) and cathepsin K contribute differently to osteoclastic activities. Microsc. Res. Tech. 61, 504–513 (2003).

121

Shi, J. et al. Membrane-type MMPs enable extracellular matrix permissiveness and mesenchymal cell proliferation during embryogenesis. Dev. Biol. 313, 196–209 (2008).

122

Parfitt, A. M. The bone remodeling compartment: a circulatory function for bone lining cells. J. Bone Miner. Res. 16, 1583–1585 (2001).

123

Hauge, E. M., Qvesel, D., Eriksen, E. F., Mosekilde, L. & Melsen, F. Cancellous bone remodeling occurs in specialized compartments lined by cells expressing osteoblastic markers. J. Bone Miner. Res. 16, 1575–1582 (2001).

124

Kristensen, H. B., Andersen, T. L., Marcussen, N., Rolighed, L. & Delaisse, J. M. Increased presence of capillaries next to remodeling sites in adult human cancellous bone. J. Bone Miner. Res. 28, 574–585 (2013).

125

Díaz-Flores, L. et al. Pericytes. Morphofunction, interactions and pathology in a quiescent and activated mesenchymal cell niche. Histol. Histopathol. 24, 909–969 (2009).

126

Delaisse, J. M. The reversal phase of the bone-remodeling cycle: cellular prerequisites for coupling resorption and formation. Bonekey Rep. 3, 561 (2014).

127

Eghbali-Fatourechi, G. Z. et al. Circulating osteoblast-lineage cells in humans. N. Engl. J. Med. 352, 1959–1966 (2005).

128

Kusumbe, A. P., Ramasamy, S. K. & Adams, R. H. Coupling of angiogenesis and osteogenesis by a specific vessel subtype in bone. Nature 507, 323–328 (2014).

129

Ramasamy, S. K., Kusumbe, A. P., Wang, L. & Adams, R. H. Endothelial Notch activity promotes angiogenesis and osteogenesis in bone. Nature 507, 376–380 (2014).

130

Xie, H. et al. PDGF-BB secreted by preosteoclasts induces angiogenesis during coupling with osteogenesis. Nat. Med. 20, 1270–1278 (2014).

131

Kim, B. J. et al. Osteoclast-secreted SLIT3 coordinates bone resorption and formation. J. Clin. Investig. 128, 1429–1441 (2018).

132

Qiu, H., Xiao, W., Yue, J. & Wang, L. Heparan sulfate modulates Slit3-induced endothelial cell migration. Methods Mol. Biol. 1229, 549–555 (2015).

133

Geutskens, S. B. et al. Control of human hematopoietic stem/progenitor cell migration by the extracellular matrix protein Slit3. Lab. Investig. 92, 1129–1139 (2012).

134

Zhou, B. O., Yue, R., Murphy, M. M., Peyer, J. G. & Morrison, S. J. Leptin-receptor-expressing mesenchymal stromal cells represent the main source of bone formed by adult bone marrow. Cell Stem Cell 15, 154–168 (2014).

135

Caire, R. et al. Parathyroid hormone remodels bone transitional vessels and the leptin receptor‐positive pericyte network in mice. J. Bone Miner. Res. 34, 1487–1501 (2019).

136

Ominsky, M. S. et al. Differential temporal effects of sclerostin antibody and parathyroid hormone on cancellous and cortical bone and quantitative differences in effects on the osteoblast lineage in young intact rats. Bone 81, 380–391 (2015).

137

Matic, I. et al. Quiescent bone lining cells are a major source of osteoblasts during adulthood. Stem Cells 34, 2930–2942 (2016).

138

Collin-Osdoby, P. et al. Receptor activator of NF-kappa B and osteoprotegerin expression by human microvascular endothelial cells, regulation by inflammatory cytokines, and role in human osteoclastogenesis. J. Biol. Chem. 276, 20659–20672 (2001).

139

Søe, K., Delaisse, J. M. & Borggaard, X. G. Osteoclast formation at the bone marrow/bone surface interface: importance of structural elements, matrix, and intercellular communication. Semin. Cell Dev. Biol. 112, 8–15 (2020).

140

Ferrier, J., Xia, S. L., Lagan, E., Aubin, J. E. & Heersche, J. N. M. Displacement and translocation of osteoblast-like cells by osteoclasts. J. Bone Miner. Res. 9, 1397–1405 (1994).

141

Karsdal, M. A., Fjording, M. S., Foged, N. T., Delaissé, J. M. & Lochter, A. Transforming growth factor-β-induced osteoblast elongation regulates osteoclastic bone resorption through a p38 mitogen-activated protein kinase- and matrix metalloproteinase-dependent pathway. J. Biol. Chem. 276, 39350–39358 (2001).

142

Saltel, F. et al. Transmigration: a new property of mature multinucleated osteoclasts. J. Bone Miner. Res. 21, 1913–1923 (2006).

143

Perez-Amodio, S., Beertsen, W. & Everts, V. (Pre-)osteoclasts induce retraction of osteoblasts before their fusion to osteoclasts. J. Bone Miner. Res. 19, 1722–1731 (2004).

144

Kajita, M. et al. Membrane-type 1 matrix metalloproteinase cleaves CD44 and promotes cell migration. J. Cell Biol. 153, 893–904 (2001).

145

Blavier, L. & Delaisse, J. M. Matrix metalloproteinases are obligatory for the migration of preosteoclasts to the developing marrow cavity of primitive long bones. J. Cell Sci. 108, 3649–3659 (1995).

146

Sato, T., Foged, N. T. & Delaissé, J. M. The migration of purified osteoclasts through collagen is inhibited by matrix metalloproteinase inhibitors. J. Bone Miner. Res. 13, 59–66 (1998).

147

Andersen, T. L. et al. A scrutiny of matrix metalloproteinases in osteoclasts: evidence for heterogeneity and for the presence of MMPs synthesized by other cells. Bone 35, 1107–1119 (2004).

148

Holmbeck, K. et al. MT1-MMP-deficient mice develop dwarfism, osteopenia, arthritis, and connective tissue disease due to inadequate collagen turnover. Cell 99, 81–92 (1999).

149

Sato, T. et al. Identification of the membrane-type matrix metalloproteinase MT1-MMP in osteoclasts. J. Cell Sci. 110, 589–596 (1997).

150

Ohuchi, E. et al. Membrane type 1 matrix metalloproteinase digests interstitial collagens and other extracellular matrix macromolecules. J. Biol. Chem. 272, 2446–2451 (1997).

151

Engsig, M. T. et al. Matrix metalloproteinase 9 and vascular endothelial growth factor are essential for osteoclast recruitment into developing long bones. J. Cell Biol. 151, 879–889 (2000).

152

Chellaiah, M. A. & Ma, T. Membrane localization of membrane type 1 matrix metalloproteinase by CD44 regulates the activation of pro-matrix metalloproteinase 9 in osteoclasts. Biomed. Res. Int. 2013, 302392 (2013).

153

Zhu, L. et al. Osteoclast-mediated bone resorption is controlled by a compensatory network of secreted and membrane-tethered metalloproteinases. Sci. Transl. Med. 12, eaaw6143 (2020).

154

Møller, A. M. J. et al. Aging and menopause reprogram osteoclast precursors for aggressive bone resorption. Bone Res. 8, 27 (2020).

155

Weinstein, R. S., Roberson, P. K. & Manolagas, S. C. Giant osteoclast formation and long-term oral bisphosphonate therapy. N. Engl. J. Med. 360, 53–62 (2009).

156

Jobke, B., Milovanovic, P., Amling, M. & Busse, B. Bisphosphonate-osteoclasts: changes in osteoclast morphology and function induced by antiresorptive nitrogen-containing bisphosphonate treatment in osteoporosis patients. Bone 59, 37–43 (2014).

157

Takagi, Y. et al. Effect of nitrogen-containing bisphosphonates on osteoclasts and osteoclastogenesis: an ultrastructural study. Microscopy 70, 302–307 (2021).

158

Liu, H. et al. Histochemical evidence of zoledronate inhibiting c-src expression and interfering with CD44/OPN-mediated osteoclast adhesion in the tibiae of mice. J. Mol. Histol. 46, 313–323 (2015).

159

Yagi, M. et al. DC-STAMP is essential for cell-cell fusion in osteoclasts and Foreign body giant cells. J. Exp. Med. 202, 345–351 (2005).

160

Takito, J. & Nakamura, M. Heterogeneity and actin cytoskeleton in osteoclast and macrophage multinucleation. Int. J. Mol. Sci. 21, 6629 (2020).

161

Streicher, C. et al. Estrogen regulates bone turnover by targeting RANKL expression in bone lining cells. Sci. Rep. 7, 6460 (2017).

162

Takayanagi, H. Osteoimmunology: shared mechanisms and crosstalk between the immune and bone systems. Nat. Rev. Immunol. 7, 292–304 (2007).

163

Karin, M. & Greten, F. R. NF-κB: linking inflammation and immunity to cancer development and progression. Nat. Rev. Immunol. 5, 749–759 (2005).

164

Hayden, M. S., West, A. P. & Ghosh, S. NF-κB and the immune response. Oncogene 25, 6758–6780 (2006).

165

Franzoso, G. et al. Requirement for NF-kappaB in osteoclast and B-cell development. Genes Dev. 11, 3482–3496 (1997).

166

Xing, L. et al. NF-κB p50 and p52 expression is not required for RANK-expressing osteoclast progenitor formation but is essential for RANK- and cytokine-mediated osteoclastogenesis. J. Bone Miner. Res. 17, 1200–1210 (2002).

167

Asagiri, M. et al. Autoamplification of NFATc1 expression determines its essential role in bone homeostasis. J. Exp. Med. 202, 1261–1269 (2005).

168

Ishida, N. et al. Large scale gene expression analysis of osteoclastogenesis in vitro and elucidation of NFAT2 as a key regulator. J. Biol. Chem. 277, 41147–41156 (2002).

169

Takayanagi, H. et al. Induction and activation of the transcription factor NFATc1 (NFAT2) integrate RANKL signaling in terminal differentiation of osteoclasts. Dev. Cell 3, 889–901 (2002).

170

Ruocco, M. G. et al. I{kappa}B kinase (IKK){beta}, but not IKK{alpha}, is a critical mediator of osteoclast survival and is required for inflammation-induced bone loss. J. Exp. Med. 201, 1677–1687 (2005).

171

Xiu, Y. et al. Chloroquine reduces osteoclastogenesis in murine osteoporosis by preventing TRAF3 degradation. J. Clin. Investig. 124, 297–310 (2014).

172

Luo, J. et al. LGR4 is a receptor for RANKL and negatively regulates osteoclast differentiation and bone resorption. Nat. Med. 22, 539–546 (2016).

173

Ko, Y. J. et al. A novel modified RANKL variant can prevent osteoporosis by acting as a vaccine and an inhibitor. Clin. Transl. Med. 11, e368 (2021).

174

Yang, X., Pande, S., Scott, C. & Friesel, R. Macrophage colony-stimulating factor pretreatment of bone marrow progenitor cells regulates osteoclast differentiation based upon the stage of myeloid development. J. Cell. Biochem. 120, 12450–12460 (2019).

175

Faccio, R., Takeshita, S., Zallone, A., Ross, F. P. & Teitelbaum, S. L. c-Fms and the αvβ3 integrin collaborate during osteoclast differentiation. J. Clin. Investig. 111, 749–758 (2003).

176

Scholtysek, C. et al. NR4A1 regulates motility of osteoclast precursors and serves as target for the modulation of systemic bone turnover. J. Bone Miner. Res. 33, 2035–2047 (2018).

177

Koga, T. et al. Costimulatory signals mediated by the ITAM motif cooperate with RANKL for bone homeostasis. Nature 428, 758–763 (2004).

178

Souza, P. P. C. & Lerner, U. H. Finding a toll on the route: the fate of osteoclast progenitors after toll-like receptor activation. Front. Immunol. 10, 1663 (2019).

179

Yim, M. The role of toll-like receptors in osteoclastogenesis. J. Bone Metab. 27, 227–235 (2020).

180

Maeda, K. et al. Wnt5a-Ror2 signaling between osteoblast-lineage cells and osteoclast precursors enhances osteoclastogenesis. Nat. Med. 18, 405–412 (2012).

181

Kobayashi, Y. et al. Wnt16 regulates osteoclast differentiation in conjunction with Wnt5a. Biochem. Biophys. Res. Commun. 463, 1278–1283 (2015).

182

Uehara, S. et al. Protein kinase N3 promotes bone resorption by osteoclasts in response to Wnt5a-Ror2 signaling. Sci. Signal. 10, eaan0023 (2017).

183

Okamoto, M. et al. Noncanonical Wnt5a enhances Wnt/β-catenin signaling during osteoblastogenesis. Sci. Rep. 4, 4493 (2014).

184

Roberts, J. L. et al. Deletion of Wnt5a in osteoclasts results in bone loss through decreased bone formation. Ann. N. Y. Acad. Sci. 1463, 45–59 (2020).

185

Akisaka, T. The ruffled border and attachment regions of the apposing membrane of resorbing osteoclasts as visualized from the cytoplasmic face of the membrane. J. Electron Microsc. 55, 53–61 (2006).

186

Yao, G., Feng, H., Cai, Y., Qi, W. & Kong, K. Characterization of vacuolar-ATPase and selective inhibition of vacuolar-H(+)-ATPase in osteoclasts. Biochem. Biophys. Res. Commun. 357, 821–827 (2007).

187

Kornak, U. et al. Loss of the ClC-7 chloride channel leads to osteopetrosis in mice and man. Cell 104, 205–215 (2001).

188

Delaissé, J. M. et al. Proteinases in bone resorption: obvious and less obvious roles. Clin. Chim. Acta 291, 223–234 (2000).

189

Stenbeck, G. & Horton, M. A. Endocytic trafficking in actively resorbing osteoclasts. J. Cell Sci. 117, 827–836 (2004).

190

Helfrich, M. H., Nesbitt, S. A., Dorey, E. L. & Horton, M. A. Rat osteoclasts adhere to a wide range of rgd (arg-gly-asp) peptide-containing proteins, including the bone sialoproteins and fibronectin, via a β3 integrin. J. Bone Miner. Res. 7, 335–343 (1992).

191

Faccio, R. et al. Activation of αvβ3 integrin on human osteoclast-like cells stimulates adhesion and migration in response to osteopontin. Biochem. Biophys. Res. Commun. 249, 522–525 (1998).

192

Ross, F. P. et al. Interactions between the bone matrix proteins osteopontin and bone sialoprotein and the osteoclast integrin alpha v beta 3 potentiate bone resorption. J. Biol. Chem. 268, 9901–9907 (1993).

193

Faccio, R. et al. Localization and possible role of two different alpha v beta 3 integrin conformations in resting and resorbing osteoclasts. J. Cell Sci. 115, 2919–2929 (2002).

194

Duong, L. T., Lakkakorpi, P., Nakamura, I. & Rodan, G. A. Integrins and signaling in osteoclast function. Matrix Biol. 19, 97–105 (2000).

195

Xiang, B. et al. The osteoclasts attach to the bone surface where the extracellular calcium concentration decreases. Cell Biochem. Biophys. 74, 553–558 (2016).

196

Sims, N. A. & Martin, T. J. Osteoclasts provide coupling signals to osteoblast lineage cells through multiple mechanisms. Annu. Rev. Physiol. 82, 507–529 (2020).

197

Jiang, X. et al. Method development of efficient protein extraction in bone tissue for proteome analysis. J. Proteome Res. 6, 2287–2294 (2007).

198

Rubinacci, A. Expanding the functional spectrum of vitamin K in bone. Focus on: “vitamin K promotes mineralization, osteoblast to osteocyte transition, and an anti-catabolic phenotype by γ-carboxylation-dependent and -independent mechanisms”. Am. J. Physiol. Cell Physiol. 297, C1336–C1338 (2009).

199

Mizokami, A., Kawakubo-Yasukochi, T. & Hirata, M. Osteocalcin and its endocrine functions. Biochem. Pharmacol. 132, 1–8 (2017).

200

Lin, X., Patil, S., Gao, Y. G. & Qian, A. The bone extracellular matrix in bone formation and regeneration. Front. Pharmacol. 11, 757 (2020).

201

Hering, S. et al. TGFβ1 and TGFβ2 mRNA and protein expression in human bone samples. Exp. Clin. Endocrinol. Diabetes 109, 217–226 (2001).

202

Oursler, M. J. Osteoclast synthesis and secretion and activation of latent transforming growth factor β. J. Bone Miner. Res. 9, 443–452 (1994).

203

Tang, Y. et al. TGF-beta1-induced migration of bone mesenchymal stem cells couples bone resorption with formation. Nat. Med. 15, 757–765 (2009).

204

Alliston, T., Choy, L., Ducy, P., Karsenty, G. & Derynck, R. TGF-beta-induced repression of CBFA1 by Smad3 decreases cbfa1 and osteocalcin expression and inhibits osteoblast differentiation. EMBO J. 20, 2254–2272 (2001).

205

Breen, E. C. et al. TGF? alters growth and differentiation related gene expression in proliferating osteoblasts in vitro, preventing development of the mature bone phenotype. J. Cell. Physiol. 160, 323–335 (1994).

206

Weivoda, M. M. et al. Osteoclast TGF-β receptor signaling induces Wnt1 secretion and couples bone resorption to bone formation. J. Bone Miner. Res. 31, 76–85 (2016).

207

Li, J. et al. TGFβ-induced degradation of TRAF3 in mesenchymal progenitor cells causes age-related osteoporosis. Nat. Commun. 10, 2795 (2019).

208

Kim, J. S. et al. Transforming growth factor-β1 regulates macrophage migration via RhoA. Blood 108, 1821–1829 (2006).

209

Takai, H. et al. Transforming growth factor-β stimulates the production of osteoprotegerin/osteoclastogenesis inhibitory factor by bone marrow stromal cells. J. Biol. Chem. 273, 27091–27096 (1998).

210

Houde, N., Chamoux, E., Bisson, M. & Roux, S. Transforming growth factor-beta1 (TGF-beta1) induces human osteoclast apoptosis by up-regulating Bim. J. Biol. Chem. 284, 23397–23404 (2009).

211

Ota, K. et al. TGF-β induces Wnt10b in osteoclasts from female mice to enhance coupling to osteoblasts. Endocrinology 154, 3745–3752 (2013).

212

Seck, T. et al. Concentration of insulin-like growth factor (IGF)-I and -II in iliac crest bone matrix from pre- and postmenopausal women: relationship to age, menopause, bone turnover, bone volume, and circulating IGFs. J. Clin. Endocrinol. Metab. 83, 2331–2337 (1998).

213

Lean, J. M., Mackay, A. G., Chow, J. W. & Chambers, T. J. Osteocytic expression of mRNA for c-fos and IGF-I: an immediate early gene response to an osteogenic stimulus. Am. J. Physiol. Endocrinol. Metab. 270, E937–E945 (1996).

214

Lau, K. H. W. et al. Osteocyte-derived insulin-like growth factor I is essential for determining bone mechanosensitivity. Am. J. Physiol. Endocrinol. Metab. 305, E271–E281 (2013).

215

Fiedler, J., Brill, C., Blum, W. F. & Brenner, R. E. IGF-I and IGF-II stimulate directed cell migration of bone-marrow-derived human mesenchymal progenitor cells. Biochem. Biophys. Res. Commun. 345, 1177–1183 (2006).

216

Yuan, Y. et al. Gene expression profiles and bioinformatics analysis of insulin-like growth factor-1 promotion of osteogenic differentiation. Mol. Genet. Genom. Med 7, e00921 (2019).

217

Xian, L. et al. Matrix IGF-1 maintains bone mass by activation of mTOR in mesenchymal stem cells. Nat. Med. 18, 1095–1101 (2012).

218

Liu, L. & Parent, C. A. Review series: TOR kinase complexes and cell migration. J. Cell Biol. 194, 815–824 (2011).

219

Feng, X. et al. Insulin-like growth factor 1 can promote proliferation and osteogenic differentiation of human dental pulp stem cells via mTOR pathway. Dev. Growth Differ. 56, 615–624 (2014).

220

Sims, N. A. & Martin, T. J. Coupling signals between the osteoclast and osteoblast: how are messages transmitted between these temporary visitors to the bone surface? Front. Endocrinol. 6, 41 (2015).

221

Karsdal, M. A., Neutzsky-Wulff, A. V., Dziegiel, M. H., Christiansen, C. & Henriksen, K. Osteoclasts secrete non-bone derived signals that induce bone formation. Biochem. Biophys. Res. Commun. 366, 483–488 (2008).

222

Thudium, C. S. et al. A comparison of osteoclast-rich and osteoclast-poor osteopetrosis in adult mice sheds light on the role of the osteoclast in coupling bone resorption and bone formation. Calcif. Tissue Int. 95, 83–93 (2014).

223

Neutzsky-Wulff, A. V. et al. Severe developmental bone phenotype in ClC-7 deficient mice. Dev. Biol. 344, 1001–1010 (2010).

224

Henriksen, K. et al. Dissociation of bone resorption and bone formation in adult mice with a non-functional V-ATPase in osteoclasts leads to increased bone strength. PLoS One 6, e27482 (2011).

225

McClung, M. R. et al. Odanacatib for the treatment of postmenopausal osteoporosis: results of the LOFT multicenter, randomised, double blind, placebo-controlled trial and LOFT Extension study. Lancet Diabetes Endocrinol. 7, 899–911 (2019).

226

Guerrini, M. M. et al. Human osteoclast-poor osteopetrosis with hypogammaglobulinemia due to TNFRSF11A (RANK) mutations. Am. J. Hum. Genet. 83, 64–76 (2008).

227

Whyte, M. P. et al. Dysosteosclerosis presents as an “osteoclast-poor” form of osteopetrosis: comprehensive investigation of a 3-year-old girl and literature review. J. Bone Miner. Res. 25, 2527–2539 (2010).

228

Walker, E. C. et al. Cardiotrophin-1 is an osteoclast-derived stimulus of bone formation required for normal bone remodeling. J. Bone Miner. Res. 23, 2025–2032 (2008).

229

Meshcheryakova, A., Mechtcheriakova, D. & Pietschmann, P. Sphingosine 1-phosphate signaling in bone remodeling: multifaceted roles and therapeutic potential. Expert Opin. Ther. Targets 21, 725–737 (2017).

230

Pederson, L., Ruan, M., Westendorf, J. J., Khosla, S. & Oursler, M. J. Regulation of bone formation by osteoclasts involves Wnt/BMP signaling and the chemokine sphingosine-1-phosphate. Proc. Natl Acad. Sci. USA 105, 20764–20769 (2008).

231

Takeshita, S. et al. Osteoclast-secreted CTHRC1 in the coupling of bone resorption to formation. J. Clin. Investig. 123, 3914–3924 (2013).

232

Matsuoka, K., Park, K. A., Ito, M., Ikeda, K. & Takeshita, S. Osteoclast-derived complement component 3a stimulates osteoblast differentiation. J. Bone Miner. Res. 29, 1522–1530 (2014).

233

Cornish, J., Callon, K. E., Edgar, S. G. & Reid, I. R. Leukemia inhibitory factor is mitogenic to osteoblasts. Bone 21, 243–247 (1997).

234

Weivoda, M. M. et al. Identification of osteoclast-osteoblast coupling factors in humans reveals links between bone and energy metabolism. Nat. Commun. 11, 87 (2020).

235

Kim, B. J. & Koh, J. M. Coupling factors involved in preserving bone balance. Cell. Mol. Life Sci. 76, 1243–1253 (2019).

236

Liu, M., Sun, Y. & Zhang, Q. Emerging role of extracellular vesicles in bone remodeling. J. Dent. Res. 97, 859–868 (2018).

237

Huynh, N. et al. Characterization of regulatory extracellular vesicles from osteoclasts. J. Dent. Res. 95, 673–679 (2016).

238

Ikebuchi, Y. et al. Coupling of bone resorption and formation by RANKL reverse signalling. Nature 561, 195–200 (2018).

239

Furuya, Y. et al. Stimulation of bone formation in cortical bone of mice treated with a receptor activator of nuclear factor-κB ligand (RANKL)-binding peptide that possesses osteoclastogenesis inhibitory activity. J. Biol. Chem. 288, 5562–5571 (2013).

240

Portal-Núñez, S. et al. Unexpected bone formation produced by RANKL blockade. Trends Endocrinol. Metab. 28, 695–704 (2017).

241

Bone, H. G. et al. 10 years of denosumab treatment in postmenopausal women with osteoporosis: results from the phase 3 randomised FREEDOM trial and open-label extension. Lancet Diabetes Endocrinol. 5, 513–523 (2017).

242

Dempster, D. W. et al. Modeling‐based bone formation in the human femoral neck in subjects treated with denosumab. J. Bone Miner. Res. 35, 1282–1288 (2020).

243

Wang, L., Huang, B., Chen, X. & Su, J. New insight into unexpected bone formation by denosumab. Drug Discov. Today 25, 1919–1922 (2020).

244

Chen, X., Zhi, X., Wang, J. & Su, J. RANKL signaling in bone marrow mesenchymal stem cells negatively regulates osteoblastic bone formation. Bone Res 6, 34 (2018).

245

Amizuka, N. et al. Defective bone remodelling in osteoprotegerin-deficient mice. J. Electron Microsc. 52, 503–513 (2003).

246

Sone, E. et al. The induction of RANKL molecule clustering could stimulate early osteoblast differentiation. Biochem. Biophys. Res. Commun. 509, 435–440 (2019).

247

Pitt, J. M., Kroemer, G. & Zitvogel, L. Extracellular vesicles: masters of intercellular communication and potential clinical interventions. J. Clin. Investig. 126, 1139–1143 (2016).

248

Ma, Q. et al. Osteoclast-derived apoptotic bodies show extended biological effects of parental cell in promoting bone defect healing. Theranostics 10, 6825–6838 (2020).

249

Li, D. et al. Osteoclast-derived exosomal miR-214-3p inhibits osteoblastic bone formation. Nat. Commun. 7, 10872 (2016).

250

Ma, Q. et al. Osteoclast-derived apoptotic bodies couple bone resorption and formation in bone remodeling. Bone Res 9, 5 (2021).

251

Zhao, C. et al. Bidirectional ephrinB2-EphB4 signaling controls bone homeostasis. Cell Metab. 4, 111–121 (2006).

252

Tazaki, Y. et al. RANKL, Ephrin-Eph and Wnt10b are key intercellular communication molecules regulating bone remodeling in autologous transplanted goldfish scales. Comp. Biochem. Physiol. A Mol. Integr. Physiol. 225, 46–58 (2018).

253

Matsuo, K. & Otaki, N. Bone cell interactions through Eph/ephrin: bone modeling, remodeling and associated diseases. Cell Adhes. Migr. 6, 148–156 (2012).

254

Allan, E. H. et al. EphrinB2 regulation by PTH and PTHrP revealed by molecular profiling in differentiating osteoblasts. J. Bone Miner. Res. 23, 1170–1181 (2008).

255

Jacome-Galarza, C. E. et al. Developmental origin, functional maintenance and genetic rescue of osteoclasts. Nature 568, 541–545 (2019).

256

Lassen, N. E. et al. Coupling of bone resorption and formation in real time: new knowledge gained from human haversian BMUs. J. Bone Miner. Res. 32, 1395–1405 (2017).

257

McDonald, M. M. et al. Osteoclasts recycle via osteomorphs during RANKL-stimulated bone resorption. Cell 184, 1330–1347.e13 (2021).

258

Kim, S. W. et al. Intermittent parathyroid hormone administration converts quiescent lining cells to active osteoblasts. J. Bone Miner. Res. 27, 2075–2084 (2012).

259

Makino, A. et al. Frequent administration of abaloparatide shows greater gains in bone anabolic window and bone mineral density in mice: a comparison with teriparatide. Bone 142, 115651 (2021).

260

Zanchetta, M. B. et al. Significant bone loss after stopping long-term denosumab treatment: a post FREEDOM study. Osteoporos. Int. 29, 41–47 (2018).

261

Sølling, A. S., Harsløf, T. & Langdahl, B. Treatment with zoledronate subsequent to denosumab in osteoporosis: a randomized trial. J. Bone Miner. Res. 35, 1858–1870 (2020).

262

Baron, R., Vignery, A. & Van Tran, P. The significance of lacunar erosion without osteoclasts: studies on the reversal phase of the remodeling sequence. Metab. Bone Dis. Relat. Res. 2, 35–40 (1980).

263

Dempster, D. W. et al. Standardized nomenclature, symbols, and units for bone histomorphometry: a 2012 update of the report of the ASBMR histomorphometry nomenclature committee. J. Bone Miner. Res. 28, 2–17 (2013).

264

Yamamoto, T., Yamagata, A. & Nagai, H. A histochemical study of tartrate-resistant acid phosphatase activity in rat osteoblasts. Acta Histochem. Cytochem. 29, 221–225 (1996).

265

Andersen, T. L. et al. Understanding coupling between bone resorption and formation. Am. J. Pathol. 183, 1–12 (2013).

266

Abdelgawad, M. E. et al. Early reversal cells in adult human bone remodeling: osteoblastic nature, catabolic functions and interactions with osteoclasts. Histochem. Cell Biol. 145, 603–615 (2016).

267

Rosenberg, N., Rosenberg, O. & Soudry, M. Osteoblasts in bone physiology-mini review. Rambam Maimonides Med. J. 3, e0013 (2012).

268

Florencio-Silva, R., Sasso, G. R. D. S., Sasso-Cerri, E., Simões, M. J. & Cerri, P. S. Biology of bone tissue: structure, function, and factors that influence bone cells. Biomed. Res. Int. 2015, 421746 (2015).

269

Yellowley, C. E., Li, Z., Zhou, Z., Jacobs, C. R. & Donahue, H. J. Functional gap junctions between osteocytic and osteoblastic cells. J. Bone Miner. Res. 15, 209–217 (2000).

270

Marotti, G., Ferretti, M., Palumbo, C. & Benincasa, M. Static anddynamic bone formation and the mechanism of collagen fiberorientation. Bone 25, 156 (1999).

271

Ferretti, M. & Palumbo, C. The osteocyte: from “prisoner” to “orchestrator”. J. Funct. Morphol. Kinesiol 6, 28 (2021).

272

Atkins, G. J. et al. RANKL expression is related to the differentiation state of human osteoblasts. J. Bone Miner. Res. 18, 1088–1098 (2003).

273

Bianco, P. & Robey, P. G. Skeletal stem cells. Development 142, 1023–1027 (2015).

274

Palumbo, C., Cavani, F., Sena, P., Benincasa, M. & Ferretti, M. Osteocyte apoptosis and absence of bone remodeling in human auditory ossicles and scleral ossicles of lower vertebrates: a mere coincidence or linked processes? Calcif. Tissue Int. 90, 211–218 (2012).

275

Bianco, P. et al. The meaning, the sense and the significance: translating the science of mesenchymal stem cells into medicine. Nat. Med. 19, 35–42 (2013).

276

Sacchetti, B. et al. No identical “mesenchymal stem cells” at different times and sites: human committed progenitors of distinct origin and differentiation potential are incorporated as adventitial cells in microvessels. Stem Cell Rep. 6, 897–913 (2016).

277

Sacchetti, B. Post-natal “mesenchymal” stem cells: the assayable skeletal potency. J. Stem Cells Regen. Med. 15, 12–15 (2019).

278

Kuznetsov, S. A. et al. Circulating skeletal stem cells. J. Cell Biol. 153, 1133–1140 (2001).

279

Alm, J. J. et al. Circulating plastic adherent mesenchymal stem cells in aged hip fracture patients. J. Orthop. Res. 28, 1634–1642 (2010).

280

Canalis, E. The fate of circulating osteoblasts. N. Engl. J. Med 352, 2014–2016 (2005).

281

Feehan, J., Nurgali, K., Apostolopoulos, V., Al Saedi, A. & Duque, G. Circulating osteogenic precursor cells: building bone from blood. EBioMedicine 39, 603–611 (2019).

282

Manolagas, S. C. & Parfitt, A. M. What old means to bone. Trends Endocrinol. Metab. 21, 369–374 (2010).

283

Weinstein, R. S., Jilka, R. L., Parfitt, A. M. & Manolagas, S. C. Inhibition of osteoblastogenesis and promotion of apoptosis of osteoblasts and osteocytes by glucocorticoids. Potential mechanisms of their deleterious effects on bone. J. Clin. Investig. 102, 274–282 (1998).

284

Moriishi, T. et al. Overexpression of BCLXL in osteoblasts inhibits osteoblast apoptosis and increases bone volume and strength. J. Bone Miner. Res. 31, 1366–1380 (2016).

285

Manolagas, S. C., Kousteni, S. & Jilka, R. L. Sex steroids and bone. Recent Prog. Horm. Res. 57, 385–409 (2002).

286

Wang, X. et al. p53 functions as a negative regulator of osteoblastogenesis, osteoblast-dependent osteoclastogenesis, and bone remodeling. J. Cell Biol. 172, 115–125 (2006).

287

Liu, H. & Li, B. p53 control of bone remodeling. J. Cell. Biochem. 111, 529–534 (2010).

288

Bonewald, L. F. The amazing osteocyte. J. Bone Miner. Res. 26, 229–238 (2011).

289

Zhao, W., Byrne, M. H., Wang, Y. & Krane, S. M. Osteocyte and osteoblast apoptosis and excessive bone deposition accompany failure of collagenase cleavage of collagen. J. Clin. Investig. 106, 941–949 (2000).

290

Holmbeck, K. et al. The metalloproteinase MT1-MMP is required for normal development and maintenance of osteocyte processes in bone. J. Cell Sci. 118, 147–156 (2005).

291

Staines, K. A. et al. Hypomorphic conditional deletion of E11/podoplanin reveals a role in osteocyte dendrite elongation. J. Cell. Physiol. 232, 3006–3019 (2017).

292

Wang, J. S. et al. Control of osteocyte dendrite formation by Sp7 and its target gene osteocrin. Nat. Commun. 12, 6271 (2021).

293

Lerebours, C. & Buenzli, P. R. Towards a cell-based mechanostat theory of bone: the need to account for osteocyte desensitisation and osteocyte replacement. J. Biomech. 49, 2600–2606 (2016).

Bone Research
Article number: 48
Cite this article:
Bolamperti S, Villa I, Rubinacci A. Bone remodeling: an operational process ensuring survival and bone mechanical competence. Bone Research, 2022, 10: 48. https://doi.org/10.1038/s41413-022-00219-8

201

Views

5

Downloads

140

Crossref

120

Web of Science

129

Scopus

Altmetrics

Received: 02 August 2021
Revised: 02 May 2022
Accepted: 15 May 2022
Published: 18 July 2022
© The Author(s) 2022

This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons license, and indicate if changes were made. The images or other third party material in this article are included in the article’s Creative Commons license, unless indicated otherwise in a credit line to the material. If material is not included in the article’s Creative Commons license and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this license, visit http://creativecommons.org/licenses/by/4.0/.

Return