Next Article in Journal
Genome-Wide Mapping of Growth-Related Quantitative Trait Loci in Orange-Spotted Grouper (Epinephelus coioides) Using Double Digest Restriction-Site Associated DNA Sequencing (ddRADseq)
Next Article in Special Issue
GL-1196 Suppresses the Proliferation and Invasion of Gastric Cancer Cells via Targeting PAK4 and Inhibiting PAK4-Mediated Signaling Pathways
Previous Article in Journal
Power Frequency Magnetic Fields Affect the p38 MAPK-Mediated Regulation of NB69 Cell Proliferation Implication of Free Radicals
Previous Article in Special Issue
Molecular Insight into Gut Microbiota and Rheumatoid Arthritis
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Present Advances and Future Perspectives of Molecular Targeted Therapy for Osteosarcoma

1
Institute for Advancing Translational Medicine in Bone & Joint Diseases, School of Chinese Medicine, Hong Kong Baptist University, Hong Kong 999077, China
2
Department of Orthopaedic Surgery, Shenzhen Hospital, Southern Medical University, Shenzhen 518100, China
3
Orthopaedic Surgery Department, Nanfang Hospital, Southern Medical University, Guangzhou 510515, China
4
Department of Orthopaedic Surgery, the First Hospital of Huizhou, Huizhou 516000, China
*
Authors to whom correspondence should be addressed.
These authors contributed equally to this work.
Int. J. Mol. Sci. 2016, 17(4), 506; https://0-doi-org.brum.beds.ac.uk/10.3390/ijms17040506
Submission received: 27 February 2016 / Accepted: 30 March 2016 / Published: 6 April 2016
(This article belongs to the Special Issue Translational Molecular Medicine & Molecular Drug Discovery)

Abstract

:
Osteosarcoma (OS) is a bone cancer mostly occurring in pediatric population. Current treatment regime of surgery and intensive chemotherapy could cure about 60%–75% patients with primary osteosarcoma, however only 15% to 30% can be cured when pulmonary metastasis or relapse has taken place. Hence, novel precise OS-targeting therapies are being developed with the hope of addressing this issue. This review summarizes the current development of molecular mechanisms and targets for osteosarcoma. Therapies that target these mechanisms with updated information on clinical trials are also reviewed. Meanwhile, we further discuss novel therapeutic targets and OS-targeting drug delivery systems. In conclusion, a full insight in OS pathogenesis and OS-targeting strategies would help us explore novel targeted therapies for metastatic osteosarcoma.

Graphical Abstract

1. Introduction

Osteosarcoma (OS) is a primary bone cancer, predominantly affecting children and adolescence population [1]. OS originates from primitive mesenchymal bone forming cells and often occurs in long bones, such as proximal tibia and distal femur [2,3]. Current OS treatment regime consists of the combination of surgery and intensive multi-agent chemotherapy. Ever since the introduction of chemotherapy, five-year survival rate among OS patients has improved to 60%–75% [4]. However, 30%–40% of OS patients are associated with pulmonary metastasis and relapse, which have significantly poor prognosis, with an overall five-year survival rate of about 20% [5]. Moreover, over the past few decades, no substantial improvement in survival rate has been achieved, though efforts were made by intensifying dosing, varying timing and using multi-combinational chemotherapy. Additionally, several adverse effects are accompanied by high-dose chemotherapy [6]. Hence, there is an increasing sense of urgency to identify new biological markers and develop novel, innovative and specific molecular targeted therapeutic approaches to improve the outcome in osteosarcoma patients with poor prognosis. In this review, we will discuss about molecular mechanism underlying osteosarcoma, current molecular therapeutic targets against immune system, extracellular and Intercellular signaling transduction pathway of the bone metabolism, as well as novel therapeutic targets and drug delivery systems that have been investigated or are currently undergoing investigation in translational studies (Figure 1).
Molecular mechanism underlying osteosarcoma is characterized by complex karyotype, extensively unstable genome and complicated interaction between multiple proteins and signaling pathways [7,8]. Several groups have reported association between mutations and tumor suppressor genes. Retinoblastoma 1 (RB1) and tumor protein p53 (TP53) have been implicated in osteosarcoma pathogenesis. Mutations in the Rb1 gene and high frequency of allelic loss of Rb1 gene at 13q and p53 gene at 17p have been detected in OS samples [9,10]. Moreover RB1 inactivation was detected in almost 20%–40% of OS cases and associated with poor prognosis [11]. Inactivation or loss of TP53 causes cells to be resistant to DNA damage as well as loss control of cell growth thus resulting in extensive proliferation, transformation, and evade toxic effects of DNA damaging agents [12]. Recent studies have identified, several genes including Apurinic/Apyrimidinic exonuclease 1 (APEX1), Myc and epidermal growth factor receptor 2 (HER2/neu, ErbB2) in osteosarcoma pathogenesis. APEX1 gene were found to be amplified with their respective protein overexpressed and could also correlate well with recurrence, metastasis, and survival in osteosarcoma patients [13]. Myc is a transcription factor that stimulates cell growth and mitosis. High expression of Myc in bone marrow stromal cells caused loss of adipogenesis and transformation into osteosarcoma [14]. Myc was also found to be amplified in OS cells lines resistant to conventional chemotherapy [15]. Higher levels of human epidermal growth factor receptor 2 (HER2/neu, ErbB2) in OS patient samples were demonstrated to be associated with early pulmonary metastases and poor prognosis [16]. In addition, other molecules which have been found to be associated with osteosarcoma pathogenesis are bone morphogenetic protein type II receptor (BMPR2) and high mobility group box1 (HMGB1) [17,18]. OS tumor pathogenesis is not bound to a specific gene mutation or gene loss or to aberrant signaling pathways, but a complex entangled mechanism. Therefore, several molecules are being investigated in order to develop targeted therapies for OS.

2. Immunomodulators

2.1. Interferons

Interferons (IFNs) are a group of signaling proteins known as cytokines, which stimulate immune system and induce anti-angiogenic and antitumor activity. IFNs have successfully demonstrated high clinical activity associated with several types of cancers [19]. IFN-α-2b has shown to exhibit inhibition on the growth of patient-derived OS cells and also on the tumor growth in PDX mouse models. In a pilot study of non-metastatic OS patients treated with IFN-α for 3–5 years, five-year disease-free survival at 63% was reported [20]. Additionally, high-grade osteosarcoma patients treated with IFN-α for 3–5 years achieved 10-year disease-free survival at 43% [21]. A recent report investigated the efficacy of IFN-α-2b as a maintenance therapy in OS patients with good response to MAP (methotrexate, doxorubicin, and cisplatin) induction chemotherapy, but no improvement in outcomes with event-free survival (EFS) were presented [22].

2.2. GM-CSF

Granulocyte-macrophage colony-stimulating factor (GM-CSF) is also an immunomodulatory cytokine tested in osteosarcoma. As an effective immunomodulator, the activity of GM-CSF against melanoma and Ewing sarcoma has been well demonstrated [23,24]. A phase I study of aerosolized GM-CSF showed that it was well-tolerated and effective in selected melanoma and Ewing sarcoma patients [25]. Moreover, a phase II trial of aerosolized GM-CSF on OS patients with first recurrence of pulmonary metastasis demonstrated that it was a feasible and safe treatment option. However, GM-CSF had no signs of immunostimulatory effects on pulmonary metastases or any improvement in outcome in OS patients [26].

2.3. Mifamurtide

With immune stimulatory activity, muramyl dipeptide (MDP) is the minimal bioactive peptidoglycan component in the cell wall of Bacille Calmette–Guerin mycobacterium [27]. Mifamurtide (MTP-PE or muramyl tripeptide phosphatidyl ethanolamine) is a synthetic derivative of MDP. Liposome encapsulation of mifamurtide favors targeted delivery of MTP-PE to monocytes and macrophages, which activates the tumoricidal activity of macrophages and monocytes via secreting interleukin(IL)-6, tumor necrosis factor (TNF)-αand increased phagocytosis[28,29]. Early in vivo activity of mifamurtide was reported in dogs with spontaneous OS, and the treatment with MTP-PE following amputation had significantly improved disease-free survival to 222 days, compared to 77 days in the placebo group [30,31]. Since then, several clinical trials have been performed in humans. A Phase III, randomized, prospective intergroup trial (INT-0133) study of mifamurtide on patients with newly diagnosed osteosarcoma, showed significant improvement in six-year overall survival from 70% to 78% and in patients with metastatic disease showed improvement in five-year overall survival from 40% to 53% [32,33]. Numerous studies have reported of promising clinical benefits when mifamurtide is combined with chemotherapy in treatment of metastatic OS [34]. The drug has been currently approved as an adjuvant treatment of osteosarcoma by European Medical Agency, but has not been approved by the US FDA. Hence, given the promising data, further investigation is needed to clarify the role of mifamurtide in treatment of OS. Currently, several clinical trials of mifamurtide’s effectiveness in treating OS are being conducted.

3. Tyrosine Kinase Receptor Inhibitors

3.1. Receptor Tyrosine Kinases (RTKs)

RTKs are cell-surface receptors which play a key role in the activation of multiple downstream signaling pathways including, phosphatidylinositol 3 (PI3)/Akt kinase and extracellular signal regulated kinase (Erk) [35]. And hence is an important mediator in regulation of normal cellular as well as physiological processes such as cell growth, survival and proliferation. Moreover, RTKs have been arraigned as a key factor in growth and progression of several tumors and several gene mutation, amplification have been implicated in the disruption of RTKs signaling cascade [36]. Here we list a few RTKs undergoing clinical trials that are involved in pathogenesis of OS (Table 1).

3.2. Insulin-Like Growth Factor Receptor (IGF-R)

IGF-R is a membrane tyrosine kinase receptor and presents in two subtypes IGF-1R and IGF-2R. IGF-1R is a transmembrane RTK and has been implicated in mediating cell differentiation, proliferation, and apoptosis in several cancers [48]. The insulin-like growth factor-1 and 2 (IGF-1 and IGF-2) bind to the IGF-1R, undergoes auto-phosphorylation and activates two major oncogenic signaling cascades: the phosphoinositide 3-kinase (PI3K/Akt) pathway and the mitogen-activated protein kinase (MAPK) pathway [49,50]. IGF-R overexpression has been observed in several OS cell lines, and is reliant on activation of IGF-1R by IGF- 1 to stimulate OS cell growth and metastatic dormancy in in vivo and in vitro [48,51]. Also IGF-R levels were seen to be elevated among OS patients tumor samples and further the elevated expression of IGF-1R and IGF-1 ligand correlated with the poor prognosis and survival rate in OS patients [52,53].
Current anti-IGF-R therapeutic approaches consist of human monoclonal antibodies (mAbs) targeting IGF-1R, IGF ligand-neutralizing antibodies and small-molecule tyrosine kinase inhibitors of IGF-1R. Several human monoclonal antibodies (mAbs) targeting IGF-1R has been developed and some of them has been or are being investigated in different clinical trials. Cixutumumab is a fully human IgG1 mAbs specifically targeting IGF-R. Phase I/II clinical trial of cixutumumab on children with refractory solid tumors including OS, reported cixutumumab to be well tolerated but with limited single-agent activity [37,38]. Initial phase II trials, combination of cixutumumab and the mTOR inhibitor temsirolimus had shown clinical activity, but a recent phase II trial could not achieve the objective response. Studies on another fully human mAb SCH 717454 (robatumumab), had revealed it to be less effective in vitro but had significant tumor regression by inhibiting cell proliferation and angiogenesis in several OS xenograft models [54]. Furthermore, SCH 717454 in combination with cisplatin or cyclophosphamide had demonstrated a remarkable increase in in vivo antitumor activity compared with single agent treatment [54]. However, a phase 1/1B trial of SCH 717454 in combination with different treatment regimens in pediatric patients with advanced solid tumors (NCT00960063 *) and a phase II trial on activity of SCH 717454 in patients with relapsed OS and Ewing’s sarcoma (NCT00617890 *) were recently terminated (Table 1).
Two IGF ligand-neutralizing antibodies against IGF ligands IGF-I and -II have been found: BI836845 and MEDI-573. Both BI 836845 and MEDI-573 are fully human monoclonal antibody with high affinity towards IGF ligands IGF-I and IGF-II, and selectively neutralizes the bioactivity of IGF ligands thereby down regulating IGF signaling through both the IGF-1R and IR-A pathways [55,56]. At present, several phase I trials of BI836845 for patients with solid tumors are ongoing recruitment (Table 1). In addition to mAbs, small-molecule tyrosine kinase inhibitors of IGF-1R are also being investigated. BMS-754807 is a reversible small molecule ATP-competitive inhibitor of IGF-1R. Evaluation of BMS-754807 against the Pediatric Preclinical Testing Program (PPTP) has reported to have a broad activity towards pediatric sarcomas and xenograft models through IGF-1R inhibition and efficacy of BMS-754807 could be enhanced when used in combinations, by targeting multiple, related signaling pathways [57]. However, even though anti-IGF-1R therapeutic approach has shown clinical benefits in some pediatric subjects, several clinical trials on anti-IGF-1R agents have been terminated, because the observed clinical benefits of targeting IGF-1R pathway in single and multi-agent strategies could not met the primary response.

3.3. Platelet-Derived Growth Factor/Receptors (PDGF/PDGFR)

Platelet-Derived Growth Factor/receptors (PDGF/PDGFR) are signaling molecules which play an important role in cell proliferation and differentiation of osteoblasts and osteoclasts [58]. PDGF/PDGFR signaling pathway, more specifically PDGFR-α/PDGF-A has been implicated in tumor growth and metastasis in variety of human solid tumors, including OS and the overexpression of PDGFR correlates with the disease progression and poor prognosis [59,60]. While some conflicting reports disagree about correlation between PDGF-A expression and prognosis in OS [61]. A recent in vivo study on OS cells–platelet interactions reported that OS cells, when in contact with platelets, could induce platelet aggregation and stimulate release of platelet-derived growth factor (PDGF) from platelets. Furthermore, platelets cause phosphorylation of PDGFR and Akt, there by promoting the proliferation of OS cells [62].
Imatinib mesylate (STI-571) is potent tyrosine kinase inhibitors against PDGFR signaling and was originally developed for the treatment of chronic myelogenous leukemia (CML). Imatinib mesylate also has multi-targeted inhibitory activity towards several kinases including, c-KIT and Bcr/Abl [63]. In vitro studies of imatinib mesylate on various OS cell lines had demonstrated imatinib mesylate to have variable toxicity and could not be viewed as a single agent for the treatment of osteosarcoma [64], while another study had reported imatinib mesylate to exert a dose-dependent anti-proliferative effect in most of the human, mouse and rat OS cell lines through cell cycle arrest, inducing caspase-dependent cell death and inhibiting cell migration. Additionally, in murine syngeneic models of undifferentiated- and mixed osteoblastic/osteolytic forms of OS, imatinib mesylate could inhibit the tumor growth in both preventive and curative approaches [65]. However, in a phase II clinical trial of imatinib mesylate in children with refractory or relapsed solid tumors including OS, it failed to demonstrate significant response [42] (Table 1).
Other multi-targeted RTK inhibitors, which also block PDGF/PDGFR signaling, are sunitinib and dasatinib. Sunitinib is a small molecule, which is approved for treatment of renal cell carcinoma (RCC) and imatinib-resistant gastrointestinal stromal tumor (GIST). Sunitinib can also inhibit VEGFR and FLT-3 (fms-related tyrosine kinase-3). Sunitinib showed anti-tumor activity in in vitro as well as in OS xenograft models, but this may be due to anti-angiogenic effect instead of PDGFR inhibition [66,67]. In a phase 1 trial of sunitinib in pediatric population with solid tumors, initially treated with anthracyclins, reported best response in four patients including one OS patient, with toxicities mainly representing hematological and cardiac symptoms [43,68] (Table 1). Dasatinib a multi-targeted RTK inhibitors also inhibits SRC and BCR-ABL [69].

3.4. Vascular Endothelial Growth Factor (VEGF)

The vascular endothelial growth factor (VEGF) ligands and receptors (VEGFR) are implicated in tumor angiogenesis and play an important part in tumor growth and development. VEGF receptor consists of a family of receptors (VEGFR-1, VEGFR-2 and VEGFR-3) which could be activated by their ligands the VEGFs (A, B, C and D) [70]. VEGF overexpression has been observed in several tumors including OS. In particular, expression levels of VEGFR-3 correlated with pulmonary metastasis development and overall survival in patients with OS [71,72]. Several preclinical studies have revealed that specifically targeting VEGFs ligands or its receptors, may be an effective strategy in treating OS [73].
At present VEGF-based anti-angiogenic therapies, include anti-VEGF antibodies and small-molecule tyrosine kinase inhibitors against VEGFR. Bevacizumab, is a humanized monoclonal antibody (mAb) targeted towards the VEGF receptor and had shown some beneficial preclinical results [74]. However, phase I evaluation of bevacizumab in children with refractory solid tumors including OS, reported no observable response [44]. Currently a phase I and II trail of bevacizumab in combination with several chemotherapy agents on OS is ongoing (NCT00458731 *) (Table 1).
The PPTP has successfully identified three potent small molecule TKIs, which selectively target VEGF receptor family. These are sunitinib, sorafenib and cediranib. Sunitinib and sorafenib are multi- targeted inhibitors, with overlapping targets (PDGFR, FLT3, RET, KIT and VEGFR). Sorafenib was first approved for the treatment of un-resectable hepatocellular carcinoma (HCC). Preclinical studies of sorafenib, reported promising results with reduction in tumor growth and inhibit pulmonary metastasis in OS models [75,76]. In a phase II study of sorafenib as single treatment agent in patients with relapsed and unresectable OS, progression-free survival (PFS) at four months was 46%, while overall survival was seven months, objective response was seen in 14% of patients and 29% of patients had stable disease [45]. Further clinical evaluation of sorafenib in combination with bevacizumab (NCT00665990) and irinotecan (NCT01518413) are in progress (Table 1).

3.5. Human Epidermal Growth Factor Receptor 2 (HER-2)

HER-2 is RTKs and belongs to the human epidermal growth factor receptor (HER/EGFR/ERBB) family. HER-2 expression is linked to the development and progression of several tumors. Recent development of targeted therapies directed towards HER-2 have achieved significant benefits in the treatment of several solid tumors [77]. However, there are some conflicting reports regarding HER-2 expression in OS. While, some studies have reported over expression of HER-2 in 40% of OS patient samples and its increased expression correlated with metastasis, relapse and poor prognosis [78,79]. However, other studies have reported minimal HER-2 expression in OS tumor samples or no association between cytoplasmic HER-2 expression in OS tumor samples and patient’s prognosis [80,81].
Trastuzumab is a monoclonal antibody targeting HER-2 receptors and was primarily used in the treatment of breast cancer. In a phase II study of trastuzumab in addition to conventional chemotherapy has been shown to be safe, but with no clinical benefits. Briefly, this study also demonstrated non-significant difference in event-free and overall survival between the HER-2-positive group treated with trastuzumab and the HER-2 negative group treated with conventional chemotherapy [46,82] (Table 1).

4. Intracellular Signaling Inhibitors

4.1. Steroid Receptor Co-Activator (Src)

Src is non-receptor protein TK. Src is an essential molecule required for normal osteoclast activity. However, over expression and activation of Src is also well associated with cancerous cell survival, development, adhesion and invasion, through various downstream signaling pathways such as Ras/mitogen-activated protein kinase pathway, STAT-3, FAK and c-Jun kinase. There by resulting in an overall aggressive and metastatic potential. High level of Src have been expressed in variety of cancers including OS [83]. Inhibition of Scr Phosphorylation in vitro induces pro-apoptotic activity and decreases cell invasion, migration, and adhesion. However, similar results could not be replicated using OS in vivo models, suggesting that Src activation may not be the primary pathway of pulmonary metastases in OS [84,85,86]. Dasatinib is a small-molecule tyrosine kinase inhibitor, directed against Src. Dasatinib is presently used in the treatment of chronic myeloid leukemia (CML) and Acute lymphoblastic leukemia (ALL). A clinical trial on pharmacokinetics and drug interaction of dasatinib showed no complete or partial responses, but maintained a stable disease condition in OS patient [8,87,88]. Two clinical trials to investigate either dasatinib alone or in combination are ongoing and awaiting results (Table 2). Saracatinib is a dual-specific inhibitor of Src and Abl. A few phase I trials of saracatinibin solid tumors have been conducted, but no OS patients were included in those studies [89,90]. Currently, Phase II studies of Src inhibitor saracatinibin recurrent pulmonary metastatic OS are being conducted (Table 2).

4.2. The Mammalian Target of Rapamycin (mTOR)

mTOR is a serine/threonine protein kinase, involved in PI3K/Akt signaling pathway and is responsible for regulation of protein synthesis, cell cycle and cell survival [96]. Activation of mTOR is attributed to be one of the key mechanisms in development and progression of several cancers including OS. Moreover, high level of mTOR in OS patients have been consistently correlated with poorer prognosis [97,98]. Indeed, suggesting mTOR as a potential therapeutic target.
The first generation of mTOR inhibitors are rapamycin (sirolimus) and rapamycin analogs like everolimus, temsirolimus, and ridaforolimus, which inhibit the kinase activityof mTORC1 through binding to FKBP-12 [99]. Preclinical testing of rapamycin monotherapy, demonstrated broad antitumor activity in in vivo as well as in OS xenograft models. Moreover, combination of rapamycin with cyclophosphamide or vincristine, enhanced the response in the OS models [100,101]. However, a clinical trial of sirolimus and cyclophosphamide in OS patients did not have a satisfactory response [91]. Clinical studies of everolimus in pediatric patients with refractory solid tumors (including 2 OS patient) had reported everolimus to be safe and had pharmacokinetic properties to those observed in adults, with stable disease observed in a single OS patient [92]. While the combination therapy of mTOR inhibitor everolimus and IGF-1R inhibitor figitumumab had no positive response in OS patients [93]. A non-randomized phase II trial on combination of sorafenib and everolimus in high-grade non-resectable osteosarcoma patients reported favorable antitumor activity as a 2nd or 3rd line of treatment, but could not attain the target of 6 months progression free survival in 50% patients [45]. Temsirolimus an analog of rapamycin, when used in combination with cisplatin or bevacizumab has demonstrated superior antitumor activity in comparison to their respective monotherapies in OS models [102]. Furthermore, in clinical setting, combination of temsirolimus with cixutumumab has shown positive clinical activity in patients with bone and soft-tissue sarcoma [40]. In a phase II trial of ridaforolimus, a selective mTOR inhibitor as a single agent therapy in patients with advance sarcoma, shows partial response in 2 OS patients (2/4 patients with osteosarcoma), while a larger double-blind phase III study has shown an increase in progression free survival (PFS) in patients with metastatic sarcoma after effective prior chemotherapy (50 bone sarcoma patients). However the study does not provide subset analysis for the bone sarcoma patients [103,104].
The second generation of dual mTOR inhibitors like, AZD8055, which inhibit mTOR1 as well as mTOR2, has demonstrated weak anti-tumor activity in OS xenograft. While, the effectiveness of other novel small-molecule inhibitors such as dual PI3K/mTOR inhibitors (NVP-BEZ235) and dual mTOR/DNA-PK inhibitor (CC-115) are still under clinical investigation [105,106]. Currently, several clinical trials of different types of mTOR inhibitors in treatment of OS are ongoing, either as a single agent therapy or as combinational therapy with other targeted agents (Table 2). These studies might give a better insight into the effectiveness of mTOR inhibitors in OS treatment.

4.3. Aurora Kinase

Loss of cell cycle control is a hallmark characteristic of cancer cells, thus leading to uncontrolled cell proliferation. Aurora kinase are serine/threonine protein kinases that regulate the normal cell mitosis, cell cycle transit from G2 and are potential oncogenic activators, as their abrupt expression or activation could cause normal cells to transform into cancerous cells [107]. In mammalian cells the aurora kinase gene encodes the three aurora kinase proteins: AURK-A, AURK-B, and AURK-C. Among these three types, AURK-A has been mostly involved in the tumorigenesis of several cancers including OS. Preclinical studies on silencing of AURK-A protein expression in osteosarcoma cells revealed that it could induce cell cycle arrest in G2/M and result in cell apoptosis in vitro [108,109]. Many aurora kinase inhibitors have been found and are being investigated in different stages of drug development.
MLN8237 (alisertib) is a second generation aurora kinase inhibitors, specifically targeting AURK-A via competing with ATP binding. Initial testing of MLN8237’s effectiveness, showed complete responses in one of six OS xenografts models [110]. A recent in vitro study has reported that MLN8237 through aurora kinase inhibition could induce autophagy and apoptotic cell death via PI3K/Akt/mTOR, p38 MAPK, and AMPK signaling pathway and that the reactive oxygen species (ROS) as well as sirtuin-1 pathways are also involved [111]. MLN8237 has been inducted in several clinical trials some of which are ongoing or awaiting results (Table 2).

5. Future Perspectives

5.1. Rank Inhibitors

OS is strongly associated with the receptor activator of nuclear factor-κB (RANK), along with its ligand (RANKL) and decoy osteoprotegerin (OPG). The molecular triad of RANK, RANKL and OPG play a key role in normal homeostasis between bone formation and bone loss, during bone remodeling. An imbalance in this triad has been associated with pathological bone remodeling and metabolism. Hence, the cancerous cells capitalize on the molecular triad of RANK, RANKL and OPG, thereby causing osteolytic or osteoblastic bone metastases [112,113]. An increased expression of RANK in human OS cell lines have been reported and the RANK activation further resisted anchorage-independent cell death [114]. However, inhibiting RANK expression in OS cells with shRNA, had reduced cell invasion and motility, while had limited effect on cell proliferation [115]. Additionally, RANKL has high expression in OS patients. A retrospective study reported two-thirds of OS biopsy specimens had positive staining for RANKL and this is poor prognosis in OS patients with high RANKL expression [116]. Furthermore, an in vivo study had demonstrated that the decoy osteoprotegerin (OPG) could indirectly prevent tumor-induced osteolysis and also inhibit tumor growth [117]. Hence, the molecular triad of RANK, RANKL and OPG could be a valuable therapeutic target in treating RANK-positive osteosarcoma.
Denosumab is a recombinant, humanized IgG 2 monoclonal antibody that is specifically targeted towards RANKL. Denosumab was initially introduced for treatment of osteoporosis as it prevents bone resorption by inhibiting differentiation of osteoclast precursors into mature osteoclasts. And it was later found to be beneficial in treatment of patients with skeletal-related events (SREs) from breast cancer and giant cell tumors of bone [118,119]. A phase II trial of denosumab has shown rapid and sustained effects in patients with giant cell tumor of bone. A randomized, double-blind study, presented superior efficacy of denosumab to zoledronic acid in delaying or preventing SREs in patients with breast cancer metastatic to the bones [120]. Given these results, further investigation in development of denosumab as treatment choice for patients suffering from high expression of RANKL in OS may be warranted.

5.2. Programmed Cell Death 1 (PD-1)/Programmed Cell Death Ligand 1 (PD-L1) Pathway

Recently, there has been interest in better understanding of cellular immune mechanism, especially about PD-1 and its ligand (PD-L1). PD-1 is a transmembrane protein, which is expressed on immune cells, activated T cells, B cells, natural killer (NK) T cells, activated monocytes, and dendritic cells. PD-1, functions as an immune checkpoint, by activation of PD-1/PD-L1 signaling pathway, promotes suppression of autoimmunity by preventing T-cells activation and by reducing cytokine production [121]. Over expression of PD-1 was found in several human cancers and its expression is associated with the disease progression as well as poor prognosis [122]. A recent report, revealed that programmed cell death ligand 1 (PD-L1) were highly expressed in OS patients, while another study reported that human metastatic tumors and not primary, OS tumors express PD-L1. This study indicated that blockade of PD-1/PD-L1 signaling pathway could drastically improve the functions of tumor-infiltrating lymphocytes, reduce the tumor burden and increase survival in OS mouse models [123]. Hence, based on supporting clinical evidence, targeting PD-1/PD-L1 interactions by anti-PD-1 antibody or PD-1 inhibitors may be a potential therapeutic strategy for the treatment of OS.

5.3. MicroRNAs

MicroRNAs, a class of 22-nucleotide small noncoding RNAs, are crucial for regulation of several pathophysiological processes and have been found to be involved in tumorigenesis, tumor progression and metastasis. It was found that miR-144 expression was significantly lower, while ROCK1 and ROCK2 levels were elevated in OS cell lines and OS tissues, further studies on restoration of miR-144 expression resulted in significant reduced cell proliferation, migration, invasion suppressed metastasis in vivo and in vitro, suggesting that miR-144 is a potent tumor suppressor by down regulating ROCK1 and ROCK1 expression in OS [124,125]. Thus, restoring miR-144 expression could be beneficial in treating osteosarcoma. Similarly, introduction of miR-132 suppressed OS cell proliferation in vitro, as well as in OS xenograft via cyclin E1 (CCNE1), while another study showed that miR-132 suppressed OS cell malignant behavior via Sox4 [126,127]. Elevated levels of miR-27a and overexpression of miR-27a/miR-27a pair have shown to suppress CBFA2T3 expression and thus contributes to development of pulmonary metastasis in OS [128,129]. MicroRNAs are also associated with the chemo resistance in OS. High expression of miR33a was found in OS cells that are resistant to cisplatin, through down regulation of TWIST and that inhibition of miR33a could increase the cell apoptosis induced by cisplatin [130] Therefore, microRNAs like miR-144, miR132, miR-27a, miR33a and may be able to serve as a diagnostic and therapeutic target for OS treatment.

5.4. Anti-Folate

Initially, high-dose methotrexate was considered as a standard treatment for osteosarcoma, but there is an increase in the number of cases of methotrexate resistant OS due to loss or reduced function of the reduced folate carrier protein and less retention of methotrexate within the cell, which led to the discovery of anti-folate agents like trimetrexate and pemetrexed. Trimetrexate is a better alternative to methotrexate for patients with osteosarcoma, as it does not rely on reduced folate carrier to enter into the cells. Also higher impaired methotrexate transport was found in OS patients samples, which is thought to have been caused due to defect in functionality of the reduced folate carrier protein [131]. Trimetrexate has demonstrated positive response in OS. A clinical study showed one complete response and one partial response in trimetrexate treated OS patients (seven evaluable patients) [132]. Pemetrexed is a multi-targeted anti-folate that inhibits several functional folate-dependent enzymes and DNA synthesis including thymidylate synthase (TS), dihydrofolate reductase and glycinamide ribonucleotide formyl transferase (GARFT) and is also dependent on reduced-folate carrier to be transported into the cells [133,134]. Pemetrexed has shown significant anti-tumor activity in OS cell lines but not superior to methotrexate [135]. In clinical setting, pemetrexed as a second line therapy in patients with advanced/metastatic OS has demonstrate partial response in one and stable disease in five patients (total 32 patients) suggesting insufficient minimal response [94]. Similar results were reported in another phase II study of pemetrexed in pediatric patients with refractory solid tumors including OS, with stable disease in one OS patient, with overall no observable response [95]. Different anti folate agents are still being tested in clinical trials (Table 2).

5.5. Improved Delivery Systems

A new therapeutic strategy in treating pulmonary metastasis in OS is to modify the delivery system of current chemotherapy agents so as to enhance local efficacy of chemotherapy agents. One such strategy is by liposomal encapsulation of drug and aerosolized delivery. Initial phase Ib/IIa study of the sustained release lipid inhalation targeting (SLIT) cisplatin in patients with OS metastatic to the lung has reported to be well tolerated and had pulmonary disease free in two patients (2/14 patients) at one year after start of therapy [136]. Further clinical study on efficacy of inhaled lipid cisplatin (ILC) in recurrent/progressive metastatic OS patients demonstrated partial response in one and stable disease in two patients (19 patients under evaluation) [137]. A phase II trial of ILC in patients with recurrent pulmonary OS is ongoing. Doxorubicin is one of the most active first-line chemotherapy agents used in the treatment of OS patients. Preclinical studies of liposomal encapsulation of doxorubicin has shown significant anti-tumor activity in vitro and in vivo [138]. Furthermore, a phase I/II clinical study of liposomal encapsulated doxorubicin have shown to be active against doxorubicin sensitive tumors as well as doxorubicin resistant tumor types [139]. The use of an aerosolized version of IL-2 could cause site-specific immunomodulation in OS patients with pulmonary malignancies and thus reduce drug-induced toxicities caused by the systemic administration. Phase I and II clinical trials (NCT01590069 *) are presently ongoing and the results of these studies will provide deep insights into the clinical benefits of these therapeutic approaches.

5.5.1. Doxorubicin Conjugate

Efforts are being made to tailor-fit active first-line chemotherapy agents, to improve their efficacy in treating drug resistant OS. Aldoxorubicin (INNO-206) a prodrug of doxorubicin (formerly known as DOXO-EMCH) has surfaced as a potential therapeutic agent against doxorubicin resistant tumors. INNO-206 is a (6-maleimidocaproyl) hydrazone derivative of doxorubicin. It is an albumin-binding prodrug with doxorubicin attached to an acid-sensitive linker (EMCH). Mechanistically, aldoxorubicin binds covalently to serum albumin, which predominantly accumulates in and around the cancerous cells. The acidic environment at tumor site causes cleavage of the acid-sensitive linker and release of free doxorubicin. A series of phase I studies on DOXO-EMCH showed it to be safe, stable in circulation without accumulating in body compartments with fewer cases of doxorubicin-associated cardiotoxicity and could induce tumor regressions in tumor types known to be anthracycline-sensitive tumors such as breast cancer, small cell lung cancer and sarcomas [140,141,142]. Aldoxorubicin is presently under investigation in different phases of clinical trial (NCT02235688 *). However, clinical efficacy of aldoxorubicin in treating metastatic OS patients has yet to be explored.

5.5.2. Nanoparticle Albumin-Bound (Nab) Paclitaxel

Paclitaxel (Taxol and Onxal) is cytotoxic microtubule stabilizing agent, against broad range of malignancies, including ovarian cancer, breast cancer, bladder cancer, head and neck cancer. However, its application in treating pediatric cancer patients has been limited due to its very low aqueous solubility, which requires paclitaxel to be formulated with Cremophor EL/ethanol as the solvent, which causes severe allergic, hypersensitivity and anaphylactic reactions. Nab-paclitaxel (Abraxane) is a novel, solvent-free, nanoparticle albumin-bound (nab) paclitaxel modified to reduce the toxicity of solvent-based paclitaxel. Nab-paclitaxel showed significant antitumor activity against most pediatric solid tumor cell lines including OS, while in OS xenograft model, nab-paclitaxel demonstrated significantly higher tumor inhibition (98.8%) compared to adriamycin (46.1%) and paclitaxel (40.8%). Further supporting the biologic rationale for clinical studies, nab-paclitaxel has shown to overcome paclitaxel resistance, mediated through ABCB1 observed in human OS cell lines [143,144,145]. Globally, nab-paclitaxel is approved for the treatment of metastatic breast cancer (MBC). At present, phase I/II clinical studies of nab-paclitaxel’s activity in pediatric patients with recurrent/refractory solid tumors including OS, are in the recruiting stage (NCT01962103 *).

6. Conclusions

Osteosarcoma is a highly heterogeneous bone cancer, which predominantly affect children and adolescence population. Current OS treatment regime consists of methotrexate, doxorubicin, cisplatin and ifosfamide, whose clinical benefits seem to have been maximized. Furthermore, strategies of intensifying chemotherapy dosing, varying its timing and introduction of multi-combinational chemotherapy has not helped to further improve the outcome of OS patients presenting with pulmonary metastasis and relapse. The identification of several new therapeutic agents, which specifically target immune system, cell proliferation, angiogenesis, extracellular and intercellular signaling pathways in tumor cells, either as single therapeutic agent or in combination with conventional chemotherapy has demonstrated promising data in preclinical and clinical trials. Indeed, a deeper insight into understanding genetic, molecular basis and tumorigenesis of OS, in addition to identification of biomarkers associated with poor prognosis and poor response to conventional treatment regimes, could pave the way to the discovery of novel molecular candidates that could be used in the development of potential therapeutic agents. In addition, the challenge ahead is to identify and harness these molecular therapeutic targets for multi-targeted combinational therapy or tailor-made personalized therapies or through modified drug delivery. This new approaches could help us explore the full potential of targeted therapies in OS.

Acknowledgments

We thank the other academic staff members in Aiping Lu and Ge Zhang’s group at Hong Kong Baptist University. We also thank Hong Kong Baptist University for providing critical comments and technical support. This study was supported by the Hong Kong General Research Fund (HKBU12102914 to Ge Zhang) and the Faculty Research Grant of Hong Kong Baptist University (FRG2/12-13/027 to Ge Zhang).

Author Contributions

Atik Badshah Shaikh, Fangfei Li and Min Li wrote the manuscript; Bing He contributed the figure designing along with literature research; Xiaojuan He, Guofen Chen, Baosheng Guo, Defang Li, Feng Jiang, Lei Dang, Shaowei Zheng, Chao Liang, Jin Liu, Cheng Lu, Biao Liu, Jun Lu, and Luyao Wang contributed to the manuscript for literature research; and Aiping Lu and Ge Zhang revised and approved the manuscript.

Conflicts of Interest

The authors declare no conflict of interest.

Abbreviations

OSOsteosarcoma
SRCSteroid receptor co-activator
PD-1Programmed cell death 1
PD-L1Programmed cell death ligand 1
RB1Retinoblastoma 1
TP53Tumor protein p53
APEX1Apurinic/Apyrimidinic exonuclease 1
BMPR2Bone morphogenetic protein type II receptor
HMGB1High mobility group box1
RANKLReceptor activator of nuclear factor-κB ligand
OPGOsteoprotegerin
RTKsReceptor tyrosine kinases
PDGFRPlatelet-derived growth factor receptor
VEGFRVascular endothelial growth factor receptor
IFNsInterferons
MAPMethotrexate, doxorubicin, and cisplatin
EFSEvent-free survival
GM-CSFGranulocyte-macrophage colony-stimulating factor
MDPMuramyl dipeptide
MTP-PEMuramyl tripeptide phosphatidyl ethanolamine
ILInterleukin
TNFTumor necrosis factor
PI3Phosphatidylinositol 3
ErkExtracellular signal regulated kinase
IGF-RInsulin-Like Growth Factor Receptor
IGFInsulin-Like Growth Factor
MAPKMitogen-activated protein kinase
PDGF/PDGFRPlatelet-Derived Growth Factor/receptors
RCCRenal cell carcinoma
GISTGastrointestinal stromal tumor
FLT-3Fms-related tyrosine kinase-3
HCCHepatocellular carcinoma
CMLChronic myeloid leukemia
ALLAcute lymphoblastic leukemia
AURKAurora Kinase
ROSReactive Oxygen Species
AMPK5′ Adenosine monophosphate-activated protein kinase
RANK/RANKLReceptor activator of nuclear factor-κB/Ligand
ROCK1Rho-associated, coiled-coil-containing protein kinase 1
NKNatural killer
nabNanoparticle albumin-bound
MBCMetastatic breast cancer
TSThymidylate synthase
GARFTGlycinamide ribonucleotide formyl transferase

References

  1. Damron, T.A.; Ward, W.G.; Stewart, A. Osteosarcoma, chondrosarcoma, and Ewing’s sarcoma: National cancer data base report. Clin. Orthop. Relat. Res. 2007, 459, 40–47. [Google Scholar] [CrossRef] [PubMed]
  2. Jones, K.B. Osteosarcomagenesis: Modeling cancer initiation in the mouse. Sarcoma 2011, 2011, 694136. [Google Scholar] [CrossRef] [PubMed]
  3. Saha, D.; Saha, K.; Banerjee, A.; Jash, D. Osteosarcoma relapse as pleural metastasis. South Asian J. Cancer 2013, 2, 56. [Google Scholar] [CrossRef] [PubMed]
  4. Allison, D.C.; Carney, S.C.; Ahlmann, E.R.; Hendifar, A.; Chawla, S.; Fedenko, A.; Angeles, C.; Menendez, L.R. A meta-analysis of osteosarcoma outcomes in the modern medical era. Sarcoma 2012, 2012. [Google Scholar] [CrossRef] [PubMed]
  5. Wu, P.K.; Chen, W.M.; Chen, C.F.; Lee, O.K.; Haung, C.K.; Chen, H.; Chen, T.H. Primary osteogenic sarcoma with pulmonary metastasis: Clinical results and prognostic factors in 91 patients. Jpn. J. Clin. Oncol. 2009, 39, 514–522. [Google Scholar] [CrossRef] [PubMed]
  6. Wachtel, M.; Schäfer, B.W. Targets for cancer therapy in childhood sarcomas. Cancer Treat. Rev. 2010, 36, 318–327. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  7. Lu, C.; Ding, L.; Mardis, E.E.R.; Wilson, R.K.R.; Chen, X.; Bahrami, A.; Pappo, A.; Easton, J.; Dalton, J.; Hedlund, E.; et al. Recurrent somatic structural variations contribute to tumorigenesis in pediatric osteosarcoma. Cell. Rep. 2014, 7, 104–112. [Google Scholar]
  8. Sampson, V.B.; Gorlick, R.; Kamara, D.; Anders Kolb, E. A review of targeted therapies evaluated by the pediatric preclinical testing program for osteosarcoma. Front. Oncol. 2013, 3, 132. [Google Scholar] [CrossRef] [PubMed]
  9. Kaya, M. Prognostic Factors for Osteosarcoma Patients. In Osteosarcoma; Springer Japan: Tokyo, Japan, 2016; pp. 73–79. [Google Scholar]
  10. Morrow, J.; Khanna, C. Osteosarcoma genetics and epigenetics: Emerging biology and candidate therapies. Crit. Rev. Oncog. 2015, 20, 173–197. [Google Scholar] [CrossRef] [PubMed]
  11. Kansara, M.; Leong, H.S.; Lin, D.M.; Popkiss, S.; Pang, P.; Garsed, D.W.; Walkley, C.R.; Cullinane, C.; Ellul, J.; Haynes, N.M.; et al. Immune response to RB1-regulated senescence limits radiation-induced osteosarcoma formation. J. Clin. Investig. 2013, 123, 5351–5360. [Google Scholar] [CrossRef] [PubMed]
  12. Chandar, N.; Billig, B.; McMaster, J.; Novak, J. Inactivation of p53 gene in human and murine osteosarcoma cells. Br. J. Cancer 1992, 65, 208–214. [Google Scholar] [CrossRef] [PubMed]
  13. Yang, J.; Yang, D.; Cogdell, D.; Du, X. APEX1 gene amplification and its protein overexpression in osteosarcoma: Correlation with recurrence, metastasis, and survival. Technol. Cancer Res. Treat. 2010, 9, 161–169. [Google Scholar] [CrossRef] [PubMed]
  14. Shimizu, T.; Ishikawa, T.; Sugihara, E.; Kuninaka, S.; Miyamoto, T.; Mabuchi, Y.; Matsuzaki, Y.; Tsunoda, T.; Miya, F.; Morioka, H.; 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, 5687–5699. [Google Scholar] [CrossRef] [PubMed]
  15. Hattinger, C.M.; Stoico, G.; Michelacci, F.; Pasello, M.; Scionti, I.; Remondini, D.; Castellani, G.C.; Fanelli, M.; Scotlandi, K.; Picci, P.; et al. Mechanisms of gene amplification and evidence of coamplification in drug-resistant human osteosarcoma cell lines. Genes Chromosomes Cancer 2009, 48, 289–309. [Google Scholar] [CrossRef] [PubMed]
  16. Abdou, A.G.; Kandil, M.; Asaad, N.Y.; Dawoud, M.M.; Shahin, A.A.; Abd Eldayem, A.F. The prognostic role of Ezrin and HER2/neu expression in osteosarcoma. Appl. Immunohistochem. Mol. Morphol. 2015, 9, 10. [Google Scholar] [CrossRef] [PubMed]
  17. Meng, Q.; Zhao, J.; Liu, H.; Zhou, G.; Zhang, W.; Xu, X.; Zheng, M. HMGB1 promotes cellular proliferation and invasion, suppresses cellular apoptosis in osteosarcoma. Tumor Biol. 2014, 35, 12265–12274. [Google Scholar] [CrossRef] [PubMed]
  18. Wang, L.; Park, P.; La Marca, F.; Than, K.; Rahman, S.; Lin, C.Y. Bone formation induced by BMP-2 in human osteosarcoma cells. Int. J. Oncol. 2013, 43, 1095–1102. [Google Scholar] [PubMed]
  19. Tarhini, A.A.; Kirkwood, J.M. How much of a good thing? What duration for interferon alfa-2b adjuvant therapy? J. Clin. Oncol. 2012, 30, 3773–3776. [Google Scholar] [CrossRef] [PubMed]
  20. Pollack, S.M.; Loggers, E.T.; Rodler, E.T.; Yee, C.; Jones, R.L. Immune-based therapies for sarcoma. Sarcoma 2011, 2011, 438940. [Google Scholar] [CrossRef] [PubMed]
  21. Müller, C.R.; Smeland, S.; Bauer, H.C.F.; Saeter, G.; Strander, H. Interferon-alpha as the only adjuvant treatment in high-grade osteosarcoma: Long term results of the Karolinska Hospital series. Acta Oncol. 2005, 44, 475–480. [Google Scholar] [CrossRef] [PubMed]
  22. Bielack, S.S.; Smeland, S.; Whelan, J.; Marina, N.; Hook, J.; Jovic, G.; Krailo, M.D.; Butterfass-Bahloul, T.; Kuhne, T.; Eriksson, M.; et al. MAP plus maintenance pegylated interferon α-2b (MAPIfn) versus MAP alone in patients with resectable high-grade osteosarcoma and good histologic response to preoperative MAP: First results of the EURAMOS-1 “good response” randomization. ASCO Meet. Abstr. 2013, 31, LBA10504. [Google Scholar]
  23. Kaufman, H.; Ruby, C.; Hughes, T.; Slingluff, C. Current status of granulocyte-macrophage colony-stimulating factor in the immunotherapy of melanoma. J. Immunother. Cancer 2014, 2, 11. [Google Scholar] [CrossRef] [PubMed]
  24. Takahashi, H.; Kato, M.; Kikuchi, A.; Hanada, R.; Koh, K. Delayed short-term administration of granulocyte colony-stimulating factor is a good mobilization strategy for harvesting autologous peripheral blood stem cells in pediatric patients with solid tumors. Pediatr. Transplant. 2013, 17, 688–693. [Google Scholar] [CrossRef] [PubMed]
  25. Anderson, P.M.; Markovic, S.N.; Sloan, J.A.; Clawson, M.L.; Wylam, M.; Arndt, C.A.S.; Smithson, W.A.; Burch, P.; Gornet, M.; Rahman, E. Aerosol granulocyte macrophage-colony stimulating factor: A low toxicity, lung-specific biological therapy in patients with lung metastases. Clin. Cancer Res. 1999, 5, 2316–2323. [Google Scholar] [PubMed]
  26. Arndt, C.A.S.; Koshkina, N.V.; Inwards, C.Y.; Hawkins, D.S.; Krailo, M.D.; Villaluna, D.; Anderson, P.M.; Goorin, A.M.; Blakely, M.L.; Bernstein, M.; et al. Inhaled granulocyte-macrophage colony stimulating factor for first pulmonary recurrence of osteosarcoma: Effects on disease-free survival and immunomodulation. a report from the Children’s Oncology Group. Clin. Cancer Res. 2010, 16, 4024–4030. [Google Scholar] [CrossRef] [PubMed]
  27. Ando, K.; Heymann, M.-F.; Stresing, V.; Mori, K.; Rédini, F.; Heymann, D. Current therapeutic strategies and novel approaches in osteosarcoma. Cancers 2013, 5, 591–616. [Google Scholar] [CrossRef] [PubMed]
  28. Frampton, J.E. Mifamurtide: A review of its use in the treatment of osteosarcoma. Pediatr. Drugs 2010, 12, 141–153. [Google Scholar] [CrossRef] [PubMed]
  29. Anderson, P.M.; Tomaras, M.; McConnell, K. Mifamurtide in osteosarcoma—A practical review. Drugs Today 2010, 46, 327–337. [Google Scholar] [CrossRef] [PubMed]
  30. MacEwen, E.G.; Kurzman, I.D.; Rosenthal, R.C.; Smith, B.W.; Manley, P.A.; Roush, J.K.; Howard, P.E. Therapy for osteosarcoma in dogs with intravenous injection of liposome-encapsulated muramyl tripeptide. J. Natl. Cancer Inst. 1989, 81, 935–938. [Google Scholar] [CrossRef] [PubMed]
  31. Meyers, P.A.; Schwartz, C.L.; Krailo, M.D.; Healey, J.H.; Bernstein, M.L.; Betcher, D.; Ferguson, W.S.; Gebhardt, M.C.; Goorin, A.M.; Harris, M.; et al. Osteosarcoma: The addition of muramyl tripeptide to chemotherapy improves overall survival—A report from the Children’s Oncology Group. J. Clin. Oncol. 2008, 26, 633–638. [Google Scholar] [CrossRef] [PubMed]
  32. Chou, A.J.; Kleinerman, E.S.; Krailo, M.D.; Chen, Z.; Betcher, D.L.; Healey, J.H.; Conrad, E.U.; Nieder, M.L.; Weiner, M.A.; Wells, R.J.; et al. Addition of muramyl tripeptide to chemotherapy for patients with newly diagnosed metastatic osteosarcoma. Cancer 2009, 115, 5339–5348. [Google Scholar] [CrossRef] [PubMed]
  33. Anderson, P.M.; Meyers, P.; Kleinerman, E.; Venkatakrishnan, K.; Hughes, D.P.; Herzog, C.; Huh, W.; Sutphin, R.; Vyas, Y.M.; Shen, V.; et al. Mifamurtide in metastatic and recurrent osteosarcoma: A patient access study with pharmacokinetic, pharmacodynamic, and safety assessments. Pediatr. Blood Cancer 2014, 61, 238–244. [Google Scholar] [CrossRef] [PubMed]
  34. Huh, W.; Egas-Bejar, D.; Anderson, P. New treatment options for nonmetastatic osteosarcoma: Focus on mifamurtide in adolescents. Clin. Oncol. Adolesc. Young Adults 2012, 2012, 1–6. [Google Scholar] [CrossRef]
  35. Fleuren, E.D.G.; Versleijen-Jonkers, Y.M.H.; Boerman, O.C.; van der Graaf, W.T.A. Targeting receptor tyrosine kinases in osteosarcoma and Ewing sarcoma: Current hurdles and future perspectives. Biochim. Biophys. Acta 2014, 1845, 266–276. [Google Scholar] [CrossRef] [PubMed]
  36. Lemmon, M.A.; Schlessinger, J. Cell signaling by receptor tyrosine kinases. Cell. 2010, 141, 1117–1134. [Google Scholar] [CrossRef] [PubMed]
  37. Malempati, S.; Weigel, B.; Ingle, A.M.; Ahern, C.H.; Carroll, J.M.; Roberts, C.T.; Reid, J.M.; Schmechel, S.; Voss, S.D.; Cho, S.Y.; et al. Phase I/II trial and pharmacokinetic study of cixutumumab in pediatric patients with refractory solid tumors and Ewing sarcoma: A report from the Children’s Oncology Group. J. Clin. Oncol. 2012, 30, 256–262. [Google Scholar] [CrossRef] [PubMed]
  38. Weigel, B.; Malempati, S.; Reid, J.M.; Voss, S.D.; Cho, S.Y.; Chen, H.X.; Krailo, M.; Villaluna, D.; Adamson, P.C.; Blaney, S.M. Phase 2 trial of cixutumumab in children, adolescents, and young adults with refractory solid tumors: A report from the Children’s Oncology Group. Pediatr. Blood Cancer 2014, 61, 452–456. [Google Scholar] [CrossRef] [PubMed]
  39. Schwartz, G.K.; Tap, W.D.; Qin, L.-X.; Livingston, M.B.; Undevia, S.D.; Chmielowski, B.; Agulnik, M.; Schuetze, S.; Reed, D.R.; Okuno, S.H.; et al. A phase II multicenter study of the IGF-1 receptor antibody cixutumumab (A12) and the mTOR inhibitor temsirolimus (TEM) in patients (pts) with refractory IGF-1R positive (+) and negative (–) bone and soft tissue sarcomas (STS). ASCO Meet. Abstr. 2012, 30, 10003. [Google Scholar]
  40. Schwartz, G.K.; Tap, W.D.; Qin, L.-X.; Livingston, M.B.; Undevia, S.D.; Chmielowski, B.; Agulnik, M.; Schuetze, S.M.; Reed, D.R.; Okuno, S.H.; et al. Cixutumumab and temsirolimus for patients with bone and soft-tissue sarcoma: A multicentre, open-label, phase 2 trial. Lancet Oncol. 2013, 14, 371–382. [Google Scholar] [CrossRef]
  41. Wagner, L.M.; Fouladi, M.; Ahmed, A.; Krailo, M.D.; Weigel, B.; DuBois, S.G.; Doyle, L.A.; Chen, H.; Blaney, S.M. Phase II study of cixutumumab in combination with temsirolimus in pediatric patients and young adults with recurrent or refractory sarcoma: A report from the Children’s Oncology Group. Pediatr. Blood Cancer 2015, 62, 440–444. [Google Scholar] [CrossRef] [PubMed]
  42. Bond, M.; Bernstein, M.L.; Pappo, A.; Schultz, K.R.; Krailo, M.; Blaney, S.M.; Adamson, P.C. A phase II study of imatinib mesylate in children with refractory or relapsed solid tumors: A Children’s Oncology Group study. Pediatr. Blood Cancer 2008, 50, 254–258. [Google Scholar] [CrossRef] [PubMed]
  43. Dubois, S.G.; Shusterman, S.; Ingle, A.M.; Ahern, C.H.; Reid, J.M.; Wu, B.; Baruchel, S.; Glade-Bender, J.; Ivy, P.; Grier, H.E.; et al. Phase I and pharmacokinetic study of sunitinib in pediatric patients with refractory solid tumors: A children’s oncology group study. Clin. Cancer Res. 2011, 17, 5113–5122. [Google Scholar] [CrossRef] [PubMed]
  44. Glade Bender, J.L.; Adamson, P.C.; Reid, J.M.; Xu, L.; Baruchel, S.; Shaked, Y.; Kerbel, R.S.; Cooney-Qualter, E.M.; Stempak, D.; Chen, H.X.; et al. Phase I trial and pharmacokinetic study of bevacizumab in pediatric patients with refractory solid tumors: A Children’s Oncology Group Study. J. Clin. Oncol. 2008, 26, 399–405. [Google Scholar] [CrossRef] [PubMed]
  45. Grignani, G.; Palmerini, E.; Ferraresi, V.; D’Ambrosio, L.; Bertulli, R.; Asaftei, S.D.; Tamburini, A.; Pignochino, Y.; Sangiolo, D.; Marchesi, E.; et al. Sorafenib and everolimus for patients with unresectable high-grade osteosarcoma progressing after standard treatment: A non-randomised phase 2 clinical trial. Lancet Oncol. 2015, 16, 98–107. [Google Scholar] [CrossRef]
  46. Ebb, D.; Meyers, P.; Grier, H.; Bernstein, M.; Gorlick, R.; Lipshultz, S.E.; Krailo, M.; Devidas, M.; Barkauskas, D.A.; Siegal, G.P.; et al. Phase II trial of trastuzumab in combination with cytotoxic chemotherapy for treatment of metastatic osteosarcoma with human epidermal growth factor receptor 2 overexpression: A report from the children’s oncology group. J. Clin. Oncol. 2012, 30, 2545–2551. [Google Scholar] [CrossRef] [PubMed]
  47. ClinicalTrials.gov. Bethesda (MD): National Library of Medicine (US). Available online: http://clinicaltrials.gov (accessed on 2 February 2016).
  48. Hassan, S.E.; Bekarev, M.; Kim, M.Y.; Lin, J.; Piperdi, S.; Gorlick, R.; Geller, D.S. Cell surface receptor expression patterns in osteosarcoma. Cancer 2012, 118, 740–749. [Google Scholar] [CrossRef] [PubMed]
  49. Pollak, M. The insulin and insulin-like growth factor receptor family in neoplasia: An update. Nat. Rev. Cancer 2012, 12, 159–169. [Google Scholar] [CrossRef] [PubMed]
  50. Tognon, C.E.; Sorensen, P.H.B. Targeting the insulin-like growth factor 1 receptor (IGF1R) signaling pathway for cancer therapy. Expert Opin. Ther. Targets 2012, 16, 33–48. [Google Scholar] [CrossRef] [PubMed]
  51. Kuijjer, M.L.; Peterse, E.F.P.; van den Akker, B.E.W.M.; Briaire-de Bruijn, I.H.; Serra, M.; Meza-Zepeda, L.A.; Myklebost, O.; Hassan, A.B.; Hogendoorn, P.C.W.; et al. IR/IGF1R signaling as potential target for treatment of high-grade osteosarcoma. BMC Cancer 2013, 13, 245. [Google Scholar] [CrossRef] [PubMed]
  52. Wang, Y.-H.; Han, X.-D.; Qiu, Y.; Xiong, J.; Yu, Y.; Wang, B.; Zhu, Z.-Z.; Qian, B.-P.; Chen, Y.-X.; Wang, S.-F.; et al. Increased expression of insulin-like growth factor-1 receptor is correlated with tumor metastasis and prognosis in patients with osteosarcoma. J. Surg. Oncol. 2012, 105, 235–243. [Google Scholar] [CrossRef] [PubMed]
  53. Burrow, S.; Andrulis, I.L.; Pollak, M.; Bell, R.S. Expression of insulin-like growth factor receptor, IGF-1, and IGF-2 in primary and metastatic osteosarcoma. J. Surg. Oncol. 1998, 69, 21–27. [Google Scholar] [CrossRef]
  54. Wang, Y.L.Y.; Lipari, P.; Wang, X.Y.; Hailey, J.; Liang, L.Z.; Ramos, R.; Liu, M.; Pachter, J.A.; Bishop, W.R.; Wang, Y. A fully human insulin-like growth factor-I receptor antibody SCH 717454 (Robatumumab) has antitumor activity as a single agent and in combination with cytotoxics in pediatric tumor xenografts. Mol. Cancer Ther. 2010, 9, 410–418. [Google Scholar] [CrossRef] [PubMed]
  55. Friedbichler, K.; Hofmann, M.H.; Kroez, M.; Ostermann, E.; Lamche, H.R.; Koessl, C.; Borges, E.; Pollak, M.N.; Adolf, G.; Adam, P.J. Pharmacodynamic and antineoplastic activity of BI 836845, a fully human IGF ligand-neutralizing antibody, and mechanistic rationale for combination with rapamycin. Mol. Cancer Ther. 2014, 13, 399–409. [Google Scholar] [CrossRef] [PubMed]
  56. Gao, J.; Chesebrough, J.W.; Cartlidge, S.A.; Ricketts, S.A.; Incognito, L.; Veldman-Jones, M.; Blakey, D.C.; Tabrizi, M.; Jallal, B.; Trail, P.A.; et al. Dual IGF-I/II-neutralizing antibody MEDI-573 potently inhibits IGF signaling and tumor growth. Cancer Res. 2011, 71, 1029–1040. [Google Scholar] [CrossRef] [PubMed]
  57. Kolb, E.A.; Gorlick, R.; Lock, R.; Carol, H.; Morton, C.L.; Keir, S.T.; Reynolds, C.P.; Kang, M.H.; Maris, J.M.; Billups, C.; et al. Initial testing (stage 1) of the IGF-1 receptor inhibitor BMS-754807 by the pediatric preclinical testing program. Pediatr. Blood Cancer 2011, 56, 595–603. [Google Scholar] [CrossRef] [PubMed]
  58. Zhang, L.; Leeman, E.; Carnes, D.C.; Graves, D.T. Human osteoblasts synthesize and respond to platelet-derived growth factor. Am. J. Physiol. 1991, 261, C348–C354. [Google Scholar] [PubMed]
  59. Sulzbacher, I.; Birner, P.; Trieb, K.; Träxler, M.; Lang, S.; Chott, A. Expression of platelet-derived growth factor-AA is associated with tumor progression in osteosarcoma. Mod. Pathol. 2003, 16, 66–71. [Google Scholar] [CrossRef] [PubMed]
  60. Bozzi, F.; Tamborini, E.; Negri, T.; Pastore, E.; Ferrari, A.; Luksch, R.; Casanova, M.; Pierotti, M.A.; Bellani, F.F.; Pilotti, S. Evidence for activation of KIT, PDGFRα, and PDGFRβ receptors in the Ewing sarcoma family of tumors. Cancer 2007, 109, 1638–1645. [Google Scholar] [CrossRef] [PubMed]
  61. Sulzbacher, I.; Birner, P.; Dominkus, M.; Pichlhofer, B.; Mazal, P.R. Expression of platelet-derived growth factor-alpha receptor in human osteosarcoma is not a predictor of outcome. Pathology 2010, 42, 664–668. [Google Scholar] [CrossRef] [PubMed]
  62. Takagi, S.; Takemoto, A.; Takami, M.; Oh-Hara, T.; Fujita, N. Platelets promote osteosarcoma cell growth through activation of the platelet-derived growth factor receptor-Akt signaling axis. Cancer Sci. 2014, 105, 983–988. [Google Scholar] [CrossRef] [PubMed]
  63. McGary, E.C.; Weber, K.; Mills, L.; Doucet, M.; Lewis, V.; Lev, D.C.; Fidler, I.J.; Bar-Eli, M. Inhibition of platelet-derived growth factor-mediated proliferation of osteosarcoma cells by the novel tyrosine kinase inhibitor STI. Clin. Cancer Res. 2002, 8, 3584–3591. [Google Scholar]
  64. Kubo, T.; Piperdi, S.; Rosenblum, J.; Antonescu, C.R.; Chen, W.; Kim, H.S.; Huvos, A.G.; Sowers, R.; Meyers, P.A.; Healey, J.H.; et al. Platelet-derived growth factor receptor as a prognostic marker and a therapeutic target for imatinib mesylate therapy in osteosarcoma. Cancer 2008, 112, 2119–2129. [Google Scholar] [CrossRef] [PubMed]
  65. Gobin, B.; Moriceau, G.; Ory, B.; Charrier, C.; Brion, R.; Blanchard, F.; Redini, F.; Heymann, D. Imatinib mesylate exerts anti-proliferative effects on osteosarcoma cells and inhibits the tumour growth in immunocompetent murine models. PLoS ONE 2014, 9, e90795. [Google Scholar] [CrossRef] [PubMed]
  66. Maris, J.M.; Courtright, J.; Houghton, P.J.; Morton, C.L.; Kolb, E.A.; Lock, R.; Tajbakhsh, M.; Reynolds, C.P.; Keir, S.T.; Wu, J.; et al. Initial testing (stage 1) of sunitinib by the pediatric preclinical testing program. Pediatr. Blood Cancer 2008, 51, 42–48. [Google Scholar] [CrossRef] [PubMed]
  67. Kumar, R.M.R.; Arlt, M.J.; Kuzmanov, A.; Born, W.; Fuchs, B. Sunitinib malate (SU-11248) reduces tumour burden and lung metastasis in an intratibial human xenograft osteosarcoma mouse model. Am. J. Cancer Res. 2015, 5, 2156–2168. [Google Scholar] [PubMed]
  68. Rivera-Valentin, R.K.; Zhu, L.; Hughes, D.P.M. Bone sarcomas in pediatrics: Progress in our understanding of tumor biology and implications for therapy. Pediatr. Drugs 2015, 17, 257–271. [Google Scholar] [CrossRef] [PubMed]
  69. Gidwani, P.; Zhang, W.; Gorlick, R.; Kolb, E. Activity of dasatinib in human osteosarcoma. Cancer Res. 2007, 67, 3255. [Google Scholar]
  70. Guan, H.; Zhou, Z.; Wang, H.; Jia, S.-F.; Liu, W.; Kleinerman, E.S. A small interfering RNA targeting vascular endothelial growth factor inhibits Ewing’s sarcoma growth in a xenograft mouse model. Clin. Cancer Res. 2005, 11, 2662–2669. [Google Scholar] [CrossRef] [PubMed]
  71. Ługowska, I.; Woźniak, W.; Klepacka, T.; Michalak, E.; Szamotulska, K. A prognostic evaluation of vascular endothelial growth factor in children and young adults with osteosarcoma. Pediatr. Blood Cancer 2011, 57, 63–68. [Google Scholar] [CrossRef] [PubMed]
  72. Yu, X.-W.; Wu, T.-Y.; Yi, X.; Ren, W.-P.; Zhou, Z.-B.; Sun, Y.-Q.; Zhang, C.-Q. Prognostic significance of VEGF expression in osteosarcoma: A meta-analysis. Tumour Biol. 2014, 35, 155–160. [Google Scholar] [CrossRef] [PubMed]
  73. Niswander, L.M.; Kim, S.Y. Stratifying osteosarcoma: Minimizing and maximizing therapy. Curr. Oncol. Rep. 2010, 12, 266–270. [Google Scholar] [CrossRef] [PubMed]
  74. Scharf, V.F.; Farese, J.P.; Coomer, A.R.; Milner, R.J.; Taylor, D.P.; Salute, M.E.; Chang, M.N.; Neal, D.; Siemann, D.W. Effect of bevacizumab on angiogenesis and growth of canine osteosarcoma cells xenografted in athymic mice. Am. J. Vet. Res. 2013, 74, 771–778. [Google Scholar] [CrossRef] [PubMed]
  75. Keir, S.T.; Maris, J.M.; Lock, R.; Kolb, E.A.; Gorlick, R.; Carol, H.; Morton, C.L.; Reynolds, C.P.; Kang, M.H.; Watkins, A.; et al. Initial testing (stage 1) of the multi-targeted kinase inhibitor sorafenib by the pediatric preclinical testing program. Pediatr. Blood Cancer 2010, 55, 1126–1133. [Google Scholar] [CrossRef] [PubMed]
  76. Pignochino, Y.; Grignani, G.; Cavalloni, G.; Motta, M.; Tapparo, M.; Bruno, S.; Bottos, A.; Gammaitoni, L.; Migliardi, G.; Camussi, G.; et al. Sorafenib blocks tumour growth, angiogenesis and metastatic potential in preclinical models of osteosarcoma through a mechanism potentially involving the inhibition of ERK1/2, MCL-1 and ezrin pathways. Mol. Cancer 2009, 8, 118. [Google Scholar] [CrossRef] [PubMed]
  77. Martin, V.; Cappuzzo, F.; Mazzucchelli, L.; Frattini, M. HER2 in solid tumors: More than 10 years under the microscope; where are we now? Future Oncol. 2014, 10, 1469–1486. [Google Scholar] [CrossRef] [PubMed]
  78. Gorlick, R.; Huvos, A.G.; Heller, G.; Aledo, A.; Beardsley, G.P.; Healey, J.H.; Meyers, P.A. Expression of HER2/erbB-2 correlates with survival in osteosarcoma. J. Clin. Oncol. 1999, 17, 2781. [Google Scholar] [PubMed]
  79. Scotlandi, K.; Manara, M.C.; Hattinger, C.M.; Benini, S.; Perdichizzi, S.; Pasello, M.; Bacci, G.; Zanella, L.; Bertoni, F.; Picci, P.; et al. Prognostic and therapeutic relevance of HER2 expression in osteosarcoma and Ewing’s sarcoma. Eur. J. Cancer 2005, 41, 1349–1361. [Google Scholar] [CrossRef] [PubMed]
  80. Kilpatrick, S.E.; Geisinger, K.R.; King, T.S.; Sciarrotta, J.; Ward, W.G.; Gold, S.H.; Bos, G.D. Clinicopathologic analysis of HER-2/neu immunoexpression among various histologic subtypes and grades of osteosarcoma. Mod. Pathol. 2001, 14, 1277–1283. [Google Scholar] [CrossRef] [PubMed]
  81. Thomas, D.G.; Giordano, T.J.; Sanders, D.; Biermann, J.S.; Baker, L. Absence of HER2/neu gene expression in osteosarcoma and skeletal Ewing’s sarcoma. Clin. Cancer Res. 2002, 8, 788–793. [Google Scholar] [PubMed]
  82. Gill, J.; Ahluwalia, M.K.; Geller, D.; Gorlick, R. New targets and approaches in osteosarcoma. Pharmacol. Ther. 2013, 137, 89–99. [Google Scholar] [CrossRef] [PubMed]
  83. Hu, C.; Deng, Z.; Zhang, Y.; Yan, L.; Cai, L.; Lei, J.; Xie, Y. The prognostic significance of Src and p-Src expression in patients with osteosarcoma. Med. Sci. Monit. 2015, 21, 638–645. [Google Scholar] [PubMed]
  84. Akiyama, T.; Dass, C.R.; Choong, P.F.M. Novel therapeutic strategy for osteosarcoma targeting osteoclast differentiation, bone-resorbing activity, and apoptosis pathway. Mol. Cancer Ther. 2008, 7, 3461–3469. [Google Scholar] [CrossRef] [PubMed]
  85. Hingorani, P.; Zhang, W.; Gorlick, R.; Kolb, E.A. Inhibition of Src phosphorylation alters metastatic potential of osteosarcoma in vitro but not in vivo. Clin. Cancer Res. 2009, 15, 3416–3422. [Google Scholar] [CrossRef] [PubMed]
  86. Díaz-Montero, C.M.; Wygant, J.N.; McIntyre, B.W. PI3-K/Akt-mediated anoikis resistance of human osteosarcoma cells requires Src activation. Eur. J. Cancer 2006, 42, 1491–1500. [Google Scholar] [CrossRef] [PubMed]
  87. Johnson, F.M.; Agrawal, S.; Burris, H.; Rosen, L.; Dhillon, N.; Hong, D.; Blackwood-Chirchir, A.; Luo, F.R.; Sy, O.; Kaul, S.; et al. A Phase 1 pharmacokinetic and drug-interaction study of dasatinib in patients with advanced solid tumors. Cancer 2010, 116, 1582–1591. [Google Scholar] [CrossRef]
  88. Schuetze, S.; Wathen, K.; Choy, E.; Samuels, B.L.; Ganjoo, K.N.; Staddon, A.P.; von Mehren, M.; Chow, W.A.; Trent, J.C.; Baker, L.H. Results of a Sarcoma Alliance for Research through Collaboration (SARC) phase II trial of dasatinib in previously treated, high-grade, advanced sarcoma. ASCO Meet. Abstr. 2010, 28, 10009. [Google Scholar]
  89. Baselga, J.; Cervantes, A.; Martinelli, E.; Chirivella, I.; Hoekman, K.; Hurwitz, H.I.; Jodrell, D.I.; Hamberg, P.; Casado, E.; Elvin, P.; et al. Phase I safety, pharmacokinetics, and inhibition of SRC activity study of saracatinib in patients with solid tumors. Clin. Cancer Res. 2010, 16, 4876–4883. [Google Scholar] [CrossRef] [PubMed]
  90. Fujisaka, Y.; Onozawa, Y.; Kurata, T.; Yasui, H.; Goto, I.; Yamazaki, K.; Machida, N.; Watanabe, J.; Shimada, H.; Shi, X.; et al. First report of the safety, tolerability, and pharmacokinetics of the Src kinase inhibitor saracatinib (AZD0530) in Japanese patients with advanced solid tumours. Investig. New Drugs 2013, 31, 108–114. [Google Scholar] [CrossRef] [PubMed]
  91. Schuetze, S.M.; Zhao, L.; Chugh, R.; Thomas, D.G.; Lucas, D.R.; Metko, G.; Zalupski, M.M.; Baker, L.H. Results of a phase II study of sirolimus and cyclophosphamide in patients with advanced sarcoma. Eur. J. Cancer 2012, 48, 1347–1353. [Google Scholar] [CrossRef] [PubMed]
  92. Fouladi, M.; Laningham, F.; Wu, J.; O’Shaughnessy, M.A.; Molina, K.; Broniscer, A.; Spunt, S.L.; Luckett, I.; Stewart, C.F.; Houghton, P.J.; et al. Phase I study of everolimus in pediatric patients with refractory solid tumors. J. Clin. Oncol. 2007, 25, 4806–4812. [Google Scholar] [CrossRef] [PubMed]
  93. Quek, R.; Wang, Q.; Morgan, J.A.; Shapiro, G.I.; Butrynski, J.E.; Ramaiya, N.; Huftalen, T.; Jederlinic, N.; Manola, J.; Wagner, A.J.; et al. Combination mTOR and IGF-1R inhibition: Phase I trial of everolimus and figitumumab in patients with advanced sarcomas and other solid tumors. Clin. Cancer Res. 2011, 17, 871–879. [Google Scholar] [CrossRef] [PubMed]
  94. Duffaud, F.; Egerer, G.; Ferrari, S.; Rassam, H.; Boecker, U.; Bui-Nguyen, B. A phase II trial of second-line pemetrexed in adults with advanced/metastatic osteosarcoma. Eur. J. Cancer 2012, 48, 564–570. [Google Scholar] [CrossRef] [PubMed]
  95. Warwick, A.B.; Malempati, S.; Krailo, M.; Melemed, A.; Gorlick, R.; Ames, M.M.; Safgren, S.L.; Adamson, P.C.; Blaney, S.M. Phase 2 trial of pemetrexed in children and adolescents with refractory solid tumors: A Children’s Oncology Group study. Pediatr. Blood Cancer 2013, 60, 237–241. [Google Scholar] [CrossRef] [PubMed]
  96. Ory, B.; Moriceau, G.; Redini, F.; Heymann, D. mTOR inhibitors (rapamycin and its derivatives) and nitrogen containing bisphosphonates: Bi-functional compounds for the treatment of bone tumours. Curr. Med. Chem. 2007, 14, 1381–1387. [Google Scholar] [CrossRef] [PubMed]
  97. Perry, J.A.; Kiezun, A.; Tonzi, P.; van Allen, E.M.; Carter, S.L.; Baca, S.C.; Cowley, G.S.; Bhatt, A.S.; Rheinbay, E.; Pedamallu, C.S.; et al. A Complementary genomic approaches highlight the PI3K/mTOR pathway as a common vulnerability in osteosarcoma. Proc. Natl. Acad. Sci. USA. 2014, 111, E5564–E5573. [Google Scholar] [CrossRef] [PubMed]
  98. Zhou, Q.; Deng, Z.; Zhu, Y.; Long, H.; Zhang, S.; Zhao, J. mTOR/p70S6K signal transduction pathway contributes to osteosarcoma progression and patients’ prognosis. Med. Oncol. 2010, 27, 1239–1245. [Google Scholar] [CrossRef] [PubMed]
  99. Meng, L.-H.; Zheng, X.F.S. Toward rapamycin analog (rapalog)-based precision cancer therapy. Acta Pharmacol. Sin. 2015, 36, 1163–1169. [Google Scholar] [CrossRef] [PubMed]
  100. Houghton, P.J.; Morton, C.L.; Kolb, E.A.; Gorlick, R.; Lock, R.; Carol, H.; Reynolds, C.P.; Maris, J.M.; Keir, S.T.; Billups, C.A.; et al. Initial testing (stage 1) of the mTOR inhibitor rapamycin by the pediatric preclinical testing program. Pediatr. Blood Cancer 2008, 50, 799–805. [Google Scholar] [CrossRef] [PubMed]
  101. Houghton, P.J.; Morton, C.L.; Gorlick, R.; Lock, R.B.; Carol, H.; Reynolds, C.P.; Kang, M.H.; Maris, J.M.; Keir, S.T.; Kolb, E.A.; et al. Stage 2 combination testing of rapamycin with cytotoxic agents by the Pediatric Preclinical Testing Program. Mol. Cancer Ther. 2010, 9, 101–112. [Google Scholar] [CrossRef] [PubMed]
  102. Fleuren, E.D.; Versleijen-Jonkers, Y.M.; Roeffen, M.H.; Franssen, G.M.; Houghton, P.J.; Oyen, W.J.; Boerman, O.C.; van der Graaf, W.T. Abstract 2761: Temsirolimus is effective as a single agent and in combination with cisplatin or bevacizumab in preclinical osteosarcoma models. Cancer Res. 2014, 73, 2761. [Google Scholar] [CrossRef]
  103. Chawla, S.P.; Staddon, A.P.; Baker, L.H.; Schuetze, S.M.; Tolcher, A.W.; D’Amato, G.Z.; Blay, J.-Y.; Mita, M.M.; Sankhala, K.K.; Berk, L.; et al. Phase II study of the mammalian target of rapamycin inhibitor ridaforolimus in patients with advanced bone and soft tissue sarcomas. J. Clin. Oncol. 2012, 30, 78–84. [Google Scholar] [CrossRef] [PubMed]
  104. Demetri, G.D.; Chawla, S.P.; Ray-Coquard, I.; Le Cesne, A.; Staddon, A.P.; Milhem, M.M.; Penel, N.; Riedel, R.F.; Bui-Nguyen, B.; Cranmer, L.D.; et al. Results of an international randomized phase III trial of the mammalian target of rapamycin inhibitor ridaforolimus versus placebo to control metastatic sarcomas in patients after benefit from prior chemotherapy. J. Clin. Oncol. 2013, 31, 2485–2492. [Google Scholar] [CrossRef] [PubMed]
  105. Houghton, P.J.; Gorlick, R.; Kolb, E.A.; Lock, R.; Carol, H.; Morton, C.L.; Keir, S.T.; Reynolds, C.P.; Kang, M.H.; Phelps, D.; et al. Initial testing (stage 1) of the mTOR kinase inhibitor AZD8055 by the pediatric preclinical testing program. Pediatr. Blood Cancer 2012, 58, 191–199. [Google Scholar] [CrossRef] [PubMed]
  106. Gobin, B.; Battaglia, S.; Lanel, R.; Chesneau, J.; Amiaud, J.; Rédini, F.; Ory, B.; Heymann, D. NVP-BEZ235, a dual PI3K/mTOR inhibitor, inhibits osteosarcoma cell proliferation and tumor development in vivo with an improved survival rate. Cancer Lett. 2014, 344, 291–298. [Google Scholar] [CrossRef] [PubMed]
  107. Cheung, C.H.A.; Lin, W.-H.; Hsu, J.T.-A.; Hour, T.-C.; Yeh, T.-K.; Ko, S.; Lien, T.-W.; Coumar, M.S.; Liu, J.-F.; Lai, W.-Y.; et al. BPR1K653, a novel Aurora kinase inhibitor, exhibits potent anti-proliferative activity in MDR1 (P-gp170)-mediated multidrug-resistant cancer cells. PLoS ONE 2011, 6, e23485. [Google Scholar] [CrossRef] [PubMed]
  108. Bayani, J.; Zielenska, M.; Pandita, A.; Al-Romaih, K.; Karaskova, J.; Harrison, K.; Bridge, J.A.; Sorensen, P.; Thorner, P.; Squire, J.A. Spectral karyotyping identifies recurrent complex rearrangements of chromosomes 8, 17, and 20 in osteosarcomas. Genes Chromosomes Cancer 2003, 36, 7–16. [Google Scholar] [CrossRef] [PubMed]
  109. Jiang, Z.; Jiang, J.; Yang, H.; Ge, Z.; Wang, Q.; Zhang, L.; Wu, C.; Wang, J. Silencing of Aurora kinase A by RNA interference inhibits tumor growth in human osteosarcoma cells by inducing apoptosis and G2/M cell cycle arrest. Oncol. Rep. 2014, 31, 1249–1254. [Google Scholar] [PubMed]
  110. Maris, J.M.; Morton, C.L.; Gorlick, R.; Kolb, E.A.; Lock, R.; Carol, H.; Keir, S.T.; Reynolds, C.P.; Kang, M.H.; Wu, J.; et al. Initial testing of the aurora kinase A inhibitor MLN8237 by the Pediatric Preclinical Testing Program (PPTP). Pediatr. Blood Cancer 2010, 55, 26–34. [Google Scholar] [CrossRef] [PubMed]
  111. Niu, N.-K.; Wang, Z.-L.; Pan, S.-T.; Ding, H.-Q.; Au, G.H.T.; He, Z.; Zhou, Z.-W.; Xiao, G.; Yang, Y.-X.; Zhang, X.; et al. Pro-apoptotic and pro-autophagic effects of the Aurora kinase A inhibitor alisertib (MLN8237) on human osteosarcoma U-2 OS and MG-63 cells through the activation of mitochondria-mediated pathway and inhibition of p38 MAPK/PI3K/Akt/mTOR signaling pathway. Drug Des. Dev. Ther. 2015, 9, 1555–1584. [Google Scholar]
  112. Theoleyre, S.; Wittrant, Y.; Tat, S.K.; Fortun, Y.; Redini, F.; Heymann, D. The molecular triad OPG/RANK/RANKL: Involvement in the orchestration of pathophysiological bone remodeling. Cytokine Growth Factor Rev. 2004, 15, 457–475. [Google Scholar] [CrossRef] [PubMed]
  113. Boyce, B.F.; Xing, L. Functions of RANKL/RANK/OPG in bone modeling and remodeling. Arch. Biochem. Biophys. 2008, 473, 139–146. [Google Scholar] [CrossRef] [PubMed]
  114. Mori, K.; Le Goff, B.; Berreur, M.; Riet, A.; Moreau, A.; Blanchard, F.; Chevalier, C.; Guisle-Marsollier, I.; Léger, J.; Guicheux, J.; et al. Human osteosarcoma cells express functional receptor activator of nuclear factor-kappa B. J. Pathol. 2007, 211, 555–562. [Google Scholar] [CrossRef] [PubMed]
  115. Beristain, A.G.; Narala, S.R.; Di Grappa, M.A.; Khokha, R. Homotypic RANK signaling differentially regulates proliferation, motility and cell survival in osteosarcoma and mammary epithelial cells. J. Cell Sci. 2012, 125, 943–955. [Google Scholar] [CrossRef] [PubMed]
  116. Lee, J.A.; Jung, J.S.; Kim, D.H.; Lim, J.S.; Kim, M.S.; Kong, C.B.; Song, W.S.; Cho, W.H.; Jeon, D.G.; Lee, S.Y.; et al. RANKL expression is related to treatment outcome of patients with localized, high-grade osteosarcoma. Pediatr. Blood Cancer 2011, 56, 738–743. [Google Scholar] [CrossRef] [PubMed]
  117. Lamoureux, F.; Richard, P.; Wittrant, Y.; Battaglia, S.; Pilet, P.; Trichet, V.; Blanchard, F.; Gouin, F.; Pitard, B.; Heymann, D.; et al. Therapeutic relevance of osteoprotegerin gene therapy in osteosarcoma: Blockade of the vicious cycle between tumor cell proliferation and bone resorption. Cancer Res. 2007, 67, 7308–7318. [Google Scholar] [CrossRef] [PubMed]
  118. Gobin, B.; Baud’Huin, M.; Isidor, B.; Heymann, D.; Heymann, M.-F. mAbs targeting RANKL in bone metastasis treatment. In Monoclonal Antibodies in Oncology; Future Medicine Ltd.: London, UK, 2013; pp. 42–53. [Google Scholar]
  119. Heymann, D. Anti-RANKL therapy for bone tumours: Basic, pre-clinical and clinical evidences. J. Bone Oncol. 2012, 1, 2–11. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  120. Stopeck, A.T.; Lipton, A.; Body, J.-J.; Steger, G.G.; Tonkin, K.; de Boer, R.H.; Lichinitser, M.; Fujiwara, Y.; Yardley, D.A.; Viniegra, M.; et al. Denosumab compared with zoledronic acid for the treatment of bone metastases in patients with advanced breast cancer: A randomized, double-blind study. J. Clin. Oncol. 2010, 28, 5132–5139. [Google Scholar] [CrossRef] [PubMed]
  121. Zitvogel, L.; Kroemer, G. Targeting PD-1/PD-L1 interactions for cancer immunotherapy. Oncoimmunology 2012, 1, 1223–1225. [Google Scholar] [CrossRef] [PubMed]
  122. Dong, H.; Strome, S.E.; Salomao, D.R.; Tamura, H.; Hirano, F.; Flies, D.B.; Roche, P.C.; Lu, J.; Zhu, G.; Tamada, K.; et al. Tumor-associated B7-H1 promotes T-cell apoptosis: A potential mechanism of immune evasion. Nat. Med. 2002, 8, 793–800. [Google Scholar] [CrossRef] [PubMed]
  123. Lussier, D.M.; O’Neill, L.; Nieves, L.M.; McAfee, M.S.; Holechek, S.A.; Collins, A.W.; Dickman, P.; Jacobsen, J.; Hingorani, P.; Blattman, J.N. Enhanced T-cell immunity to osteosarcoma through antibody blockade of PD-1/PD-L1 interactions. J. Immunother. 2015, 38, 96–106. [Google Scholar] [CrossRef] [PubMed]
  124. Huang, J.; Shi, Y.; Li, H.; Yang, M.; Liu, G. MicroRNA-144 acts as a tumor suppressor by targeting Rho-associated coiled-coil containing protein kinase 1 in osteosarcoma cells. Mol. Med. Rep. 2015, 12, 4554–4559. [Google Scholar] [CrossRef] [PubMed]
  125. Wang, W.; Zhou, X.; Wei, M. MicroRNA-144 suppresses osteosarcoma growth and metastasis by targeting ROCK1 and ROCK. Oncotarget 2015, 6, 10297–10308. [Google Scholar] [CrossRef] [PubMed]
  126. Wang, J.; Xu, G.; Shen, F.; Kang, Y. miR-132 targeting cyclin E1 suppresses cell proliferation in osteosarcoma cells. Tumour Biol. 2014, 35, 4859–4865. [Google Scholar] [CrossRef] [PubMed]
  127. Liu, Y.; Li, Y.; Liu, J.; Wu, Y.; Zhu, Q. MicroRNA-132 inhibits cell growth and metastasis in osteosarcoma cell lines possibly by targeting Sox. Int. J. Oncol. 2015, 47, 1672–1684. [Google Scholar] [PubMed]
  128. Salah, Z.; Arafeh, R.; Maximov, V. miR-27a and miR-27a* contribute to metastatic properties of osteosarcoma cells. Oncotarget 2015, 6, 4920. [Google Scholar] [CrossRef] [PubMed]
  129. Tang, J.; Zhao, H.; Cai, H.; Wu, H. Diagnostic and prognostic potentials of microRNA-27a in osteosarcoma. Biomed. Pharmacother. 2015, 71, 222–226. [Google Scholar] [CrossRef] [PubMed]
  130. Zhou, Y.; Huang, Z.; Wu, S.; Zang, X. miR-33a is up-regulated in chemoresistant osteosarcoma and promotes osteosarcoma cell resistance to cisplatin by down-regulating TWIST. J. Exp. Clin. Cancer Res. 2014, 33, 12. [Google Scholar] [CrossRef] [PubMed]
  131. Gorlick, R.; Sowers, R.; Wenzel, B.D.; Richardson, C.; Meyers, P.A.; Healey, J.H.; Levy, A.S. Impairment of methotrexate transport is common in osteosarcoma tumor samples. Sarcoma 2011, 2011, 834170. [Google Scholar]
  132. Trippett, T.; Meyers, P.; Gorlick, R.; Steinherz, P. High dose trimetrexate with leucovorin protection in recurrent childhood malignancies: A Phase II trial. Proc. Am. Soc. Clin. Oncol. 1999, 18, 231a. [Google Scholar]
  133. Rollins, K.D.; Lindley, C. Pemetrexed: A multitargeted antifolate. Clin. Ther. 2005, 27, 1343–1382. [Google Scholar] [CrossRef] [PubMed]
  134. Adjei, A.A. Pemetrexed (Alimta): A novel multitargeted antifolate agent. Expert Rev. Anticancer Ther. 2003, 3, 145–156. [Google Scholar] [CrossRef] [PubMed]
  135. Bodmer, N.; Walters, D.; Fuchs, B. Pemetrexed, a multitargeted antifolate drug, demonstrates lower efficacy in comparison to methotrexate against osteosarcoma cell lines. Pediatr. Blood Cancer 2008, 50, 905–908. [Google Scholar] [CrossRef] [PubMed]
  136. Chou, A.J.; Bell, M.D.; Mackinson, C.; Gupta, R.; Meyers, P.A.; Gorlick, R. Phase Ib/IIa study of sustained release lipid inhalation targeting cisplatin by inhalation in the treatment of patients with relapsed/progressive osteosarcoma metastatic to the lung. ASCO Meet. Abstr. 2007, 25, 9525. [Google Scholar]
  137. Chou, A.J.; Gupta, R.; Bell, M.D.; Riewe, K.O.; Meyers, P.A.; Gorlick, R. Inhaled lipid cisplatin (ILC) in the treatment of patients with relapsed/progressive osteosarcoma metastatic to the lung. Pediatr. Blood Cancer 2013, 60, 580–586. [Google Scholar] [CrossRef] [PubMed]
  138. Wu, D.; Wan, M. Methylene diphosphonate-conjugated adriamycin liposomes: Preparation, characteristics, and targeted therapy for osteosarcomas in vitro and in vivo. Biomed. Microdevices 2012, 14, 497–510. [Google Scholar] [CrossRef] [PubMed]
  139. Alberts, D.S.; Muggia, F.M.; Carmichael, J.; Winer, E.P.; Jahanzeb, M.; Venook, A.P.; Skubitz, K.M.; Rivera, E.; Sparano, J.A.; Dibella, N.J.; et al. Efficacy and safety of liposomal anthracyclines in Phase I/II clinical trials. Semin. Oncol. 2004, 31, 53–90. [Google Scholar] [CrossRef] [PubMed]
  140. Kratz, F. DOXO-EMCH (INNO-206): The first albumin-binding prodrug of doxorubicin to enter clinical trials. Expert Opin. Investig. Drugs 2007, 16, 855–866. [Google Scholar] [CrossRef] [PubMed]
  141. Unger, C.; Häring, B.; Medinger, M.; Drevs, J.; Steinbild, S.; Kratz, F.; Mross, K. Phase I and pharmacokinetic study of the (6-maleimidocaproyl)hydrazone derivative of doxorubicin. Clin. Cancer Res. 2007, 13, 4858–4866. [Google Scholar] [CrossRef] [PubMed]
  142. Mita, M.M.; Natale, R.B.; Wolin, E.M.; Laabs, B.; Dinh, H.; Wieland, S.; Levitt, D.J.; Mita, A.C. Pharmacokinetic study of aldoxorubicin in patients with solid tumors. Investig. New Drugs 2015, 33, 341–348. [Google Scholar] [CrossRef] [PubMed]
  143. Zhang, L.; Marrano, P.; Kumar, S.; Leadley, M.; Elias, E.; Thorner, P.; Baruchel, S. Nab-Paclitaxel is an active drug in preclinical model of pediatric solid tumors. Clin. Cancer Res. 2013, 19, 5972–5983. [Google Scholar] [CrossRef] [PubMed]
  144. Yang, Y.; Niu, X.; Zhang, Q.; Hao, L.; Ding, Y.; Xu, H. The efficacy of abraxane on osteosarcoma xenografts in nude mice and expression of secreted protein, acidic and rich in cysteine. Am. J. Med. Sci. 2012, 344, 199–205. [Google Scholar] [CrossRef] [PubMed]
  145. Wagner, L.M.; Yin, H.; Eaves, D.; Currier, M.; Cripe, T.P. Preclinical evaluation of nanoparticle albumin-bound paclitaxel for treatment of pediatric bone sarcoma. Pediatr Blood Cancer 2014, 61, 2096–2098. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Overview of the targeted therapies in Osteosarcoma. This figure schematically shows molecular targets and associated drugs identified for therapeutic intervention in osteosarcoma. Therapeutic targets include specific cell surface receptor tyrosine kinases (RTKs): HER2, insulin-like growth factor 1 receptor (IGF1R), platelet-derived growth factor receptor (PDGFR), vascular endothelial growth factor receptor (VEGFR); and intracellular signaling targets: steroid receptor co-activator (SRC), RAF, mTOR and aurora kinases. Other potential targets include receptor activator of nuclear factor-κB ligand (RANKL), AKT, mTOR, programmed cell death 1 (PD-1)/programmed cell death ligand 1 (PD-L1), and folate.
Figure 1. Overview of the targeted therapies in Osteosarcoma. This figure schematically shows molecular targets and associated drugs identified for therapeutic intervention in osteosarcoma. Therapeutic targets include specific cell surface receptor tyrosine kinases (RTKs): HER2, insulin-like growth factor 1 receptor (IGF1R), platelet-derived growth factor receptor (PDGFR), vascular endothelial growth factor receptor (VEGFR); and intracellular signaling targets: steroid receptor co-activator (SRC), RAF, mTOR and aurora kinases. Other potential targets include receptor activator of nuclear factor-κB ligand (RANKL), AKT, mTOR, programmed cell death 1 (PD-1)/programmed cell death ligand 1 (PD-L1), and folate.
Ijms 17 00506 g001
Table 1. Clinical trials of tyrosine kinase receptor inhibitors in osteosarcoma.
Table 1. Clinical trials of tyrosine kinase receptor inhibitors in osteosarcoma.
TargetClassTherapyMechanismClinical TrialOutcomeReference
IGF-1RAnti- IGF-R antibodiesCixutumumabActivation of (PI3K/Akt ) and MAPKPhase I/IIPD[37]
Phase IINo OR[38]
Phase IIMPFS at 6 weeks[39]
Phase IIMPFS at 6 weeks[40]
Phase IIMPFS at 21.4 weeks[41]
Human mAb SCH 717454 (robatumumab)Inhibits IGF-R binding and signalingPhase I/IB (Terminated)NCT00960063 *
Phase II (Terminated)NCT00617890 *
IGF ligand-neutralizing antibodiesBI 836845 mAbNeutralizes IGF ligandPhase INCT01403974 *
Phase INCT02145741 *
Phase INCT01317420 *
Small-molecule TKI’sBMS-754807ATP-competitive inhibitor of IGFPhase 1INCT00898716 *
Phase IIINCT00134030 *
PDGFRSmall-molecule TKI’sImatinib mesylatePDFGR, c-KITPhase IIno OR[42]
Phase IINCT00030667 *
Phase IINCT00031915 *
Multi-targeted RTK inhibitorsSunitinibPDGFR, FLT3, RET, KIT and VEGFR inhibitorPhase ISD in 1(2)[43]
VEGF/VEGFRAnti-VEGF antibodiesBevacizumabVEGF inhibitorPhase Ino CPR[44]
Phase II NCT00458731 *
Small-molecule TKI’sSorafenibPDGFR, FLT3, RET, c-KIT, VEGFR inhibitorPhase II45% PFS at 6 months[45]
Phase INCT00665990 *
Phase INCT01518413 *
HER-2Anti-HER-2 antibodyTrastuzumabHER-2 inhibitorPhase IINSD[46]
Phase IINCT00005033 *
Abbreviations: PR, partial response; PD, progressive disease; SD, stable disease; MPFS, median progression-free survival; PFS, progression free survival; CPR, complete or partial response; OR, Objective response; NSD, no significant difference; * clinical trial number [47].
Table 2. Clinical trials of Intracellular signaling inhibitors in Osteosarcoma.
Table 2. Clinical trials of Intracellular signaling inhibitors in Osteosarcoma.
TargetClassTherapyMechanismClinical TrialOutcomeReference
Srcsmall-molecule TKI’s (Multi targeted)DasatinibPDGF/PDGFR, SRC, BCR-ABL inhibitorPhase ISD in 1(1)[87]
Phase IISD in 5(45)[88]
Phase IINCT00464620 *
Phase I/IINCT00788125 *
Dual-inhibitorSaracatinibSrc and Abl specific inhibitorPhase IINCT00752206 *
mTOR1st generation of mTOR inhibitorsSirolimusInhibit mTORC1 by binding toFKBP-12Phase IINo CPR[91]
EverolimusPhase ISD in 1(2), no OR[92]
Phase ISD in 3(3)[93]
Phase IINCT01216826 *
TemsirolimusPhase IIMPFS at 21.4 weeks[41]
Phase IIMPFS at 6 weeks[40]
Phase IINCT01614795 *
2nd generation of dual mTOR inhibitorsAZD8055mTOR1, mTOR2 inhibitorPhase INCT00731263 *
Phase I withdrawnNCT01194193 *
Aurora kinase A2nd generation aurora kinase inhibitorsMLN8237 (alisertib)AURK-Ainhibitor via ATP bindingPhase INCT02214147 *
Phase INCT01898078 *
Phase IINCT01154816 *
Phase INCT02444884 *
FolateMultitargeted antifolatePemetrexedfolate-dependent enzymes and DNA synthesis enzymes inhibitorPhase IICPR in 1[94]
SD in 5
PD in 22
MPFS at 1.4 months
Phase IIno CPR[95]
Phase IINCT00003776 *
Phase IINCT00002738 *
Phase INCT00119301 *
Abbreviations: PR, partial response; PD, progressive disease; SD, stable disease; MPFS, median progression-free survival; PFS, progression free survival; CPR, complete or partial response; OR, Objective response; NSD, no significant difference;* clinical trial number [47].

Share and Cite

MDPI and ACS Style

Shaikh, A.B.; Li, F.; Li, M.; He, B.; He, X.; Chen, G.; Guo, B.; Li, D.; Jiang, F.; Dang, L.; et al. Present Advances and Future Perspectives of Molecular Targeted Therapy for Osteosarcoma. Int. J. Mol. Sci. 2016, 17, 506. https://0-doi-org.brum.beds.ac.uk/10.3390/ijms17040506

AMA Style

Shaikh AB, Li F, Li M, He B, He X, Chen G, Guo B, Li D, Jiang F, Dang L, et al. Present Advances and Future Perspectives of Molecular Targeted Therapy for Osteosarcoma. International Journal of Molecular Sciences. 2016; 17(4):506. https://0-doi-org.brum.beds.ac.uk/10.3390/ijms17040506

Chicago/Turabian Style

Shaikh, Atik Badshah, Fangfei Li, Min Li, Bing He, Xiaojuan He, Guofen Chen, Baosheng Guo, Defang Li, Feng Jiang, Lei Dang, and et al. 2016. "Present Advances and Future Perspectives of Molecular Targeted Therapy for Osteosarcoma" International Journal of Molecular Sciences 17, no. 4: 506. https://0-doi-org.brum.beds.ac.uk/10.3390/ijms17040506

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop