Hostname: page-component-848d4c4894-8kt4b Total loading time: 0 Render date: 2024-06-22T10:08:50.578Z Has data issue: false hasContentIssue false

Centrosome Fine Ultrastructure of the Osteocyte Mechanosensitive Primary Cilium

Published online by Cambridge University Press:  21 November 2012

R.E. Uzbekov*
Affiliation:
Department of Microscopy, François Rabelais University, Tours, France Faculty of Bioengineering and Bioinformatics, Moscow State University, Moscow, Russia
D.B. Maurel
Affiliation:
EA4708 I3MTO, Orléans University, IPROS, Hospital Porte Madeleine, Orleans, France
P.C. Aveline
Affiliation:
EA4708 I3MTO, Orléans University, IPROS, Hospital Porte Madeleine, Orleans, France
S. Pallu
Affiliation:
EA4708 I3MTO, Orléans University, IPROS, Hospital Porte Madeleine, Orleans, France
C.L. Benhamou
Affiliation:
EA4708 I3MTO, Orléans University, IPROS, Hospital Porte Madeleine, Orleans, France
G.Y. Rochefort*
Affiliation:
EA4708 I3MTO, Orléans University, IPROS, Hospital Porte Madeleine, Orleans, France
*
*Corresponding author. E-mail: rustem.uzbekov@univ-tours.fr
**Corresponding author. E-mail: gael.rochefort@gmail.com
Get access

Abstract

The centrosome is the principal microtubule organization center in cells, giving rise to microtubule-based organelles (e.g., cilia, flagella). The aim was to study the osteocyte centrosome morphology at an ultrastructural level in relation to its mechanosensitive function. Osteocyte centrosomes and cilia in tibial cortical bone were explored by acetylated alpha-tubulin (AαTub) immunostaining under confocal microscopy. For the first time, fine ultrastructure and spatial orientation of the osteocyte centrosome were explored by transmission electron microscopy on serial ultrathin sections. AαTub-positive staining was observed in 94% of the osteocytes examined (222/236). The mother centriole formed a short primary cilium and was longer than the daughter centriole due to an intermediate zone between centriole and cilium. The proximal end of the mother centriole was connected with the surface of daughter centriole by striated rootlets. The mother centriole exhibited distal appendages that interacted with the cell membrane and formed a particular structure called “cilium membrane prolongation.” The primary cilium was mainly oriented perpendicular to the long axis of bone. Mother and daughter centrioles change their original mutual orientation during the osteocyte differentiation process. The short primary cilium is hypothesized as a novel type of fluid-sensing organelle in osteocytes.

Type
Biological Applications
Copyright
Copyright © Microscopy Society of America 2012

Access options

Get access to the full version of this content by using one of the access options below. (Log in options will check for institutional or personal access. Content may require purchase if you do not have access.)

References

Adams, G.M., Wright, R.L. & Jarvik, J.W. (1985). Defective temporal and spatial control of flagellar assembly in a mutant of Chlamydomonas reinhardtii with variable flagellar number. J Cell Biol 100(3), 955964.Google Scholar
Albrecht-Buehler, G. & Bushnell, A. (1979). The orientation of centrioles in migrating 3T3 cells. Exp Cell Res 120(1), 111118.Google Scholar
Alieva, I.B. & Uzbekov, R.E. (2008). The centrosome is a polyfunctional multiprotein cell complex. Biochemistry (Mosc) 73(6), 626643.Google Scholar
Alieva, I.B. & Vorobjev, I.A. (2004). Vertebrate primary cilia: A sensory part of centrosomal complex in tissue cells, but a “sleeping beauty” in cultured cells? Cell Biol Int 28(2), 139150.Google Scholar
Anderson, R.G. (1972). The three-dimensional structure of the basal body from the rhesus monkey oviduct. J Cell Biol 54(2), 246265.Google Scholar
Baud, C.A. (1968). Submicroscopic structure and functional aspects of the osteocyte. Clin Orthop Relat Res 56, 227236.Google Scholar
Bonewald, L.F. (2006). Mechanosensation and transduction in osteocytes. Bonekey Osteovision 3(10), 715.Google Scholar
Bonewald, L.F. (2007). Osteocytes as dynamic multifunctional cells. Ann NY Acad Sci 1116, 281290.Google Scholar
Bonewald, L.F. (2011). The amazing osteocyte. J Bone Miner Res 26(2), 229238.Google Scholar
Bonewald, L.F. & Johnson, M.L. (2008). Osteocytes, mechanosensing and Wnt signaling. Bone 42(4), 606615.Google Scholar
Burger, E.H., Klein-Nulend, J. & Smit, T.H. (2003). Strain-derived canalicular fluid flow regulates osteoclast activity in a remodelling osteon—A proposal. J Biomech 36(10), 14531459.Google Scholar
Cambray-Deakin, M.A. & Burgoyne, R.D. (1987). Acetylated and detyrosinated alpha-tubulins are co-localized in stable microtubules in rat meningeal fibroblasts. Cell Motil Cytoskeleton 8(3), 284291.Google Scholar
Connolly, J.A., Kiosses, B.W. & Kalnins, V.I. (1986). Centrioles are lost as embryonic myoblasts fuse into myotubes in vitro . Eur J Cell Biol 39(2), 341345.Google Scholar
Corbit, K.C., Shyer, A.E., Dowdle, W.E., Gaulden, J., Singla, V., Chen, M.H., Chuang, P.T. & Reiter, J.F. (2008). Kif3a constrains beta-catenin-dependent Wnt signalling through dual ciliary and non-ciliary mechanisms. Nat Cell Biol 10(1), 7076.Google Scholar
Davenport, J.R. & Yoder, B.K. (2005). An incredible decade for the primary cilium: A look at a once-forgotten organelle. Am J Physiol Renal Physiol 289(6), F1159–1169.Google Scholar
DeRouen, M.C. & Oro, A.E. (2009). The primary cilium: A small yet mighty organelle. J Invest Dermatol 129(2), 264265.Google Scholar
Dixon, W.E. & Inchley, O. (1905). The cilioscribe, an instrument for recording the activity of cilia. J Physiol 32(5-6), 395400.Google Scholar
Federman, M. & Nichols, G. Jr. (1974). Bone cell cilia: Vestigial or functional organelles? Calcif Tissue Res 17(1), 8185.Google Scholar
Gerdes, J.M., Liu, Y., Zaghloul, N.A., Leitch, C.C., Lawson, S.S., Kato, M., Beachy, P.A., Beales, P.L., DeMartino, G.N., Fisher, S., Badano, J.L. & Katsanis, N. (2007). Disruption of the basal body compromises proteasomal function and perturbs intracellular Wnt response. Nat Genet 39(11), 13501360.Google Scholar
Greenwood, M. (1892). On retractile cilia in the intestine of lumbricus terrestris. J Physiol 13(3-4), 239259.Google Scholar
Greer, K., Maruta, H., L'Hernault, S.W. & Rosenbaum, J.L. (1985). Alpha-tubulin acetylase activity in isolated Chlamydomonas flagella . J Cell Biol 101(6), 20812084.Google Scholar
Han, Y., Cowin, S.C., Schaffler, M.B. & Weinbaum, S. (2004). Mechanotransduction and strain amplification in osteocyte cell processes. Proc Natl Acad Sci USA 101(47), 1668916694.Google Scholar
Handel, M., Schulz, S., Stanarius, A., Schreff, M., Erdtmann-Vourliotis, M., Schmidt, H., Wolf, G. & Hollt, V. (1999). Selective targeting of somatostatin receptor 3 to neuronal cilia. Neuroscience 89(3), 909926.Google Scholar
Haycraft, C.J. & Serra, R. (2008). Cilia involvement in patterning and maintenance of the skeleton. Curr Top Dev Biol 85, 303332.Google Scholar
Heino, T.J., Kurata, K., Higaki, H. & Vaananen, H.K. (2009). Evidence for the role of osteocytes in the initiation of targeted remodeling. Technol Health Care 17(1), 4956.Google Scholar
Hibberd, D.J. (1975). Observations on the ultrastructure of the choanoflagellate Codosiga botrytis (Ehr.) Saville-Kent with special reference to the flagellar apparatus. J Cell Sci 17(1), 191219.Google Scholar
Huangfu, D. & Anderson, K.V. (2005). Cilia and hedgehog responsiveness in the mouse. Proc Natl Acad Sci USA 102(32), 1132511330.Google Scholar
Jacobs, C.R., Temiyasathit, S. & Castillo, A.B. (2010). Osteocyte mechanobiology and pericellular mechanics. Annu Rev Biomed Eng 12, 369400.Google Scholar
Kitase, Y., Barragan, L., Qing, H., Kondoh, S., Jiang, J.X., Johnson, M.L. & Bonewald, L.F. (2010). Mechanical induction of PGE2 in osteocytes blocks glucocorticoid-induced apoptosis through both the beta-catenin and PKA pathways. J Bone Miner Res 25(12), 26572668.Google Scholar
Koyama, E., Young, B., Nagayama, M., Shibukawa, Y., Enomoto-Iwamoto, M., Iwamoto, M., Maeda, Y., Lanske, B., Song, B., Serra, R. & Pacifici, M. (2007). Conditional Kif3a ablation causes abnormal hedgehog signaling topography, growth plate dysfunction, and excessive bone and cartilage formation during mouse skeletogenesis. Development 134(11), 21592169.Google Scholar
Kwon, R.Y., Temiyasathit, S., Tummala, P., Quah, C.C. & Jacobs, C.R. (2010). Primary cilium-dependent mechanosensing is mediated by adenylyl cyclase 6 and cyclic AMP in bone cells. FASEB J 24(8), 28592868.Google Scholar
Larson, D.E. & Dingle, A.D. (1981). Isolation, ultrastructure, and protein composition of the flagellar rootlet of Naegleria gruberi . J Cell Biol 89(3), 424432.Google Scholar
L'Hernault, S.W. & Rosenbaum, J.L. (1985a). Chlamydomonas alpha-tubulin is posttranslationally modified by acetylation on the epsilon-amino group of a lysine. Biochemistry 24(2), 473478.Google Scholar
L'Hernault, S.W. & Rosenbaum, J.L. (1985b). Reversal of the posttranslational modification on Chlamydomonas flagellar alpha-tubulin occurs during flagellar resorption. J Cell Biol 100(2), 457462.Google Scholar
Lin, F., Hiesberger, T., Cordes, K., Sinclair, A.M., Goldstein, L.S., Somlo, S. & Igarashi, P. (2003). Kidney-specific inactivation of the KIF3A subunit of kinesin-II inhibits renal ciliogenesis and produces polycystic kidney disease. Proc Natl Acad Sci USA 100(9), 52865291.Google Scholar
Liu, W., Murcia, N.S., Duan, Y., Weinbaum, S., Yoder, B.K., Schwiebert, E. & Satlin, L.M. (2005). Mechanoregulation of intracellular Ca2+ concentration is attenuated in collecting duct of monocilium-impaired orpk mice. Am J Physiol Renal Physiol 289(5), F978–988.Google Scholar
Low, S.H., Vasanth, S., Larson, C.H., Mukherjee, S., Sharma, N., Kinter, M.T., Kane, M.E., Obara, T. & Weimbs, T. (2006). Polycystin-1, STAT6, and P100 function in a pathway that transduces ciliary mechanosensation and is activated in polycystic kidney disease. Dev Cell 10(1), 5769.Google Scholar
Malone, A.M., Anderson, C.T., Stearns, T. & Jacobs, C.R. (2007a). Primary cilia in bone. J Musculoskelet Neuronal Interact 7(4), 301.Google Scholar
Malone, A.M., Anderson, C.T., Tummala, P., Kwon, R.Y., Johnston, T.R., Stearns, T. & Jacobs, C.R. (2007b). Primary cilia mediate mechanosensing in bone cells by a calcium-independent mechanism. Proc Natl Acad Sci USA 104(33), 1332513330.Google Scholar
Maruta, H., Greer, K. & Rosenbaum, J.L. (1986). The acetylation of alpha-tubulin and its relationship to the assembly and disassembly of microtubules. J Cell Biol 103(2), 571579.Google Scholar
Maurel, D.B., Jaffre, C., Rochefort, G.Y., Aveline, P.C., Boisseau, N., Uzbekov, R., Gosset, D., Pichon, C., Fazzalari, N.L., Pallu, S. & Benhamou, C.L. (2011). Low bone accrual is associated with osteocyte apoptosis in alcohol-induced osteopenia. Bone 49(3), 543552.Google Scholar
Mullins, R. & Wette, R. (1966). On the statistical expectation and evaluation of centriole orientations in cell profiles. J Cell Biol 30(3), 652655.Google Scholar
Nauli, S.M., Alenghat, F.J., Luo, Y., Williams, E., Vassilev, P., Li, X., Elia, A.E., Lu, W., Brown, E.M., Quinn, S.J., Ingber, D.E. & Zhou, J. (2003). Polycystins 1 and 2 mediate mechanosensation in the primary cilium of kidney cells. Nat Genet 33(2), 129137.Google Scholar
Nicolella, D.P., Moravits, D.E., Gale, A.M., Bonewald, L.F. & Lankford, J. (2006). Osteocyte lacunae tissue strain in cortical bone. J Biomech 39(9), 17351743.Google Scholar
Noble, B.S. (2008). The osteocyte lineage. Arch Biochem Biophys 473(2), 106111.Google Scholar
Nonaka, S., Yoshiba, S., Watanabe, D., Ikeuchi, S., Goto, T., Marshall, W.F. & Hamada, H. (2005). De novo formation of left-right asymmetry by posterior tilt of nodal cilia. PLoS Biol 3(8), e268. Google Scholar
Olsen, B. (2005). Nearly all cells in vertebrates and many cells in invertebrates contain primary cilia. Matrix Biol 24(7), 449450.Google Scholar
Ong, A.C. & Wheatley, D.N. (2003). Polycystic kidney disease—The ciliary connection. Lancet 361(9359), 774776.Google Scholar
Pacheco, M., Valencia, M., Caparros-Martin, J.A., Mulero, F., Goodship, J.A. & Ruiz-Perez, V.L. (2012). Evc works in chondrocytes and osteoblasts to regulate multiple aspects of growth plate development in the appendicular skeleton and cranial base. Bone 50(1), 2841.Google Scholar
Parfitt, A.M. (1977). The cellular basis of bone turnover and bone loss: A rebuttal of the osteocytic resorption—Bone flow theory. Clin Orthop Relat Res 127, 236247.Google Scholar
Pazour, G.J. (2004). Intraflagellar transport and cilia-dependent renal disease: The ciliary hypothesis of polycystic kidney disease. J Am Soc Nephrol 15(10), 25282536.Google Scholar
Piperno, G. & Fuller, M.T. (1985). Monoclonal antibodies specific for an acetylated form of alpha-tubulin recognize the antigen in cilia and flagella from a variety of organisms. J Cell Biol 101(6), 20852094.Google Scholar
Piperno, G., LeDizet, M. & Chang, X.J. (1987). Microtubules containing acetylated alpha-tubulin in mammalian cells in culture. J Cell Biol 104(2), 289302.Google Scholar
Plotkin, L.I., Lezcano, V., Thostenson, J., Weinstein, R.S., Manolagas, S.C. & Bellido, T. (2008). Connexin 43 is required for the anti-apoptotic effect of bisphosphonates on osteocytes and osteoblasts in vivo . J Bone Miner Res 23(11), 17121721.Google Scholar
Plotkin, L.I., Manolagas, S.C. & Bellido, T. (2007). Glucocorticoids induce osteocyte apoptosis by blocking focal adhesion kinase-mediated survival. Evidence for inside-out signaling leading to anoikis. J Biol Chem 282(33), 2412024130.Google Scholar
Plotnikova, O.V., Golemis, E.A. & Pugacheva, E.N. (2008). Cell cycle-dependent ciliogenesis and cancer. Cancer Res 68(7), 20582061.Google Scholar
Poole, C.A., Flint, M.H. & Beaumont, B.W. (1985). Analysis of the morphology and function of primary cilia in connective tissues: A cellular cybernetic probe? Cell Motil 5(3), 175193.Google Scholar
Poole, C.A., Zhang, Z.J. & Ross, J.M. (2001). The differential distribution of acetylated and detyrosinated alpha-tubulin in the microtubular cytoskeleton and primary cilia of hyaline cartilage chondrocytes. J Anat 199(Pt 4), 393405.Google Scholar
Praetorius, H.A. & Spring, K.R. (2001). Bending the MDCK cell primary cilium increases intracellular calcium. J Membr Biol 184(1), 7179.Google Scholar
Qiu, N., Cao, L., David, V., Quarles, L.D. & Xiao, Z. (2010). Kif3a deficiency reverses the skeletal abnormalities in Pkd1 deficient mice by restoring the balance between osteogenesis and adipogenesis. PLoS One 5(12), e15240. Google Scholar
Rattner, J.B. & Phillips, S.G. (1973). Independence of centriole formation and DNA synthesis. J Cell Biol 57(2), 359372.Google Scholar
Robbins, E., Jentzsch, G. & Micali, A. (1968). The centriole cycle in synchronized HeLa cells. J Cell Biol 36(2), 329339.Google Scholar
Rochefort, G.Y., Pallu, S. & Benhamou, C.L. (2010). Osteocyte: The unrecognized side of bone tissue. Osteoporos Int 21(9), 14571469.Google Scholar
Rosenbaum, J.L. & Witman, G.B. (2002). Intraflagellar transport. Nat Rev Mol Cell Biol 3(11), 813825.Google Scholar
Santos, A., Bakker, A.D., Zandieh-Doulabi, B., Semeins, C.M. & Klein-Nulend, J. (2009). Pulsating fluid flow modulates gene expression of proteins involved in Wnt signaling pathways in osteocytes. J Orthop Res 27(10), 12801287.Google Scholar
Sasse, R., Glyn, M.C., Birkett, C.R. & Gull, K. (1987). Acetylated alpha-tubulin in Physarum: immunological characterization of the isotype and its usage in particular microtubular organelles. J Cell Biol 104(1), 4149.Google Scholar
Scherft, J.P. & Daems, W.T. (1967). Single cilia in chondrocytes. J Ultrastruct Res 19(5), 546555.Google Scholar
Schneider, L., Clement, C.A., Teilmann, S.C., Pazour, G.J., Hoffmann, E.K., Satir, P. & Christensen, S.T. (2005). PDGFRalphaalpha signaling is regulated through the primary cilium in fibroblasts. Curr Biol 15(20), 18611866.Google Scholar
Scholey, J.M. (2003). Intraflagellar transport. Annu Rev Cell Dev Biol 19, 423443.Google Scholar
Simons, M., Gloy, J., Ganner, A., Bullerkotte, A., Bashkurov, M., Kronig, C., Schermer, B., Benzing, T., Cabello, O.A., Jenny, A., Mlodzik, M., Polok, B., Driever, W., Obara, T. & Walz, G. (2005). Inversin, the gene product mutated in nephronophthisis type II, functions as a molecular switch between Wnt signaling pathways. Nat Genet 37(5), 537543.Google Scholar
Singla, V. & Reiter, J.F. (2006). The primary cilium as the cell's antenna: Signaling at a sensory organelle. Science 313(5787), 629633.Google Scholar
Siroky, B.J., Ferguson, W.B., Fuson, A.L., Xie, Y., Fintha, A., Komlosi, P., Yoder, B.K., Schwiebert, E.M., Guay-Woodford, L.M. & Bell, P.D. (2006). Loss of primary cilia results in deregulated and unabated apical calcium entry in ARPKD collecting duct cells. Am J Physiol Renal Physiol 290(6), F1320–1328.Google Scholar
Tassin, A.M., Maro, B. & Bornens, M. (1985). Fate of microtubule-organizing centers during myogenesis in vitro . J Cell Biol 100(1), 3546.Google Scholar
Temiyasathit, S. & Jacobs, C.R. (2010). Osteocyte primary cilium and its role in bone mechanotransduction. Ann NY Acad Sci 1192, 422428.Google Scholar
Teti, A. & Zallone, A. (2009). Do osteocytes contribute to bone mineral homeostasis? Osteocytic osteolysis revisited. Bone 44(1), 1116.Google Scholar
Uzbekov, R. & Prigent, C. (2007). Clockwise or anticlockwise? Turning the centriole triplets in the right direction! FEBS Lett 581(7), 12511254.Google Scholar
Uzbekov, R.E. & Alieva, I.B. (2008). The centrosome—A riddle of the “cell processor.” Tsitologiia 50(2), 91112 (in Russian).Google Scholar
Vasil'ev, N.B., Vorob'ev, I.A., Leontovich, A.M. & Petrovskaia, M.B. (1988). An analysis of the centriole orientation in tissue culture cells. Tsitologiia 30(9), 10911100 (in Russian).Google Scholar
Vorobjev, I.A. & Chentsov, Yu. S. (1982). Centrioles in the cell cycle. I. Epithelial cells. J Cell Biol 93(3), 938949.Google Scholar
Whitfield, J.F. (2003). Primary cilium—Is it an osteocyte's strain-sensing flowmeter? J Cell Biochem 89(2), 233237.Google Scholar
Xiao, Z., Zhang, S., Mahlios, J., Zhou, G., Magenheimer, B.S., Guo, D., Dallas, S.L., Maser, R., Calvet, J.P., Bonewald, L. & Quarles, L.D. (2006). Cilia-like structures and polycystin-1 in osteoblasts/osteocytes and associated abnormalities in skeletogenesis and Runx2 expression. J Biol Chem 281(41), 3088430895.Google Scholar
Xiao, Z.S. & Quarles, L.D. (2010). Role of the polycytin-primary cilia complex in bone development and mechanosensing. Ann NY Acad Sci 1192, 410421.Google Scholar
You, L.D., Weinbaum, S., Cowin, S.C. & Schaffler, M.B. (2004). Ultrastructure of the osteocyte process and its pericellular matrix. Anat Rec 278(2), 505513.Google Scholar
Supplementary material: PDF

Rustem Uzbekov Supplementary Material

Appendix

Download Rustem Uzbekov Supplementary Material(PDF)
PDF 570.6 KB