决定骨骼发育的转录因子的转录调控机制的研究
详细信息    本馆镜像全文|  推荐本文 |  |   获取CNKI官网全文
摘要
骨是一个极其复杂、不断变化的组织,它的生长过程涉及诸多因素。转录因子是影响骨生长非常重要的因素。Runx2和Osterix是成骨细胞分化和功能及骨形成的两个重要的转录因子。在成骨细胞分化途径中,Osterix处于Runx2下游,但Osterix是否受Runx2的直接调控目前仍不清楚。
     通过RT-PCR技术,我们发现在许多非成骨细胞系(包括已分化细胞和间质细胞)和成骨细胞系中瞬时表达Runx2能够诱导Osterix的表达。我们克隆了3.2kb人的Osterix基因5’侧翼区,并将启动子连接到报告基因上进行功能分析。结果显示Runx2能够上调该启动子的活性,并且进一步的实验又证明了在转录起始位点上游1.1kb和0.5kb之间的区域是维持启动子基本活性所必需,同时该区间也同Runx2的上调活性密切相关。在此区间内我们确定了一个Runx2的功能性结合位点“AGTGGTT”,此外还有另一个弱调节位点“TGTGGT”。以上实验数据支持Runx2参与调控Osterix基因的表达的假说。此外,在非成骨细胞中,瞬时转染和荧光素酶双报告实验显示Osterix能上调I-胶原蛋白的2.3kb启动子活性,但Runx2不能。这点差别暗示了在成骨细胞分化中处于Runx2下游的Osterix是激活I-胶原蛋白表达所必需的。
     Sox9、L-Sox5和Sox6这三种转录因子是软骨细胞分化和软骨形成所必需的。它们共同作用足以诱导永久性软骨的形成。在软骨细胞分化过程中,这三种转录因子更为精细的调控机制还不为所知。我们利用反转录病毒系统建立了以ATDC5细胞为基础的,过表达L-Sox5、Sox6、Sox9和L-Sox5+Sox9的稳转细胞系,并通过Alcian blue、RT-PCR等方法检测软骨细胞分化程度。用抑制剂来确定与Stat1有关的各个稳转细胞的增殖情况以及MAPK所调节的分化情况。结果表明过表达L-Sox5和Sox6能加快软骨细胞增殖,抑制分化;过表达Sox9显著抑制增殖,但却能提高软骨特异性基因的表达。如果外源L-Sox5与Sox9的比例接1:1,便能使增殖和分化在很大程度上接近正常。此外Sox转录因子表达状态的变化能够调控Stat1的表达变化,也使MAPK途径产生不同应答。这些实验结果说明Sox转录因子之间的比例变化能够调控软骨细胞的增殖和分化。在软骨中,胞外信号之所以造就了不同类型的软骨细胞(增殖或分化)至少部分是通过建立不同的Sox比例实现的。
The skeleton is a very complicated and continuously changing tissue. Multiple factors influence skeletal development. Among these factors, transcription factors should play a central regulatory role. Runx2 and Osterix are both key transcription factors involved in osteoblast differentiation and bone formation. However, whether Runx2 positively regulats Osterix is not known.
     By using RT-PCR, we found that transient expression of Runx2 in a number of non-osteoblastic cell lines, either pluripotent or differentiated, and in the osteoblastic cell lines was able to induce the expression of Osterix. We cloned the 3.2 kb 5’-flanking region of human Osterix gene and analyzed the promoter linked to a firefly luciferase reporter gene. We found that Runx2 could upregulate the promoter activity. The fragment between 1.1 kb and 0.5 kb upstream of the first translation start codon was needed for basal promoter activity. A main functional Runx2 binding site“AGTGGTT”was then identified within the promoter, whereas, another Runx2 binding site“TGTGGT”was canonical, but not conserved. Moreover, the transient transfection and dual-luciferase assay showed Osterix up-regulated the activity of the 2.3kb type I collagen promoter in the non-osteoblastic cells, but Runx2 did not. This defference implies that Osterix, the down stream transcription factor of Runx2 during osteoblast differentiation, is needed to stimulate the osteoblast-specific gene expression of type I collagen.
     The three Sox transcription factors, L-Sox5, Sox6, and Sox9, are essentially required for chondrocyte differentiation and cartilage development. They are sufficient to induce permanent cartilage. However, their precise mode of action is still poorly understood when they function together in chondrocytes. By using a retroviral system, we generate several stable cell lines overexpressing L-Sox5, Sox6, Sox9, and L-Sox5 and Sox9, respectively. Chondrogenic differentiation was examined by Alcian Blue staining, RT-PCR etc. Specific inhibitors were used to characterize the Stat1-related proliferation or the MAPK-mediated differentiation of the cell lines. The aberrant overexpression of L-Sox5 or Sox6 in ATDC5 dramatically accelerated proliferation and inhibited differentiation during chondrogenesis. In contrast, the aberrant overexpression of Sox9 markedly inhibited proliferation but enhanced the expression of cartilage-specific genes. Exogenous L-Sox5 and Sox9 in the ratio near 1:1 largely secured normal proliferation and differentiation. In addition, the altered expression statuses of the Sox transcription factors differently regulated Stat1 expression. And these alterations also resulted in different responses to the MAPK pathway. The alterations in the starting ratio of the Sox transcription factors modulate proliferation and differentiation during chondrogenesis. Extracellular signals generate different chondrocytes at least in part through establishing varied ratios of the Sox transcription factors in the cartilage.
引文
[1]. Lefebvre V, Smits P. Transcriptional control of chondrocyte fate and differentiation. Birth Defects Res C Embryo Today. 2005;75:200-212.
    [2]. Ito Y. Molecular basis of tissue-specific gene expression mediated by the runt domain transcription factor PEBP2/CBF. Genes Cells. 1999;4:685-696.
    [3]. Schroeder TM, Jensen ED, Westendorf JJ. Runx2: a master organizer of gene transcription in developing and maturing osteoblasts. Birth Defects Res C Embryo Today. 2005;75:213-225.
    [4]. Okuda T, van Deursen J, Hiebert SW, et al. AML1, the target of multiple chromosomal translocations in human leukemia, is essential for normal fetal liver hematopoiesis. Cell. 1996;84:321-330.
    [5]. Li QL, Ito K, Sakakura C, et al. Causal relationship between the loss of RUNX3 expression and gastric cancer. Cell. 2002;109:113-124.
    [6]. Otto F, Thornell AP, Crompton T, et al. Cbfa1, a candidate gene for cleidocranial dysplasia syndrome, is essential for osteoblast differentiation and bone development. Cell. 1997;89:765-771.
    [7]. Inada M, Yasui T, Nomura S, et al. Maturational disturbance of chondrocytes in Cbfa1-deficient mice. Dev Dyn. 1999;214:279-290.
    [8]. Ducy P, Zhang R, Geoffroy V, et al. Osf2/Cbfa1: a transcriptional activator of osteoblast differentiation. Cell. 1997;89:747-754.
    [9]. Mundlos S, Otto F, Mundlos C, et al. Mutations involving the transcription factor CBFA1 cause cleidocranial dysplasia. Cell. 1997;89:773-779.
    [10]. Komori T, Yagi H, Nomura S, et al. Targeted disruption of Cbfa1 results in a complete lack of bone formation owing to maturational arrest of osteoblasts. Cell. 1997;89:755-764.
    [11]. Yoshida CA, Furuichi T, Fujita T, et al. Core-binding factor beta interacts with Runx2 and is required for skeletal development. Nat Genet. 2002;32:633-638.
    [12]. Kundu M, Javed A, Jeon JP, et al. Cbfbeta interacts with Runx2 and has a critical role in bone development. Nat Genet. 2002;32:639-644.
    [13]. Miller J, Horner A, Stacy T, et al. The core-binding factor beta subunit is required for bone formation and hematopoietic maturation. Nat Genet. 2002;32:645-649.
    [14]. Takeda S, Bonnamy JP, Owen MJ, et al. Continuous expression of Cbfa1 in nonhypertrophic chondrocytes uncovers its ability to induce hypertrophic chondrocyte differentiation and partially rescues Cbfa1-deficient mice. Genes Dev. 2001;15:467-481.
    [15]. Ueta C, Iwamoto M, Kanatani N, et al. Skeletal malformations caused by overexpression of Cbfa1 or its dominant negative form in chondrocytes. J Cell Biol. 2001;153:87-100.
    [16]. Liu W, Toyosawa S, Furuichi T, et al. Overexpression of Cbfa1 in osteoblasts inhibits osteoblast maturation and causes osteopenia with multiple fractures. J Cell Biol. 2001;155:157-166.
    [17]. Geoffroy V, Kneissel M, Fournier B, et al. High bone resorption in adult aging transgenic mice overexpressing cbfa1/runx2 in cells of the osteoblastic lineage. Mol Cell Biol. 2002;22:6222-6233.
    [18]. Choi JY, Pratap J, Javed A, et al. Subnuclear targeting of Runx/Cbfa/AML factors is essential for tissue-specific differentiation during embryonic development. Proc Natl Acad Sci U S A. 2001;98:8650-8655.
    [19]. Zhang YW, Yasui N, Ito K, et al. A RUNX2/PEBP2alpha A/CBFA1 mutation displaying impaired transactivation and Smad interaction in cleidocranial dysplasia. Proc Natl Acad Sci U S A. 2000;97:10549-10554.
    [20]. Xiao G, Jiang D, Gopalakrishnan R, et al. Fibroblast growth factor 2 induction of the osteocalcin gene requires MAPK activity and phosphorylation of the osteoblast transcription factor, Cbfa1/Runx2. JBiol Chem. 2002;277:36181-36187.
    [21]. Tintut Y, Parhami F, Le V, et al. Inhibition of osteoblast-specific transcription factor Cbfa1 by the cAMP pathway in osteoblastic cells. Ubiquitin/proteasome-dependent regulation. J Biol Chem. 1999;274:28875-28879.
    [22]. Wee HJ, Huang G, Shigesada K, et al. Serine phosphorylation of RUNX2 with novel potential functions as negative regulatory mechanisms. EMBO Rep. 2002;3:967-974.
    [23]. Thirunavukkarasu K, Mahajan M, McLarren KW, et al. Two domains unique to osteoblast-specific transcription factor Osf2/Cbfa1 contribute to its transactivation function and its inability to heterodimerize with Cbfbeta. Mol Cell Biol. 1998;18:4197-4208.
    [24]. McLarren KW, Lo R, Grbavec D, et al. The mammalian basic helix loop helix protein HES-1 binds to and modulates the transactivating function of the runt-related factor Cbfa1. J Biol Chem. 2000;275:530-538.
    [25]. Thomas DM, Carty SA, Piscopo DM, et al. The retinoblastoma protein acts as a transcriptional coactivator required for osteogenic differentiation. Mol Cell. 2001;8:303-316.
    [26]. Nakashima K, Zhou X, Kunkel G, et al. The novel zinc finger-containing transcription factor osterix is required for osteoblast differentiation and bone formation. Cell. 2002;108:17-29.
    [27]. Zelzer E, Glotzer DJ, Hartmann C, et al. Tissue specific regulation of VEGF expression during bone development requires Cbfa1/Runx2. Mech Dev. 2001;106:97-106.
    [28]. Pevny LH, Lovell-Badge R. Sox genes find their feet. Curr Opin Genet Dev. 1997;7:338-344.
    [29]. Wegner M. From head to toes: the multiple facets of Sox proteins. Nucleic Acids Res. 1999;27:1409-1420.
    [30]. Kamachi Y, Uchikawa M, Kondoh H. Pairing SOX off: with partners in the regulation of embryonic development. Trends Genet. 2000;16:182-187.
    [31]. Kamachi Y, Cheah KS, Kondoh H. Mechanism of regulatory target selection by the SOX high-mobility-group domain proteins as revealed by comparison of SOX1/2/3 and SOX9. Mol Cell Biol. 1999;19:107-120.
    [32]. Lefebvre V, Huang W, Harley VR, et al. SOX9 is a potent activator of the chondrocyte-specific enhancer of the pro alpha1(II) collagen gene. Mol Cell Biol. 1997;17:2336-2346.
    [33]. Kamachi Y, Sockanathan S, Liu Q, et al. Involvement of SOX proteins in lens-specific activation of crystallin genes. Embo J. 1995;14:3510-3519.
    [34]. Yuan H, Corbi N, Basilico C, et al. Developmental-specific activity of the FGF-4 enhancer requires the synergistic action of Sox2 and Oct-3. Genes Dev. 1995;9:2635-2645.
    [35]. Nishimoto M, Fukushima A, Okuda A, et al. The gene for the embryonic stem cell coactivator UTF1 carries a regulatory element which selectively interacts with a complex composed of Oct-3/4 and Sox-2. Mol Cell Biol. 1999;19:5453-5465.
    [36]. De Santa Barbara P, Bonneaud N, Boizet B, et al. Direct interaction of SRY-related protein SOX9 and steroidogenic factor 1 regulates transcription of the human anti-Mullerian hormone gene. Mol Cell Biol. 1998;18:6653-6665.
    [37]. Foster JW, Dominguez-Steglich MA, Guioli S, et al. Campomelic dysplasia and autosomal sex reversal caused by mutations in an SRY-related gene. Nature. 1994;372:525-530.
    [38]. Wagner T, Wirth J, Meyer J, et al. Autosomal sex reversal and campomelic dysplasia are caused by mutations in and around the SRY-related gene SOX9. Cell. 1994;79:1111-1120.
    [39]. Sudbeck P, Schmitz ML, Baeuerle PA, et al. Sex reversal by loss of the C-terminal transactivation domain of human SOX9. Nat Genet. 1996;13:230-232.
    [40]. Wright E, Hargrave MR, Christiansen J, et al. The Sry-related gene Sox9 is expressed during chondrogenesis in mouse embryos. Nat Genet. 1995;9:15-20.
    [41]. Ng LJ, Wheatley S, Muscat GE, et al. SOX9 binds DNA, activates transcription, and coexpresses with type II collagen during chondrogenesis in the mouse. Dev Biol. 1997;183:108-121.
    [42]. Zhao Q, Eberspaecher H, Lefebvre V, et al. Parallel expression of Sox9 and Col2a1 in cells undergoing chondrogenesis. Dev Dyn. 1997;209:377-386.
    [43]. Bi W, Deng JM, Zhang Z, et al. Sox9 is required for cartilage formation. Nat Genet. 1999;22:85-89.
    [44]. Akiyama H, Chaboissier MC, Martin JF, et al. The transcription factor Sox9 has essential roles in successive steps of the chondrocyte differentiation pathway and is required for expression of Sox5 and Sox6. Genes Dev. 2002;16:2813-2828.
    [45]. Mundlos S, Olsen BR. Heritable diseases of the skeleton. Part II: Molecular insights into skeletal development-matrix components and their homeostasis. Faseb J. 1997;11:227-233.
    [46]. Sandell LJ. In situ expression of collagen and proteoglycan genes in notochord and during skeletal development and growth. Microsc Res Tech. 1994;28:470-482.
    [47]. Lefebvre V, Li P, de Crombrugghe B. A new long form of Sox5 (L-Sox5), Sox6 and Sox9 are coexpressed in chondrogenesis and cooperatively activate the type II collagen gene. Embo J. 1998;17:5718-5733.
    [48]. Smits P, Li P, Mandel J, et al. The transcription factors L-Sox5 and Sox6 are essential for cartilage formation. Dev Cell. 2001;1:277-290.
    [49]. Brent AE, Braun T, Tabin CJ. Genetic analysis of interactions between the somitic muscle, cartilage and tendon cell lineages during mouse development. Development. 2005;132:515-528.
    [50]. Lefebvre V. Toward understanding the functions of the two highly related Sox5 and Sox6 genes. J Bone Miner Metab. 2002;20:121-130.
    [51]. Bell DM, Leung KK, Wheatley SC, et al. SOX9 directly regulates the type-II collagen gene. Nat Genet. 1997;16:174-178.
    [52]. Xie WF, Zhang X, Sakano S, et al. Trans-activation of the mouse cartilage-derived retinoic acid-sensitive protein gene by Sox9. J Bone Miner Res. 1999;14:757-763.
    [53]. Bridgewater LC, Lefebvre V, de Crombrugghe B. Chondrocyte-specific enhancer elements in the Col11a2 gene resemble the Col2a1 tissue-specific enhancer. J Biol Chem. 1998;273:14998-15006.
    [54]. Liu Y, Li H, Tanaka K, et al. Identification of an enhancer sequence within the first intron required for cartilage-specific transcription of the alpha2(XI) collagen gene. J Biol Chem. 2000;275:12712-12718.
    [55]. Sekiya I, Tsuji K, Koopman P, et al. SOX9 enhances aggrecan gene promoter/enhancer activity and is up-regulated by retinoic acid in a cartilage-derived cell line, TC6. J Biol Chem. 2000;275:10738-10744.
    [56]. Rentsendorj O, Nagy A, Sinko I, et al. Highly conserved proximal promoter element harbouring paired Sox9-binding sites contributes to the tissue- and developmental stage-specific activity of the matrilin-1 gene. Biochem J. 2005;389:705-716.
    [57]. Smits P, Dy P, Mitra S, et al. Sox5 and Sox6 are needed to develop and maintain source, columnar, and hypertrophic chondrocytes in the cartilage growth plate. J Cell Biol. 2004;164:747-758.
    [58]. Kobayashi T, Soegiarto DW, Yang Y, et al. Indian hedgehog stimulates periarticular chondrocyte differentiation to regulate growth plate length independently of PTHrP. J Clin Invest. 2005;115:1734-1742.
    [59]. Reimold AM, Grusby MJ, Kosaras B, et al. Chondrodysplasia and neurological abnormalities in ATF-2-deficient mice. Nature. 1996;379:262-265.
    [60]. Kronenberg HM. Developmental regulation of the growth plate. Nature. 2003;423:332-336.
    [61]. St-Jacques B, Hammerschmidt M, McMahon AP. Indian hedgehog signaling regulates proliferation and differentiation of chondrocytes and is essential for bone formation. Genes Dev. 1999;13:2072-2086.
    [62]. Akiyama H, Lyons JP, Mori-Akiyama Y, et al. Interactions between Sox9 and beta-catenin control chondrocyte differentiation. Genes Dev. 2004;18:1072-1087.
    [63]. Kim IS, Otto F, Zabel B, et al. Regulation of chondrocyte differentiation by Cbfa1. Mech Dev. 1999;80:159-170.
    [64]. Yoshida CA, Yamamoto H, Fujita T, et al. Runx2 and Runx3 are essential for chondrocyte maturation, and Runx2 regulates limb growth through induction of Indian hedgehog. Genes Dev. 2004;18:952-963.
    [65]. Huang W, Chung UI, Kronenberg HM, et al. The chondrogenic transcription factor Sox9 is a target of signaling by the parathyroid hormone-related peptide in the growth plate of endochondral bones. Proc Natl Acad Sci U S A. 2001;98:160-165.
    [66]. Seki K, Fujimori T, Savagner P, et al. Mouse Snail family transcription repressors regulate chondrocyte, extracellular matrix, type II collagen, and aggrecan. J Biol Chem. 2003;278:41862-41870.
    [67]. Tanaka K, Tsumaki N, Kozak CA, et al. A Kruppel-associated box-zinc finger protein, NT2, represses cell-type-specific promoter activity of the alpha 2(XI) collagen gene. Mol Cell Biol. 2002;22:4256-4267.
    [68]. Zheng Q, Zhou G, Morello R, et al. Type X collagen gene regulation by Runx2 contributes directly to its hypertrophic chondrocyte-specific expression in vivo. J Cell Biol. 2003;162:833-842.
    [69]. Iwamoto M, Kitagaki J, Tamamura Y, et al. Runx2 expression and action in chondrocytes are regulated by retinoid signaling and parathyroid hormone-related peptide (PTHrP). Osteoarthritis Cartilage. 2003;11:6-15.
    [70]. Li TF, Dong Y, Ionescu AM, et al. Parathyroid hormone-related peptide (PTHrP) inhibits Runx2 expression through the PKA signaling pathway. Exp Cell Res. 2004;299:128-136.
    [71]. Ferrari D, Kosher RA. Dlx5 is a positive regulator of chondrocyte differentiation during endochondral ossification. Dev Biol. 2002;252:257-270.
    [72]. Robledo RF, Rajan L, Li X, et al. The Dlx5 and Dlx6 homeobox genes are essential for craniofacial, axial, and appendicular skeletal development. Genes Dev. 2002;16:1089-1101.
    [73]. Magee C, Nurminskaya M, Faverman L, et al. SP3/SP1 transcription activity regulates specific expression of collagen type X in hypertrophic chondrocytes. J Biol Chem. 2005;280:25331-25338.
    [74]. Roca H, Phimphilai M, Gopalakrishnan R, et al. Cooperative interactions between RUNX2 and homeodomain protein-binding sites are critical for the osteoblast-specific expression of the bone sialoprotein gene. J Biol Chem. 2005;280:30845-30855.
    [75]. Vega RB, Matsuda K, Oh J, et al. Histone deacetylase 4 controls chondrocyte hypertrophy during skeletogenesis. Cell. 2004;119:555-566.
    [76]. Adams CS, Shapiro IM. The fate of the terminally differentiated chondrocyte: evidence for microenvironmental regulation of chondrocyte apoptosis. Crit Rev Oral Biol Med. 2002;13:465-473.
    [77]. Sato M, Morii E, Komori T, et al. Transcriptional regulation of osteopontin gene in vivo by PEBP2alphaA/CBFA1 and ETS1 in the skeletal tissues. Oncogene. 1998;17:1517-1525.
    [78]. Hess J, Porte D, Munz C, et al. AP-1 and Cbfa/runt physically interact and regulate parathyroid hormone-dependent MMP13 expression in osteoblasts through a new osteoblast-specific element 2/AP-1 composite element. J Biol Chem. 2001;276:20029-20038.
    [79]. Kuboki T, Kanyama M, Nakanishi T, et al. Cbfa1/Runx2 gene expression in articular chondrocytes of the mice temporomandibular and knee joints in vivo. Arch Oral Biol. 2003;48:519-525.
    [80]. Wang X, Manner PA, Horner A, et al. Regulation of MMP-13 expression by RUNX2 and FGF2 in osteoarthritic cartilage. Osteoarthritis Cartilage. 2004;12:963-973.
    [81]. Tondravi MM, McKercher SR, Anderson K, et al. Osteopetrosis in mice lacking haematopoietic transcription factor PU.1. Nature. 1997;386:81-84.
    [82]. Yoshida H, Hayashi S, Kunisada T, et al. The murine mutation osteopetrosis is in the coding region of the macrophage colony stimulating factor gene. Nature. 1990;345:442-444.
    [83]. Hershey CL, Fisher DE. Mitf and Tfe3: members of a b-HLH-ZIP transcription factor family essential for osteoclast development and function. Bone. 2004;34:689-696.
    [84]. Steingrimsson E, Tessarollo L, Pathak B, et al. Mitf and Tfe3, two members of the Mitf-Tfe family ofbHLH-Zip transcription factors, have important but functionally redundant roles in osteoclast development. Proc Natl Acad Sci U S A. 2002;99:4477-4482.
    [85]. Partington GA, Fuller K, Chambers TJ, et al. Mitf-PU.1 interactions with the tartrate-resistant acid phosphatase gene promoter during osteoclast differentiation. Bone. 2004;34:237-245.
    [86]. Luchin A, Purdom G, Murphy K, et al. The microphthalmia transcription factor regulates expression of the tartrate-resistant acid phosphatase gene during terminal differentiation of osteoclasts. J Bone Miner Res. 2000;15:451-460.
    [87]. Motyckova G, Weilbaecher KN, Horstmann M, et al. Linking osteopetrosis and pycnodysostosis: regulation of cathepsin K expression by the microphthalmia transcription factor family. Proc Natl Acad Sci U S A. 2001;98:5798-5803.
    [88]. So H, Rho J, Jeong D, et al. Microphthalmia transcription factor and PU.1 synergistically induce the leukocyte receptor osteoclast-associated receptor gene expression. J Biol Chem. 2003;278:24209-24216.
    [89]. Sato M, Morii E, Takebayashi-Suzuki K, et al. Microphthalmia-associated transcription factor interacts with PU.1 and c-Fos: determination of their subcellular localization. Biochem Biophys Res Commun. 1999;254:384-387.
    [90]. Franzoso G, Carlson L, Xing L, et al. Requirement for NF-kappaB in osteoclast and B-cell development. Genes Dev. 1997;11:3482-3496.
    [91]. Johnson RS, Spiegelman BM, Papaioannou V. Pleiotropic effects of a null mutation in the c-fos proto-oncogene. Cell. 1992;71:577-586.
    [92]. Wang ZQ, Ovitt C, Grigoriadis AE, et al. Bone and haematopoietic defects in mice lacking c-fos. Nature. 1992;360:741-745.
    [93]. Rao A, Luo C, Hogan PG. Transcription factors of the NFAT family: regulation and function. Annu Rev Immunol. 1997;15:707-747.
    [94]. Matsuo K, Galson DL, Zhao C, et al. Nuclear factor of activated T-cells (NFAT) rescues osteoclastogenesis in precursors lacking c-Fos. J Biol Chem. 2004;279:26475-26480.
    [95]. Cappellen D, Luong-Nguyen NH, Bongiovanni S, et al. Transcriptional program of mouse osteoclast differentiation governed by the macrophage colony-stimulating factor and the ligand for the receptor activator of NFkappa B. J Biol Chem. 2002;277:21971-21982.
    [96]. Ishida N, Hayashi K, Hoshijima M, et al. Large scale gene expression analysis of osteoclastogenesis in vitro and elucidation of NFAT2 as a key regulator. J Biol Chem. 2002;277:41147-41156.
    [97]. Takayanagi H, Kim S, Koga T, et al. Induction and activation of the transcription factor NFATc1 (NFAT2) integrate RANKL signaling in terminal differentiation of osteoclasts. Dev Cell. 2002;3:889-901.
    [98]. Kobayashi T, Kronenberg H. Minireview: transcriptional regulation in development of bone. Endocrinology. 2005;146:1012-1017.
    [99]. Provot S, Schipani E. Molecular mechanisms of endochondral bone development. Biochem Biophys Res Commun. 2005;328:658-665.
    [100]. Fan Q, Ding J, Zhang J, et al. Effect of the knockdown of podocin mRNA on nephrin and alpha-actinin in mouse podocyte. Exp Biol Med (Maywood). 2004;229:964-970.
    [101]. Altaf FM, Hering TM, Kazmi NH, et al. Ascorbate-enhanced chondrogenesis of ATDC5 cells. Eur Cell Mater. 2006;12:64-69; discussion 69-70.
    [102]. Shukunami C, Shigeno C, Atsumi T, et al. Chondrogenic differentiation of clonal mouse embryonic cell line ATDC5 in vitro: differentiation-dependent gene expression of parathyroid hormone (PTH)/PTH-related peptide receptor. J Cell Biol. 1996;133:457-468.
    [103]. Kitamura T, Koshino Y, Shibata F, et al. Retrovirus-mediated gene transfer and expression cloning: powerful tools in functional genomics. Exp Hematol. 2003;31:1007-1014.
    [104]. Milona MA, Gough JE, Edgar AJ. Expression of alternatively spliced isoforms of human Sp7 inosteoblast-like cells. BMC Genomics. 2003;4:43.
    [105]. Wang SW, Speck NA. Purification of core-binding factor, a protein that binds the conserved core site in murine leukemia virus enhancers. Mol Cell Biol. 1992;12:89-102.
    [106]. Meyers S, Downing JR, Hiebert SW. Identification of AML-1 and the (8;21) translocation protein (AML-1/ETO) as sequence-specific DNA-binding proteins: the runt homology domain is required for DNA binding and protein-protein interactions. Mol Cell Biol. 1993;13:6336-6345.
    [107]. Rossert J, Eberspaecher H, de Crombrugghe B. Separate cis-acting DNA elements of the mouse pro-alpha 1(I) collagen promoter direct expression of reporter genes to different type I collagen-producing cells in transgenic mice. J Cell Biol. 1995;129:1421-1432.
    [108]. Thomas DP, Sunters A, Gentry A, et al. Inhibition of chondrocyte differentiation in vitro by constitutive and inducible overexpression of the c-fos proto-oncogene. J Cell Sci. 2000;113 ( Pt 3):439-450.
    [109]. Guimont P, Grondin F, Dubois CM. Sox9-dependent transcriptional regulation of the proprotein convertase furin. Am J Physiol Cell Physiol. 2007;293:C172-183.
    [110]. Murdoch AD, Grady LM, Ablett MP, et al. Chondrogenic differentiation of human bone marrow stem cells in transwell cultures: generation of scaffold-free cartilage. Stem Cells. 2007;25:2786-2796.
    [111]. Han Y, Lefebvre V. L-Sox5 and Sox6 drive expression of the aggrecan gene in cartilage by securing binding of Sox9 to a far-upstream enhancer. Mol Cell Biol. 2008;28:4999-5013.
    [112]. Ikeda T, Kamekura S, Mabuchi A, et al. The combination of SOX5, SOX6, and SOX9 (the SOX trio) provides signals sufficient for induction of permanent cartilage. Arthritis Rheum. 2004;50:3561-3573.
    [113]. Goldring MB, Tsuchimochi K, Ijiri K. The control of chondrogenesis. J Cell Biochem. 2006;97:33-44.
    [114]. Sahni M, Ambrosetti DC, Mansukhani A, et al. FGF signaling inhibits chondrocyte proliferation and regulates bone development through the STAT-1 pathway. Genes Dev. 1999;13:1361-1366.
    [115]. Frank DA, Mahajan S, Ritz J. Fludarabine-induced immunosuppression is associated with inhibition of STAT1 signaling. Nat Med. 1999;5:444-447.
    [116]. Bobick BE, Kulyk WM. The MEK-ERK signaling pathway is a negative regulator of cartilage-specific gene expression in embryonic limb mesenchyme. J Biol Chem. 2004;279:4588-4595.
    [117]. de Crombrugghe B, Lefebvre V, Nakashima K. Regulatory mechanisms in the pathways of cartilage and bone formation. Curr Opin Cell Biol. 2001;13:721-727.
    [118]. Komori T. Runx2, a multifunctional transcription factor in skeletal development. J Cell Biochem. 2002;87:1-8.
    [119]. Komori T. Regulation of skeletal development by the Runx family of transcription factors. J Cell Biochem. 2005;95:445-453.
    [120]. Lee MH, Kim YJ, Kim HJ, et al. BMP-2-induced Runx2 expression is mediated by Dlx5, and TGF-beta 1 opposes the BMP-2-induced osteoblast differentiation by suppression of Dlx5 expression. J Biol Chem. 2003;278:34387-34394.
    [121]. Harada H, Tagashira S, Fujiwara M, et al. Cbfa1 isoforms exert functional differences in osteoblast differentiation. J Biol Chem. 1999;274:6972-6978.
    [122]. Xiao ZS, Hjelmeland AB, Quarles LD. Selective deficiency of the "bone-related" Runx2-II unexpectedly preserves osteoblast-mediated skeletogenesis. J Biol Chem. 2004;279:20307-20313.
    [123]. Hirata K, Tsukazaki T, Kadowaki A, et al. Transplantation of skin fibroblasts expressing BMP-2 promotes bone repair more effectively than those expressing Runx2. Bone. 2003;32:502-512.
    [124]. Mori-Akiyama Y, Akiyama H, Rowitch DH, et al. Sox9 is required for determination of the chondrogenic cell lineage in the cranial neural crest. Proc Natl Acad Sci U S A. 2003;100:9360-9365.
    [125]. Takahashi K, Tanabe K, Ohnuki M, et al. Induction of pluripotent stem cells from adult human fibroblasts by defined factors. Cell. 2007;131:861-872.
    [126]. Murakami S, Balmes G, McKinney S, et al. Constitutive activation of MEK1 in chondrocytes causesStat1-independent achondroplasia-like dwarfism and rescues the Fgfr3-deficient mouse phenotype. Genes Dev. 2004;18:290-305.
    [127]. Kim HS, Lee MS. STAT1 as a key modulator of cell death. Cell Signal. 2007;19:454-465.
    [128]. Sahni M, Raz R, Coffin JD, et al. STAT1 mediates the increased apoptosis and reduced chondrocyte proliferation in mice overexpressing FGF2. Development. 2001;128:2119-2129.
    [129]. Murakami S, Kan M, McKeehan WL, et al. Up-regulation of the chondrogenic Sox9 gene by fibroblast growth factors is mediated by the mitogen-activated protein kinase pathway. Proc Natl Acad Sci U S A. 2000;97:1113-1118.
    [130]. Chang SH, Oh CD, Yang MS, et al. Protein kinase C regulates chondrogenesis of mesenchymes via mitogen-activated protein kinase signaling. J Biol Chem. 1998;273:19213-19219.
    [131]. Oh CD, Chang SH, Yoon YM, et al. Opposing role of mitogen-activated protein kinase subtypes, erk-1/2 and p38, in the regulation of chondrogenesis of mesenchymes. J Biol Chem. 2000;275:5613-5619.
    [132]. Seghatoleslami MR, Roman-Blas JA, Rainville AM, et al. Progression of chondrogenesis in C3H10T1/2 cells is associated with prolonged and tight regulation of ERK1/2. J Cell Biochem. 2003;88:1129-1144.
    [133]. Yoon YM, Oh CD, Kim DY, et al. Epidermal growth factor negatively regulates chondrogenesis of mesenchymal cells by modulating the protein kinase C-alpha, Erk-1, and p38 MAPK signaling pathways. J Biol Chem. 2000;275:12353-12359.

© 2004-2018 中国地质图书馆版权所有 京ICP备05064691号 京公网安备11010802017129号

地址:北京市海淀区学院路29号 邮编:100083

电话:办公室:(+86 10)66554848;文献借阅、咨询服务、科技查新:66554700