Osteosarcoma: Cells-of-Origin, Cancer Stem Cells, and Targeted Therapies

Ander Abarrategi, Juan Tornin, Lucia Martinez-Cruzado, Ashley Hamilton, Enrique Martinez-Campos, Juan P Rodrigo, M Victoria González, Nicola Baldini, Javier Garcia-Castro, Rene Rodriguez, Ander Abarrategi, Juan Tornin, Lucia Martinez-Cruzado, Ashley Hamilton, Enrique Martinez-Campos, Juan P Rodrigo, M Victoria González, Nicola Baldini, Javier Garcia-Castro, Rene Rodriguez

Abstract

Osteosarcoma (OS) is the most common type of primary solid tumor that develops in bone. Although standard chemotherapy has significantly improved long-term survival over the past few decades, the outcome for those patients with metastatic or recurrent OS remains dismally poor and, therefore, novel agents and treatment regimens are urgently required. A hypothesis to explain the resistance of OS to chemotherapy is the existence of drug resistant CSCs with progenitor properties that are responsible of tumor relapses and metastasis. These subpopulations of CSCs commonly emerge during tumor evolution from the cell-of-origin, which are the normal cells that acquire the first cancer-promoting mutations to initiate tumor formation. In OS, several cell types along the osteogenic lineage have been proposed as cell-of-origin. Both the cell-of-origin and their derived CSC subpopulations are highly influenced by environmental and epigenetic factors and, therefore, targeting the OS-CSC environment and niche is the rationale for many recently postulated therapies. Likewise, some strategies for targeting CSC-associated signaling pathways have already been tested in both preclinical and clinical settings. This review recapitulates current OS cell-of-origin models, the properties of the OS-CSC and its niche, and potential new therapies able to target OS-CSCs.

Figures

Figure 1
Figure 1
Cell-of-origin for OS. The figure shows the most relevant cell types present in the bone microenvironment. MSCs, represented in a perivascular niche, and their derived cell types along the osteogenic lineage, such as the osteoblasts (OSB), are strong candidates to acquire the first cancer-promoting mutations and initiate OS formation.
Figure 2
Figure 2
OS-CSC niches. The figure represents possible niches for CSCs in OS. Suggested locations for CSCs are the perivascular niche, the endosteal niche, and areas of poor vascularization (hypoxic niche).

References

    1. Rosenberg A. E., Cleton-Jansen A. M., de Pinieux G., et al. Conventional osteosarcoma. In: Fletcher C. D. M., Bridge J. A., Hogendoorn P. C. W., Mertens F., editors. WHO Classification of Tumours of Soft Tissue and Bone. 4th. Lyon, France: International Agency for Research on Cancer; 2013. pp. 282–288.
    1. Alfranca A., Martinez-Cruzado L., Tornin J., et al. Bone microenvironment signals in osteosarcoma development. Cellular and Molecular Life Sciences. 2015;72(16):3097–3113. doi: 10.1007/s00018-015-1918-y.
    1. Botter S. M., Neri D., Fuchs B. Recent advances in osteosarcoma. Current Opinion in Pharmacology. 2014;16(1):15–23. doi: 10.1016/j.coph.2014.02.002.
    1. Allison D. C., Carney S. C., Ahlmann E. R., et al. A meta-analysis of osteosarcoma outcomes in the modern medical era. Sarcoma. 2012;2012:10. doi: 10.1155/2012/704872.704872
    1. Ng A. J., Mutsaers A. J., Baker E. K., Walkley C. R. Genetically engineered mouse models and human osteosarcoma. Clinical Sarcoma Research. 2012;2(1, article 19) doi: 10.1186/2045-3329-2-19.
    1. Tang N., Song W.-X., Luo J., Haydon R. C., He T.-C. Osteosarcoma development and stem cell differentiation. Clinical Orthopaedics and Related Research. 2008;466(9):2114–2130. doi: 10.1007/s11999-008-0335-z.
    1. Jones K. B. Osteosarcomagenesis: modeling cancer initiation in the mouse. Sarcoma. 2011;2011:10. doi: 10.1155/2011/694136.694136
    1. Malkin D., Jolly K. W., Barbier N., et al. Germline mutations of the p53 tumor-suppressor gene in children and young adults with second malignant neoplasms. The New England Journal of Medicine. 1992;326(20):1309–1315. doi: 10.1056/nejm199205143262002.
    1. Wong F. L., Boice J. D., Jr., Abramson D. H., et al. Cancer incidence after retinoblastoma: radiation dose and sarcoma risk. Journal of the American Medical Association. 1997;278(15):1262–1267. doi: 10.1001/jama.1997.03550150066037.
    1. Overholtzer M., Rao P. H., Favis R., et al. The presence of p53 mutations in human osteosarcomas correlates with high levels of genomic instability. Proceedings of the National Academy of Sciences of the United States of America. 2003;100(20):11547–11552.
    1. Wadayama B.-I., Toguchida J., Shimizu T., et al. Mutation spectrum of the retinoblastoma gene in osteosarcomas. Cancer Research. 1994;54(11):3042–3048.
    1. Rodriguez R., Rubio R., Menendez P. Modeling sarcomagenesis using multipotent mesenchymal stem cells. Cell Research. 2012;22(1):62–77. doi: 10.1038/cr.2011.157.
    1. Berman S. D., Calo E., Landman A. S., et al. Metastatic osteosarcoma induced by inactivation of Rb and p53 in the osteoblast lineage. Proceedings of the National Academy of Sciences of the United States of America. 2008;105(33):11851–11856. doi: 10.1073/pnas.0805462105.
    1. Lin P. P., Pandey M. K., Jin F., Raymond A. K., Akiyama H., Lozano G. Targeted mutation of p53 and Rb in mesenchymal cells of the limb bud produces sarcomas in mice. Carcinogenesis. 2009;30(10):1789–1795. doi: 10.1093/carcin/bgp180.
    1. Walkley C. R., Qudsi R., Sankaran V. G., et al. Conditional mouse osteosarcoma, dependent on p53 loss and potentiated by loss of Rb, mimics the human disease. Genes and Development. 2008;22(12):1662–1676. doi: 10.1101/gad.1656808.
    1. Calo E., Quintero-Estades J. A., Danielian P. S., Nedelcu S., Berman S. D., Lees J. A. Rb regulates fate choice and lineage commitment in vivo. Nature. 2010;466(7310):1110–1114. doi: 10.1038/nature09264.
    1. Benassi M. S., Molendini L., Gamberi G., et al. Alteration of prb/p16/cdk4 regulation in human osteosarcoma. International Journal of Cancer. 1999;84(5):489–493. doi: 10.1002/(sici)1097-0215(19991022)84:560;489::aid-ijc760;;2-d.
    1. López-Guerrero J. A., López-Ginés C., Pellín A., Carda C., Llombart-Bosch A. Deregulation of the G1 to S-phase cell cycle checkpoint is involved in the pathogenesis of human osteosarcoma. Diagnostic Molecular Pathology. 2004;13(2):81–91. doi: 10.1097/00019606-200406000-00004.
    1. Maelandsmo G. M., Berner J.-M., Florenes V. A., et al. Homozygous deletion frequency and expression levels of the CDKN2 gene in human sarcomas—relationship to amplification and mRNA levels of CDK4 and CCND1 . British Journal of Cancer. 1995;72(2):393–398. doi: 10.1038/bjc.1995.344.
    1. Zhang Y., Xiong Y., Yarbrough W. G. ARF promotes MDM2 degradation and stabilizes p53: ARF-INK4a locus deletion impairs both the Rb and p53 tumor suppression pathways. Cell. 1998;92(6):725–734. doi: 10.1016/S0092-8674(00)81401-4.
    1. Lonardo F., Ueda T., Huvos A. G., Healey J., Ladanyi M. p53 and MDM2 alterations in osteosarcomas: correlation with clinicopathologic features and proliferative rate. Cancer. 1997;79(8):1541–1547. doi: 10.1002/(sici)1097-0142(19970415)79:8<1541::aid-cncr15>;2-y.
    1. Ladanyi M., Cha C., Lewis R., Jhanwar S. C., Huvos A. G., Healey J. H. MDM2 gene amplification in metastatic osteosarcoma. Cancer Research. 1993;53(1):16–18.
    1. Lockwood W. W., Stack D., Morris T., et al. Cyclin E1 is amplified and overexpressed in osteosarcoma. Journal of Molecular Diagnostics. 2011;13(3):289–296. doi: 10.1016/j.jmoldx.2010.11.020.
    1. Gamberi G., Benassi M. S., Bohling T., et al. C-myc and c-fos in human osteosarcoma: prognostic value of mRNA and protein expression. Oncology. 1998;55(6):556–563. doi: 10.1159/000011912.
    1. Papachristou D. J., Batistatou A., Sykiotis G. P., Varakis I., Papavassiliou A. G. Activation of the JNK-AP-1 signal transduction pathway is associated with pathogenesis and progression of human osteosarcomas. Bone. 2003;32(4):364–371. doi: 10.1016/S8756-3282(03)00026-7.
    1. Wu J.-X., Carpenter P. M., Gresens C., et al. The proto-oncogene c-fos is over-expressed in the majority of human osteosarcomas. Oncogene. 1990;5(7):989–1000.
    1. Wang Z.-Q., Liang J., Schellander K., Wagner E. F., Grigoriadis A. E. c-fos-induced osteosarcoma formation in transgenic mice: cooperativity with c-jun and the role of endogenous c-fos. Cancer Research. 1995;55(24):6244–6251.
    1. Smida J., Baumhoer D., Rosemann M., et al. Genomic alterations and allelic imbalances are strong prognostic predictors in osteosarcoma. Clinical Cancer Research. 2010;16(16):4256–4267. doi: 10.1158/1078-0432.CCR-10-0284.
    1. Ueda T., Healey J. H., Huvos A. G., Ladanyi M. Amplification of the MYC gene in osteosarcoma secondary to Paget's disease of bone. Sarcoma. 1997;1(3-4):131–134. doi: 10.1080/13577149778209.
    1. Scionti I., Michelacci F., Pasello M., et al. Clinical impact of the methotrexate resistance-associated genes C-MYC and dihydrofolate reductase (DHFR) in high-grade osteosarcoma. Annals of Oncology. 2008;19(8):1500–1508. doi: 10.1093/annonc/mdn148.
    1. Lamoureux F., Picarda G., Rousseau J., et al. Therapeutic efficacy of soluble receptor activator of nuclear factor-κB-Fc delivered by nonviral gene transfer in a mouse model of osteolytic osteosarcoma. Molecular Cancer Therapeutics. 2008;7(10):3389–3398. doi: 10.1158/1535-7163.mct-08-0497.
    1. Lamoureux F., Richard P., Wittrant Y., et al. Therapeutic relevance of osteoprotegerin gene therapy in osteosarcoma: blockade of the vicious cycle between tumor cell proliferation and bone resorption. Cancer Research. 2007;67(15):7308–7318. doi: 10.1158/0008-5472.can-06-4130.
    1. Moriceau G., Ory B., Gobin B., et al. Therapeutic approach of primary bone tumours by bisphosphonates. Current Pharmaceutical Design. 2010;16(27):2981–2987. doi: 10.2174/138161210793563554.
    1. Quan G. M. Y., Choong P. F. M. Anti-angiogenic therapy for osteosarcoma. Cancer and Metastasis Reviews. 2006;25(4):707–713. doi: 10.1007/s10555-006-9031-1.
    1. Gobin B., Huin M. B., Lamoureux F., et al. BYL719, a new α-specific PI3K inhibitor: single administration and in combination with conventional chemotherapy for the treatment of osteosarcoma. International Journal of Cancer. 2015;136(4):784–796. doi: 10.1002/ijc.29040.
    1. Ma Y., Ren Y., Han E. Q., et al. Inhibition of the Wnt-β-catenin and Notch signaling pathways sensitizes osteosarcoma cells to chemotherapy. Biochemical and Biophysical Research Communications. 2013;431(2):274–279. doi: 10.1016/j.bbrc.2012.12.118.
    1. McQueen P., Ghaffar S., Guo Y., Rubin E. M., Zi X., Hoang B. H. The Wnt signaling pathway: implications for therapy in osteosarcoma. Expert Review of Anticancer Therapy. 2011;11(8):1223–1232. doi: 10.1586/era.11.94.
    1. Martins-Neves S. R., Corver W. E., Paiva-Oliveira D. I., et al. Osteosarcoma stem cells have active Wnt/β-catenin and overexpress SOX2 and KLF4. Journal of Cellular Physiology. 2016;231(4):876–886. doi: 10.1002/jcp.25179.
    1. Mu X., Isaac C., Greco N., Huard J., Weiss K. Notch signaling is associated with ALDH activity and an aggressive metastatic phenotype in murine osteosarcoma cells. Frontiers in Oncology. 2013;3, article 143 doi: 10.3389/fonc.2013.00143.
    1. Cai Y., Cai T., Chen Y. Wnt pathway in osteosarcoma, from oncogenic to therapeutic. Journal of Cellular Biochemistry. 2014;115(4):625–631. doi: 10.1002/jcb.24708.
    1. Li Y., Zhang J., Ma D., et al. Curcumin inhibits proliferation and invasion of osteosarcoma cells through inactivation of Notch-1 signaling. The FEBS Journal. 2012;279(12):2247–2259. doi: 10.1111/j.1742-4658.2012.08607.x.
    1. Chen G., Deng C., Li Y.-P. TGF-β and BMP signaling in osteoblast differentiation and bone formation. International Journal of Biological Sciences. 2012;8(2):272–288. doi: 10.7150/ijbs.2929.
    1. Lamora A., Mullard M., Amiaud J., et al. Anticancer activity of halofuginone in a preclinical model of osteosarcoma: inhibition of tumor growth and lung metastases. Oncotarget. 2015;6(16):14413–14427. doi: 10.18632/oncotarget.3891.
    1. Lamora A., Talbot J., Bougras G., et al. Overexpression of Smad7 blocks primary tumor growth and lung metastasis development in osteosarcoma. Clinical Cancer Research. 2014;20(19):5097–5112. doi: 10.1158/1078-0432.CCR-13-3191.
    1. Zhang H., Wu H., Zheng J., et al. Transforming growth factor β1 signal is crucial for dedifferentiation of cancer cells to cancer stem cells in osteosarcoma. STEM CELLS. 2013;31(3):433–446. doi: 10.1002/stem.1298.
    1. Kobayashi E., Hornicek F. J., Duan Z. MicroRNA involvement in osteosarcoma. Sarcoma. 2012;2012:8. doi: 10.1155/2012/359739.359739
    1. Di Fiore R., Drago-Ferrante R., Pentimali F., et al. MicroRNA-29b-1 impairs in vitro cell proliferation, self-renewal and chemoresistance of human osteosarcoma 3AB-OS cancer stem cells. International Journal of Oncology. 2014;45(5):2013–2023. doi: 10.3892/ijo.2014.2618.
    1. Xu M., Jin H., Xu C.-X., et al. miR-382 inhibits osteosarcoma metastasis and relapse by targeting y box-binding protein 1. Molecular Therapy. 2015;23(1):89–98. doi: 10.1038/mt.2014.197.
    1. Fujiwara T., Katsuda T., Hagiwara K., et al. Clinical relevance and therapeutic significance of microRNA-133a expression profiles and functions in malignant osteosarcoma-initiating cells. Stem Cells. 2014;32(4):959–973. doi: 10.1002/stem.1618.
    1. Song B., Wang Y., Titmus M. A., et al. Molecular mechanism of chemoresistance by miR-215 in osteosarcoma and colon cancer cells. Molecular Cancer. 2010;9, article 96 doi: 10.1186/1476-4598-9-96.
    1. Moriarity B. S., Otto G. M., Rahrmann E. P., et al. A Sleeping Beauty forward genetic screen identifies new genes and pathways driving osteosarcoma development and metastasis. Nature Genetics. 2015;47(6):615–624. doi: 10.1038/ng.3293.
    1. Kovac M., Blattmann C., Ribi S., et al. Exome sequencing of osteosarcoma reveals mutation signatures reminiscent of BRCA deficiency. Nature Communications. 2015;6, article 8940 doi: 10.1038/ncomms9940.
    1. Kuijjer M. L., Rydbeck H., Kresse S. H., et al. Identification of osteosarcoma driver genes by integrative analysis of copy number and gene expression data. Genes Chromosomes and Cancer. 2012;51(7):696–706. doi: 10.1002/gcc.21956.
    1. Ou M., Cai W.-W., Zender L., et al. MMP13, Birc2 (clAP1), and Birc3 (clAP2), amplified on chromosome 9, collaborate with p53 deficiency in mouse osteosarcoma progression. Cancer Research. 2009;69(6):2559–2567. doi: 10.1158/0008-5472.CAN-08-2929.
    1. Xiong Y., Wu S., Du Q., Wang A., Wang Z. Integrated analysis of gene expression and genomic aberration data in osteosarcoma (OS) Cancer Gene Therapy. 2015;22(11):524–529. doi: 10.1038/cgt.2015.48.
    1. Entz-Werle N., Lavaux T., Metzger N., et al. Involvement of MET/TWIST/APC combination or the potential role of ossification factors in pediatric high-grade osteosarcoma oncogenesis. Neoplasia. 2007;9(8):678–688. doi: 10.1593/neo.07367.
    1. Chen X., Bahrami A., Pappo A., et al. Recurrent somatic structural variations contribute to tumorigenesis in pediatric osteosarcoma. Cell Reports. 2014;7(1):104–112. doi: 10.1016/j.celrep.2014.03.003.
    1. Kansara M., Teng M. W., Smyth M. J., Thomas D. M. Translational biology of osteosarcoma. Nature Reviews Cancer. 2014;14(11):722–735. doi: 10.1038/nrc3838.
    1. Mendoza S., David H., Gaylord G. M., Miller C. W. Allelic loss at 10q26 in osteosarcoma in the region of the BUB3 and FGFR2 genes. Cancer Genetics and Cytogenetics. 2005;158(2):142–147. doi: 10.1016/j.cancergencyto.2004.08.035.
    1. Rao-Bindal K., Kleinerman E. S. Epigenetic regulation of apoptosis and cell cycle in osteosarcoma. Sarcoma. 2011;2011:5. doi: 10.1155/2011/679457.679457
    1. Wang L. L., Gannavarapu A., Kozinetz C. A., et al. Association between osteosarcoma and deleterious mutations in the RECQL4 gene Rothmund-Thomson syndrome. Journal of the National Cancer Institute. 2003;95(9):669–674. doi: 10.1093/jnci/95.9.669.
    1. Visvader J. E. Cells of origin in cancer. Nature. 2011;469(7330):314–322. doi: 10.1038/nature09781.
    1. Mutsaers A. J., Walkley C. R. Cells of origin in osteosarcoma: mesenchymal stem cells or osteoblast committed cells? Bone. 2014;62:56–63. doi: 10.1016/j.bone.2014.02.003.
    1. Xiao W., Mohseny A. B., Hogendoorn P. C., Cleton-Jansen A. M. Mesenchymal stem cell transformation and sarcoma genesis. Clinical Sarcoma Research. 2013;3(1, article 10) doi: 10.1186/2045-3329-3-10.
    1. Rodríguez R., García-Castro J., Trigueros C., García Arranz M., Menéndez P. Multipotent mesenchymal stromal cells: clinical applications and cancer modeling. Advances in Experimental Medicine and Biology. 2012;741:187–205. doi: 10.1007/978-1-4614-2098-9_13.
    1. Rodriguez R., Tornin J., Suarez C., et al. Expression of FUS-CHOP fusion protein in immortalized/transformed human mesenchymal stem cells drives mixoid liposarcoma formation. Stem Cells. 2013;31(10):2061–2072. doi: 10.1002/stem.1472.
    1. Mutsaers A. J., Ng A. J. M., Baker E. K., et al. Modeling distinct osteosarcoma subtypes in vivo using Cre:lox and lineage-restricted transgenic shRNA. Bone. 2013;55(1):166–178. doi: 10.1016/j.bone.2013.02.016.
    1. Molyneux S. D., Di Grappa M. A., Beristain A. G., et al. Prkar1a is an osteosarcoma tumor suppressor that defines a molecular subclass in mice. The Journal of Clinical Investigation. 2010;120(9):3310–3325. doi: 10.1172/jci42391.
    1. Sottnik J. L., Campbell B., Mehra R., Behbahani-Nejad O., Hall C. L., Keller E. T. Osteocytes serve as a progenitor cell of osteosarcoma. Journal of Cellular Biochemistry. 2014;115(8):1420–1429. doi: 10.1002/jcb.24793.
    1. Tao J., Jiang M.-M., Jiang L., et al. Notch activation as a driver of osteogenic sarcoma. Cancer Cell. 2014;26(3):390–401. doi: 10.1016/j.ccr.2014.07.023.
    1. Chan L. H., Wang W., Yeung W., Deng Y., Yuan P., Mak K. K. Hedgehog signaling induces osteosarcoma development through Yap1 and H19 overexpression. Oncogene. 2014;33(40):4857–4866. doi: 10.1038/onc.2013.433.
    1. Quist T., Jin H., Zhu J.-F., Smith-Fry K., Capecchi M. R., Jones K. B. The impact of osteoblastic differentiation on osteosarcomagenesis in the mouse. Oncogene. 2015;34(32):4278–4284. doi: 10.1038/onc.2014.354.
    1. Rubio R., García-Castro J., Gutiérrez-Aranda I., et al. Deficiency in p53 but not retinoblastoma induces the transformation of mesenchymal stem cells in vitro and initiates leiomyosarcoma in vivo . Cancer Research. 2010;70(10):4185–4194. doi: 10.1158/0008-5472.can-09-4640.
    1. Rubio R., Gutierrez-Aranda I., Sáez-Castillo A. I., et al. The differentiation stage of p53-Rb-deficient bone marrow mesenchymal stem cells imposes the phenotype of in vivo sarcoma development. Oncogene. 2013;32(41):4970–4980. doi: 10.1038/onc.2012.507.
    1. Rubio R., Abarrategi A., Garcia-Castro J., et al. Bone environment is essential for osteosarcoma development from transformed mesenchymal stem cells. Stem Cells. 2014;32(5):1136–1148. doi: 10.1002/stem.1647.
    1. Mohseny A. B., Szuhai K., Romeo S., et al. Osteosarcoma originates from mesenchymal stem cells in consequence of aneuploidization and genomic loss of Cdkn2. The Journal of Pathology. 2009;219(3):294–305. doi: 10.1002/path.2603.
    1. Shimizu T., Ishikawa T., Sugihara E., et al. C-MYC overexpression with loss of Ink4a/Arf transforms bone marrow stromal cells into osteosarcoma accompanied by loss of adipogenesis. Oncogene. 2010;29(42):5687–5699. doi: 10.1038/onc.2010.312.
    1. Cleton-Jansen A.-M., Anninga J. K., Briaire-de Bruijn I. H., et al. Profiling of high-grade central osteosarcoma and its putative progenitor cells identifies tumourigenic pathways. British Journal of Cancer. 2009;101(11):1909–1918. doi: 10.1038/sj.bjc.6605405.
    1. Lemos D. R., Eisner C., Hopkins C. I., Rossi F. M. V. Skeletal muscle-resident MSCs and bone formation. Bone. 2015;80:19–23. doi: 10.1016/j.bone.2015.06.013.
    1. Kaplan F. S., Chakkalakal S. A., Shore E. M. Fibrodysplasia ossificans progressiva: mechanisms and models of skeletal metamorphosis. Disease Models and Mechanisms. 2012;5(6):756–762. doi: 10.1242/dmm.010280.
    1. Basu-Roy U., Basilico C., Mansukhani A. Perspectives on cancer stem cells in osteosarcoma. Cancer Letters. 2013;338(1):158–167. doi: 10.1016/j.canlet.2012.05.028.
    1. Dela Cruz F. S. Cancer stem cells in pediatric sarcomas. Frontiers in Oncology. 2013;3, article 168 doi: 10.3389/fonc.2013.00168.
    1. Liu B., Ma W., Jha R. K., Gurung K. Cancer stem cells in osteosarcoma: recent progress and perspective. Acta Oncologica. 2011;50(8):1142–1150. doi: 10.3109/0284186x.2011.584553.
    1. Siclari V. A., Qin L. Targeting the osteosarcoma cancer stem cell. Journal of Orthopaedic Surgery and Research. 2010;5(1, article 78) doi: 10.1186/1749-799x-5-78.
    1. Tirino V., Desiderio V., Paino F., et al. Cancer stem cells in solid tumors: an overview and new approaches for their isolation and characterization. The FASEB Journal. 2013;27(1):13–24. doi: 10.1096/fj.12-218222.
    1. Gibbs C. P., Kukekov V. G., Reith J. D., et al. Stem-like cells in bone sarcomas: implications for tumorigenesis. Neoplasia. 2005;7(11):967–976. doi: 10.1593/neo.05394.
    1. Wang L., Park P., Lin C.-Y. Characterization of stem cell attributes in human osteosarcoma cell lines. Cancer Biology and Therapy. 2009;8(6):543–552.
    1. Wilson H., Huelsmeyer M., Chun R., Young K. M., Friedrichs K., Argyle D. J. Isolation and characterisation of cancer stem cells from canine osteosarcoma. Veterinary Journal. 2008;175(1):69–75. doi: 10.1016/j.tvjl.2007.07.025.
    1. Salerno M., Avnet S., Bonuccelli G., et al. Sphere-forming cell subsets with cancer stem cell properties in human musculoskeletal sarcomas. International Journal of Oncology. 2013;43(1):95–102. doi: 10.3892/ijo.2013.1927.
    1. He A., Yang X., Huang Y., et al. CD133+CD+ cells mediate in the lung metastasis of osteosarcoma. Journal of Cellular Biochemistry. 2015;116(8):1719–1729. doi: 10.1002/jcb.25131.
    1. Li J., Zhong X.-Y., Li Z.-Y., et al. CD133 expression in osteosarcoma and derivation of CD133+ cells. Molecular Medicine Reports. 2013;7(2):577–584. doi: 10.3892/mmr.2012.1231.
    1. Tirino V., Desiderio V., Paino F., et al. Human primary bone sarcomas contain CD133+ cancer stem cells displaying high tumorigenicity in vivo. FASEB Journal. 2011;25(6):2022–2030. doi: 10.1096/fj.10-179036.
    1. Adhikari A. S., Agarwal N., Wood B. M., et al. CD117 and Stro-1 identify osteosarcoma tumor-initiating cells associated with metastasis and drug resistance. Cancer Research. 2010;70(11):4602–4612. doi: 10.1158/0008-5472.CAN-09-3463.
    1. Tian J., Li X., Si M., Liu T., Li J. CD271+ osteosarcoma cells display stem-like properties. PLoS ONE. 2014;9(6, article e98549) doi: 10.1371/journal.pone.0098549.
    1. Murase M., Kano M., Tsukahara T., et al. Side population cells have the characteristics of cancer stem-like cells/cancer-initiating cells in bone sarcomas. British Journal of Cancer. 2009;101(8):1425–1432. doi: 10.1038/sj.bjc.6605330.
    1. Sun D.-X., Liao G.-J., Liu K.-G., Jian H. Endosialin-expressing bone sarcoma stem-like cells are highly tumor-initiating and invasive. Molecular Medicine Reports. 2015;12(4):5665–5670. doi: 10.3892/mmr.2015.4218.
    1. Honoki K., Fujii H., Kubo A., et al. Possible involvement of stem-like populations with elevated ALDH1 in sarcomas for chemotherapeutic drug resistance. Oncology Reports. 2010;24(2):501–505. doi: 10.3892/or-00000885.
    1. Wang L., Park P., Zhang H., La Marca F., Lin C.-Y. Prospective identification of tumorigenic osteosarcoma cancer stem cells in OS99-1 cells based on high aldehyde dehydrogenase activity. International Journal of Cancer. 2011;128(2):294–303. doi: 10.1002/ijc.25331.
    1. Levings P. P., McGarry S. V., Currie T. P., et al. Expression of an exogenous human Oct-4 promoter identifies tumor-initiating cells in osteosarcoma. Cancer Research. 2009;69(14):5648–5655. doi: 10.1158/0008-5472.CAN-08-3580.
    1. Yu L., Liu S., Guo W., et al. hTERT promoter activity identifies osteosarcoma cells with increased EMT characteristics. Oncology Letters. 2014;7(1):239–244. doi: 10.3892/ol.2013.1692.
    1. Yu L., Liu S., Zhang C., et al. Enrichment of human osteosarcoma stem cells based on hTERT transcriptional activity. Oncotarget. 2013;4(12):2326–2338. doi: 10.18632/oncotarget.1554.
    1. Di Fiore R., Santulli A., Ferrante R. D., et al. Identification and expansion of human osteosarcoma-cancer-stem cells by long-term 3-aminobenzamide treatment. Journal of Cellular Physiology. 2009;219(2):301–313. doi: 10.1002/jcp.21667.
    1. Tang Q.-L., Liang Y., Xie X.-B., et al. Enrichment of osteosarcoma stem cells by chemotherapy. Chinese Journal of Cancer. 2011;30(6):426–432. doi: 10.5732/cjc.011.10127.
    1. Naka N., Takenaka S., Araki N., et al. Synovial sarcoma is a stem cell malignancy. Stem Cells. 2010;28(7):1119–1131. doi: 10.1002/stem.452.
    1. Fujii H., Honoki K., Tsujiuchi T., Kido A., Yoshitani K., Takakura Y. Sphere-forming stem-like cell populations with drug resistance in human sarcoma cell lines. International Journal of Oncology. 2009;34(5):1381–1386. doi: 10.3892/ijo_00000265.
    1. Yan G.-N., Lv Y.-F., Guo Q.-N. Advances in osteosarcoma stem cell research and opportunities for novel therapeutic targets. Cancer Letters. 2016;370(2):268–274. doi: 10.1016/j.canlet.2015.11.003.
    1. Basu-Roy U., Seo E., Ramanathapuram L., et al. Sox2 maintains self renewal of tumor-initiating cells in osteosarcomas. Oncogene. 2012;31(18):2270–2282. doi: 10.1038/onc.2011.405.
    1. Salerno M., Avnet S., Bonuccelli G., Hosogi S., Granchi D., Baldini N. Impairment of lysosomal activity as a therapeutic modality targeting cancer stem cells of embryonal rhabdomyosarcoma cell line RD. PLoS ONE. 2014;9(10) doi: 10.1371/journal.pone.0110340.e110340
    1. He H., Ni J., Huang J. Molecular mechanisms of chemoresistance in osteosarcoma (Review) Oncology Letters. 2014;7(5):1352–1362. doi: 10.3892/ol.2014.1935.
    1. Martins-Neves S. R., Lopes Á. O., do Carmo A., et al. Therapeutic implications of an enriched cancer stem-like cell population in a human osteosarcoma cell line. BMC Cancer. 2012;12, article 139 doi: 10.1186/1471-2407-12-139.
    1. Yang M., Yan M., Zhang R., Li J., Luo Z. Side population cells isolated from human osteosarcoma are enriched with tumor-initiating cells. Cancer Science. 2011;102(10):1774–1781. doi: 10.1111/j.1349-7006.2011.02028.x.
    1. Gonçalves C., Martins-Neves S. R., Paiva-Oliveira D., Oliveira V. E. B., Fontes-Ribeiro C., Gomes C. M. F. Sensitizing osteosarcoma stem cells to doxorubicin-induced apoptosis through retention of doxorubicin and modulation of apoptotic-related proteins. Life Sciences. 2015;130:47–56. doi: 10.1016/j.lfs.2015.03.009.
    1. Plaks V., Kong N., Werb Z. The cancer stem cell niche: how essential is the niche in regulating stemness of tumor cells? Cell Stem Cell. 2015;16(3):225–238. doi: 10.1016/j.stem.2015.02.015.
    1. Kuhn N. Z., Tuan R. S. Regulation of stemness and stem cell niche of mesenchymal stem cells: implications in tumorigenesis and metastasis. Journal of Cellular Physiology. 2010;222(2):268–277. doi: 10.1002/jcp.21940.
    1. Zeng W., Wan R., Zheng Y., Singh S. R., Wei Y. Hypoxia, stem cells and bone tumor. Cancer Letters. 2011;313(2):129–136. doi: 10.1016/j.canlet.2011.09.023.
    1. Boyce B. F., Xing L. Functions of RANKL/RANK/OPG in bone modeling and remodeling. Archives of Biochemistry and Biophysics. 2008;473(2):139–146. doi: 10.1016/j.abb.2008.03.018.
    1. Choong P. F. M., Broadhead M. L., Clark J. C. M., Myers D. E., Dass C. R. The molecular pathogenesis of osteosarcoma: a review. Sarcoma. 2011;2011:12. doi: 10.1155/2011/959248.959248
    1. Harwood J. L., Alexander J. H., Mayerson J. L., Scharschmidt T. J. Targeted chemotherapy in bone and soft-tissue sarcoma. Orthopedic Clinics of North America. 2015;46(4):587–608. doi: 10.1016/j.ocl.2015.06.011.
    1. Hattinger C. M., Fanelli M., Tavanti E., et al. Advances in emerging drugs for osteosarcoma. Expert Opinion on Emerging Drugs. 2015;20(3):495–514. doi: 10.1517/14728214.2015.1051965.
    1. Heymann D., Rédini F. Targeted therapies for bone sarcomas. Bonekey Reports. 2013;2, article 378 doi: 10.1038/bonekey.2013.112.
    1. Cathomas R., Rothermundt C., Bode B., Fuchs B., Von Moos R., Schwitter M. RANK ligand blockade with denosumab in combination with sorafenib in chemorefractory osteosarcoma: a possible step forward? Oncology. 2014;88(4):257–260. doi: 10.1159/000369975.
    1. Sampson V. B., Gorlick R., Kamara D., Anders Kolb E. A review of targeted therapies evaluated by the pediatric preclinical testing program for osteosarcoma. Frontiers in Oncology. 2013;3, article 132 doi: 10.3389/fonc.2013.00132.
    1. DeRenzo C., Gottschalk S. Genetically modified T-cell therapy for osteosarcoma. Advances in Experimental Medicine and Biology. 2014;804:323–340. doi: 10.1007/978-3-319-04843-7_18.
    1. Kawano M., Itonaga I., Iwasaki T., Tsuchiya H., Tsumura H. Anti-TGF-β antibody combined with dendritic cells produce antitumor effects in osteosarcoma. Clinical Orthopaedics and Related Research. 2012;470(8):2288–2294. doi: 10.1007/s11999-012-2299-2.
    1. Rainusso N., Brawley V. S., Ghazi A., et al. Immunotherapy targeting HER2 with genetically modified T cells eliminates tumor-initiating cells in osteosarcoma. Cancer Gene Therapy. 2012;19(3):212–217. doi: 10.1038/cgt.2011.83.
    1. Tarek N., Lee D. A. Natural killer cells for osteosarcoma. Advances in Experimental Medicine and Biology. 2014;804:341–353. doi: 10.1007/978-3-319-04843-7_19.
    1. Di Fiore R., Fanale D., Drago-Ferrante R., et al. Genetic and molecular characterization of the human osteosarcoma 3AB-OS cancer stem cell line: a possible model for studying osteosarcoma origin and stemness. Journal of Cellular Physiology. 2013;228(6):1189–1201. doi: 10.1002/jcp.24272.
    1. Gemei M., Corbo C., D'Alessio F., Di Noto R., Vento R., Del Vecchio L. Surface proteomic analysis of differentiated versus stem-like osteosarcoma human cells. Proteomics. 2013;13(22):3293–3297. doi: 10.1002/pmic.201300170.
    1. Zuch D., Giang A.-H., Shapovalov Y., et al. Targeting radioresistant osteosarcoma cells with parthenolide. Journal of Cellular Biochemistry. 2012;113(4):1282–1291. doi: 10.1002/jcb.24002.
    1. Mongre R. K., Sodhi S. S., Ghosh M., et al. The novel inhibitor BRM270 downregulates tumorigenesis by suppression of NF-κB signaling cascade in MDR-induced stem like cancer-initiating cells. International Journal of Oncology. 2015;46(6):2573–2585. doi: 10.3892/ijo.2015.2961.
    1. Gong C., Liao H., Wang J., et al. LY294002 induces G0/G1 cell cycle arrest and apoptosis of cancer stem-like cells from human osteosarcoma via downregulation of PI3K activity. Asian Pacific Journal of Cancer Prevention. 2012;13(7):3103–3107. doi: 10.7314/APJCP.2012.13.7.3103.
    1. Yi X.-J., Zhao Y.-H., Qiao L.-X., Jin C.-L., Tian J., Li Q.-S. Aberrant Wnt/β-catenin signaling and elevated expression of stem cell proteins are associated with osteosarcoma side population cells of high tumorigenicity. Molecular Medicine Reports. 2015;12(4):5042–5048. doi: 10.3892/mmr.2015.4025.
    1. Krause U., Ryan D. M., Clough B. H., Gregory C. A. An unexpected role for a Wnt-inhibitor: Dickkopf-1 triggers a novel cancer survival mechanism through modulation of aldehyde-dehydrogenase-1 activity. Cell Death and Disease. 2014;5(2) doi: 10.1038/cddis.2014.67.e1093
    1. Song B., Wang Y., Xi Y., et al. Mechanism of chemoresistance mediated by miR-140 in human osteosarcoma and colon cancer cells. Oncogene. 2009;28(46):4065–4074. doi: 10.1038/onc.2009.274.
    1. Wang Y., Yao J., Meng H., et al. A novel long non-coding RNA, hypoxia-inducible factor-2α promoter upstream transcript, functions as an inhibitor of osteosarcoma stem cells in vitro. Molecular Medicine Reports. 2015;11(4):2534–2540. doi: 10.3892/mmr.2014.3024.
    1. Chen X., Hu C., Zhang W., et al. Metformin inhibits the proliferation, metastasis, and cancer stem-like sphere formation in osteosarcoma MG63 cells in vitro. Tumor Biology. 2015;36(12):9873–9883. doi: 10.1007/s13277-015-3751-1.
    1. Quattrini I., Conti A., Pazzaglia L., et al. Metformin inhibits growth and sensitizes osteosarcoma cell lines to cisplatin through cell cycle modulation. Oncology Reports. 2014;31(1):370–375. doi: 10.3892/or.2013.2862.
    1. Chang Y., Zhao Y., Gu W., et al. Bufalin inhibits the differentiation and proliferation of cancer stem cells derived from primary osteosarcoma cells through Mir-148a. Cellular Physiology and Biochemistry. 2015;36(3):1186–1196. doi: 10.1159/000430289.
    1. Chang Y., Zhao Y., Zhan H., Wei X., Liu T., Zheng B. Bufalin inhibits the differentiation and proliferation of human osteosarcoma cell line hMG63-derived cancer stem cells. Tumor Biology. 2014;35(2):1075–1082. doi: 10.1007/s13277-013-1143-y.
    1. Tang Q.-L., Zhao Z.-Q., Li J.-C., et al. Salinomycin inhibits osteosarcoma by targeting its tumor stem cells. Cancer Letters. 2011;311(1):113–121. doi: 10.1016/j.canlet.2011.07.016.
    1. Ni M., Xiong M., Zhang X., et al. Poly(lactic-co-glycolic acid) nanoparticles conjugated with CD133 aptamers for targeted salinomycin delivery to CD133+ osteosarcoma cancer stem cells. International Journal of Nanomedicine. 2015;10:2537–2554. doi: 10.2147/ijn.s78498.
    1. Li J., Liu W., Zhao K., et al. Diallyl trisulfide reverses drug resistance and lowers the ratio of CD133+ cells in conjunction with methotrexate in a human osteosarcoma drug-resistant cell subline. Molecular Medicine Reports. 2009;2(2):245–252. doi: 10.3892/mmr-00000091.
    1. Li Y., Zhang J., Zhang L., Si M., Yin H., Li J. Diallyl trisulfide inhibits proliferation, invasion and angiogenesis of osteosarcoma cells by switching on suppressor microRNAs and inactivating of Notch-1 signaling. Carcinogenesis. 2013;34(7):1601–1610. doi: 10.1093/carcin/bgt065.
    1. Di Pompo G., Salerno M., Rotili D., et al. Novel histone deacetylase inhibitors induce growth arrest, apoptosis, and differentiation in sarcoma cancer stem cells. Journal of Medicinal Chemistry. 2015;58(9):4073–4079. doi: 10.1021/acs.jmedchem.5b00126.
    1. Mu X., Brynien D., Weiss K. R. The HDAC inhibitor Vorinostat diminishes the in vitro metastatic behavior of Osteosarcoma cells. BioMed Research International. 2015;2015:6. doi: 10.1155/2015/290368.290368
    1. Xu J.-F., Pan X.-H., Zhang S.-J., et al. CD47 blockade inhibits tumor progression human osteosarcoma in xenograft models. Oncotarget. 2015;6(27):23662–23670. doi: 10.18632/oncotarget.4282.
    1. Vermeulen L., de Sousa e Melo F., Richel D. J., Medema J. P. The developing cancer stem-cell model: clinical challenges and opportunities. The Lancet Oncology. 2012;13(2):e83–e89. doi: 10.1016/s1470-2045(11)70257-1.
    1. Tornin J., Martinez-Cruzado L., Santos L., et al. Inhibition of SP1 by the mithramycin analog EC-8042 efficiently targets tumor initiating cells in sarcoma. Oncotarget. 2016 doi: 10.18632/oncotarget.8817.

Source: PubMed

3
Abonneren