Abstract
Osteosarcoma is the most common primary malignancy of bone, typically presenting in the first or second decade of life. Unfortunately, clinical outcomes for osteosarcoma patients have not substantially improved in over 30 years. This stagnation in therapeutic advances is perhaps explained by the genetic, epigenetic, and biological complexities of this rare tumor. In this review we provide a general background on the biology of osteosarcoma and the clinical status quo. We go on to enumerate the genetic and epigenetic defects identified in osteosarcoma. Finally, we discuss ongoing large-scale studies in the field and potential new therapies that are currently under investigation.
Keywords: epigenetics, genetics, metastasis, osteosarcoma, therapy
I. INTRODUCTION
Osteosarcoma is a rare mesenchymal tumor that is histologically characterized by the presence of malignant mesenchymal cells and the production of a bone stroma. Within this histological diagnosis, distinct histological subtypes have been defined, including osteoblastic, chondroblastic, fibroblastic, and telangiectatic osteosarcomas.1 Nonetheless, the biological behavior of and treatments for these distinct subtypes are similar. In the United States there are approximately 900 new cases of osteosarcoma reported each year.2 Despite its rarity, however, osteosarcoma represents the most common primary malignancy of bone. The biology of this tumor is most commonly characterized by an appendicular primary tumor with a high rate of metastasis to the lungs.3 This review summarizes the biological behavior of this tumor and conventional approaches to therapy. Given the stagnation in clinical outcomes seen with such conventional therapies over the past 3 decades, a focus of this review is the ongoing efforts to define the genetic and epigenetic changes that are associated with tumor formation and those associated with progression and metastasis. Using these data, we propose a biological rationale to consider novel and targeted therapies for osteosarcoma.
II. PATHOGENESIS, BIOLOGY, AND CONVENTIONAL THERAPY
Genetic aberrations or changes in gene expression in osteosarcoma tissues have not identified common or recurrent genetic lesions or pathway alterations that explain the development of this tumor type.4,5 Rather, osteosarcoma is best characterized by its disorganized genome. Indeed, the most consistent genetic finding in osteosarcoma, beyond dysregulation of p53 and Rb (retinoblastoma), is significant aneuploidy and some evidence of massive disruption in the control of chromosomal structure (i.e., chromothripsis).5,6 This has suggested the possibility of an early defect in DNA repair/surveillance as a mechanism for osteosarcomagenesis and the resultant bizarre aneuploidy.
A. Cell of Origin
Beyond the difficulty in identifying recurrent and driving genetic alterations associated with osteosarcoma development, the cell of origin for this tumor type remains unclear.7 Historically, cells committed to the bone lineage (e.g., the osteoblast and preosteoblast) were reasonably believed to represent a cellular target for transformation.8–10 More recent data have been used to suggest the hypothesis that the osteosarcoma-transforming event occurs in multipotent mesenchymal stem cells, and that this event then yields a karyotypic complexity that is tolerated by these primitive mesenchymal cells, subsequently driving them down bone-differentiation lineages.11,12 Irrespective of the actual cell of origin, it is commonly agreed that the gene expression and the cellular phenotype of osteosarcoma are related to bone.13–15
Accordingly, the characteristic presentation of osteosarcoma indicates a tight association with normal phases of skeletal development. Osteosarcoma most often develops in the appendicular skeleton in pediatric patients, during the second decade of life. The timing of tumor diagnosis in patients coincides with the developmental growth spurt that occurs in this patient population.16 This epidemiological association as well as both cellular and molecular data have suggested a role for skeletal growth regulating the growth hormone–insulin-like growth factor axis in the development and progression of osteosarcoma.17
B. Metastasis and Progression
The metastatic cascade represents a series of processes that occur as a cell leaves the primary tumor and invades the surrounding tumor microenvironment, leading to intravasation into existing or new vascular structures, survival in the circulation, and eventual arrest and extravasation at distant secondary sites, followed by the development or recruitment of a blood supply and growth at the secondary site.18 Successful metastatic cells must detect, modulate, and manage discrete cellular stresses to metastasize successfully.19–22 A critical and potentially defining determinant of metastatic success that has been observed in many cancers is their ability to overcome the substantial stress experienced as cells engage the microenvironment of the secondary site.19,20,23 Cells must rapidly adapt to this new and hostile environment to survive. This is important as tumor cells first enter a secondary metastatic environment but also as cells reengage the microenvironment later during metastatic progression. We and others believe the stresses faced by metastatic cells are distinct from those experienced during initial tumor formation. Our understanding of the timing of events leading to metastatic progression, and the events themselves, are far from complete. For example, it is unclear whether microscopic cells that leave the primary tumor early in the development of osteosarcoma move directly to the lung and reside there for long periods of time (dormancy), or whether these cells transit from the primary tumor to “protected environments,” where they become dormant and then move through all the steps of metastasis en route to the lung.24 Similarly, it is reasonable that the cells capable of completing the complex set of events required for metastasis continue to metastasize to more distant sites (within the lung in osteosarcoma).25 Indeed, if metastatic cells emerge from “protected sites,” and if metastatic lesions continue to metastasize, the opportunity to improve patient outcomes by targeting the process of metastasis would have merit not only in patients who present with localized and presumed microscopic metastasis but also in patients presenting with gross metastasis. Shortcomings in our understanding of the biology of metastasis preclude any a priori judgments of the value of agents that target any aspect of the process of metastasis as potentially valuable in osteosarcoma. Rigorous preclinical studies and innovative clinical trial designs, however, are necessary to allow the assessment of agents that target metastatic progression in patients. This is particularly true in the case of osteosarcoma because outcomes for these patients have not improved for more than 3 decades.3 Many of the aberrant genetic and epigentic mechanisms leading to the development of osteosarcoma tumors are reasonably expected to contribute to metastatic progression. These mechanisms are summarized in figure 1, which is meant to serve as a visual reference for the descriptions in the text to follow.
FIG. 1.
Identified molecular alterations leading to metastatic traits of osteosarcoma cells. This image depicts the genetic, epigenetic, and other molecular biological processes whose defects have been associated with metastatic traits of osteosarcoma cells. lncRNA, long noncoding RNA; mRNA, messenger RNA; miRNA, microRNA.
C. Conventional Treatment and Outcomes
Approximately 80% of patients present with grossly localized disease.26 Accordingly, definitive resection of the primary tumor is necessary, and this is readily achieved in most patients.3 Neoadjuvant chemotherapy has been incorporated into many treatment protocols. The use of neoadjuvant chemotherapy has many goals, but, most clearly, it seems to simplify surgical resection.27 Assessment of tumor necrosis following neoadjuvant therapy can be prognostically valuable,28 and it can also be used to prioritize chemotherapeutics that may be best used in the adjuvant setting. Nonetheless, it is unclear whether the use of neoadjuvant therapy improves long-term outcomes for patients beyond improving the management of the primary tumor. Despite the localized presentation for most patients, the successful resection of most primary tumors, and both neoadjuvant and adjuvant chemotherapy, many patients (35%) develop metastases.28,29 This is certainly the result of microscopic metastases that were not detectable at the time of presentation. Among patients who present with disseminated disease, over 80% progress within 5 years.3 Long-term outcomes for patients who present with localized or disseminated disease have largely remained unchanged over the past 30 years.3 Greater intensification of cytotoxic therapy alone is not likely to improve these long-term outcomes. Accordingly, progress is needed to define recurrent genetic or epigenetic alterations linked to osteosarcomagenesis and, perhaps more important, linked to osteosarcoma progression and metastasis.
III. GENETICS
A. Genetic Susceptibility
1. Genome-wide Association Studies
The past decade has seen an explosion of studies seeking to understand how genetic variability contributes to disease. These population-based studies, known as genome-wide association studies, seek to detect genetic variants—most commonly single nucleotide polymorphisms (SNPs)—that are associated with complex traits in populations (e.g., susceptibility to cancer).30 Recent studies of humans and dogs with osteosarcoma revealed multiple SNPs associated with risk for the development of osteosarcoma.31,32
Numerous studies associating common genetic variants with osteosarcoma risk have been published in the past 15 years.33–39 While in these studies risk SNPs have been linked to biological pathways with known relevance to osteosarcomagenesis, their statistical power has been limited by a small sample size because of the rarity of this cancer type. A recently published study sought to overcome such limitations through an international collaborative effort comparing genotypes of 941 human osteosarcoma cases with those of 3291 controls. Data from this study demonstrated a significant association of 3 SNPs with osteosarcoma risk. The first (rs1906953; P = 8.1 × 10−9) is located within intron 7 of the glutamate receptor metabotropic 4 (GRM4) gene at 6p21.3.31 GRM4 plays a role in cyclic AMP signaling, which has been linked to osteosarcoma in a number of studies,40,41 indicating its plausible ability to confer osteosarcoma risk. The locus maps to a DNase I hypersensitivity region in the Encyclopedia of DNA Elements data set, suggesting that it may contain active regulatory elements. The second and third SNPs (rs7591996 and rs10208273; P = 1.0 × 10−8 and 2.9 × 10−7 , respectively) are located in the gene desert at 2p25.2. While neither of these lead SNPs were associated with regulatory elements or transcription factor binding sites in the Encyclopedia of DNA Elements data set, several surrogate SNPs occurred within transcription factor binding sites or altered known regulatory motifs.31
Pet dogs develop osteosarcoma that shares many features with the human disease, including tumor histology, gene expression, response to chemotherapy, and risk for pulmonary metastasis.42 Accordingly, the dog with osteosarcoma provides a valuable model for the study of cancer-associated genes, drug development, and prognostic markers. A recently published genome-wide association study sought to identify risk loci for osteosarcoma in 3 dog breeds at high risk for osteosarcoma. The study included 286 greyhounds, 135 Rottweilers, and 141 Irish wolfhounds, with relatively equal numbers of cases and controls for each breed. The study identified 33 inherited risk loci accounting for 55–85% of phenotype variance within a breed. The SNP with the strongest association with osteosarcoma development in greyhounds was located 150 kilobases upstream of the CDKN2A/B genes, which are known to play a key role in osteosarcoma development and progression (see section III, B, 3, a). The top SNP in Rottweilers and Irish wolfhounds alters an evolutionarily constrained enhancer element that was active in human osteosarcoma cells. Loci among all breeds were enriched for genes with key functions in bone differentiation and development.32
2. Genetic Syndromes Associated with Osteosarcoma
Increased risk of osteosarcoma is associated with a number of well-defined genetic syndromes: hereditary retinoblastoma (germline mutation of the Rb gene), Li-Fraumeni syndrome (germline mutation of the p53 gene), Bloom syndrome (germline mutation of the RECQL2 gene), Werner syndrome (germline mutation of the RECQL3 gene), and Rothmund-Thomson syndrome (germline mutation of the RECQL4 gene).43 Many of these genes and pathways are commonly altered by somatic mutations in osteosarcoma tumors, although the mechanism of mutation is often distinct from these germline mutations.
B. Somatic Genetic Alterations in Osteosarcoma
As stated above, the rarity of osteosarcoma within the population makes comprehensive genetic and genomic analyses difficult. Numerous studies, often investigating only a subset of common genetic mutations in relatively small patient or cell line cohorts, have been reported. Not surprisingly, such studies make it difficult to confidently assess the frequency and functional consequences of the investigated genetic abnormalities in osteosarcoma. Accordingly, mutation frequencies often are reported as a range identified from multiple publications. Of note, the Therapeutically Applicable Research To Generate Effective Treatments Osteosarcoma Project is currently in progress (http://ocg.cancer.gov/programs/target/projects/osteosarcoma). This is a large-scale, multi-institutional collaborative effort to comprehensively identify genetic and epigenetic aberrations in osteosarcoma using a combination of genomic approaches. The results of this study are expected to be published within the next year and should shed new light on the genetic and epigenetic drivers of osteosarcoma. This effort will move osteosarcoma into the postgenomic era, allowing for subsequent interrogations of genetic contributions to osteosarcoma to be completed in silico, a luxury available in the study of other more common cancers but not yet possible for osteosarcoma.
1. Genetic Heterogeneity
As is described in detail below, the mutational landscape of osteosarcoma is highly complex and varies significantly between tumors.5 This high degree of intertumor heterogeneity confounds our understanding of the molecular pathogenesis of osteosarcoma and may explain some of the difficulty in identifying therapeutic agents that are likely to improve outcomes for the spectrum of patients with osteosarcoma.
2. Chromosomal Abnormalities
A hallmark of osteosarcoma is chromosomal instability (CIN),44,45 a form of genome-wide alteration characterized by a high degree of losses and gains of full chromosomes or chromosomal segments.46,47 CIN has been shown to result from a loss of function in cell cycle checkpoint and DNA damage response pathways.48,49 As described below, these pathways can be dysregulated in osteosarcoma via both genetic and epigenetic mechanisms. Aberrant maintenance of telomeres through a mechanism known as alternative lengthening of telomeres also has been shown to result in CIN in osteosarcoma.50,51
Unlike many other sarcomas, osteosarcoma lacks a canonical translocation or genetic mutation.4,5,52–57 Rather, osteosarcoma is a cancer typified by widespread and heterogeneous abnormalities in chromosomal number and substructure. Osteosarcoma ploidy can range from haploidy to hexaploidy.52,58 While myriad chromosomal losses/gains have been identified, chromosome 1 is most often gained and chromosomes 9, 10, 13, and 17 are most often lost.52,58 The most common copy number alterations are deletions of portions of chromosomes 3, 6, 9, 10, 13, 17, and 18 and amplifications of portions of chromosomes 1, 6, 8, and 17. These regions encode a number of tumor suppressors and oncogenes, respectively.5
3. Tumor Suppressors
a. Rb Pathway
Rb is a critical regulator of the G1-to-S cell cycle transition. In the absence of mitogenic stimuli, Rb remains dephosphorylated and binds to E2F family transcription factors, preventing their activation of cell cycle progression. During normal mitosis, this is reversed via Rb phosphorylation by CDK4. Loss-of-function Rb mutations remove this cell cycle checkpoint.59 The CDKN2A locus (also known as INK4A) encodes 2 functionally and structurally distinct genes via alternative splicing. The first, p16INK4a , is a negative regulator of CDK4. The second, p14ARF , is a key regulator of p53 (see below). Loss of p16INK4a function alleviates negative regulation of CDK4, resulting in Rb inactivation.60 Thus, mutations in the CDKN2A gene can phenocopy loss-of-function Rb mutations.
Loss-of-function Rb mutations occur in up to 70% of osteosarcoma cases;61-64 the most common is loss of heterozygosity.62,65,66 Other types of Rb mutations include structural rearrangements and point mutations.61–64,67–69 In one study 70% of patients possessed deletions or rearrangements in the CDKN2A gene with the potential to reduce the expression or function of p16INK4a.57,70–73
b. p53 Pathway
p53 Is a transcription factor that regulates critical genes in DNA damage response, cell cycle progression, and apoptosis pathways.74 p53 acts as a tumor suppressor in essentially all tumor types, and its function can be affected by mutations to the gene itself or by mutations to up- or downstream mediators of its activity.75 p14ARF normally acts to sequester the E3 ubiquitin ligase MDM2 in the nucleolus, preventing it from promoting p53 degradation.74 p14ARF is expressed from the same CDKN2A locus that encodes p16INK4a (see above).60 Similar to p16INK4a in the Rb pathway, loss-of-function mutations in the p14ARF gene can phenocopy mutations to TP53.74
Loss-of-function TP53 mutations occur in as many as three-fourths of osteosarcoma cases.5 These mutations include allelic loss (75–80%), rearrangements (10–20%), and point mutations (20– 30%).76–83 A recent study demonstrated that 9.5% of young patients (<30 years of age) with sporadic osteosarcoma carried either a rare germline TP53 exonic variant or the canonical Li-Fraumeni mutation, but that these variants are absent from patients who develop osteosarcoma later in life.84 As stated above, as many as 70% of osteosarcoma tumors harbor mutations with the potential to affect p14ARF expression or function and, therefore, alter p53 function.57,70–73
c. Other Tumor Suppressors
Other tumor suppressors associated with deletions or loss of heterozygosity in osteosarcoma include APC, BUB3, FGFR2, LSAMP, RECQL4, and WWOX.65,85–97
4. Oncogenes
a. Rb Pathway
E2F3 and CDK4, both of which counteract Rb control of cell cycle progression, have been estimated to possess gain-of-function mutations in 60% and 10% of tumors, respectively.88,98,99
b. p53 Pathway
MDM2 is an E3 ubiquitin ligase that acts as a negative regulator of p53 (see above). The MDM2 gene is amplified in 3–25% of osteosarcoma tumors.61,99–102 COPS3 also promotes proteosomal degradation of p53 and is estimated to cause gain-of-function mutations in 20–80% of osteosarcomas.92,94,103–106
c. c-Myc
c-Myc is a key transcription factor that acts as a general amplifier of gene expression, enhancing the transcription of essentially all genes with active promoters in a given cell,107,108 and is a well-described oncogene with gained function in most tumor types.109 c-Myc is amplified in 7–67% of osteosarcoma tumors88,94,103,110–114 and overexpressed in at least 34% of tumors.115,116
d. Other Oncogenes
Other oncogenes associated with amplifications in osteosarcoma include CDC5L, MAPK7, MET, PIM1, PMP22, PRIM1, RUNX2, and VEGFA.85,88,94,98,104,114,117–125 Collectively, the finding that near ubiquitous alterations in the Rb and p53 pathway function in osteosarcoma through both gain- and loss-of-function mutations indicates that loss of cell cycle control and inappropriate DNA damage response are key drivers of osteosarcoma development. The role that these genetic alterations play in tumor progression and metastasis, however, remains less clear.
C. Mechanisms of Genetic Aberration in Osteosarcoma
As described above, osteosarcoma tumors display a compendium of genetic abnormalities with a high degree of intertumor heterogeneity. While certain genes are commonly altered across tumors, the most common genetic characteristic of osteosarcoma tumors is the remarkable breadth of genetic changes relative to normal tissue. Like the mutations themselves, the mechanisms by which these genetic alterations are acquired are likely to represent a broad spectrum both within and across tumors. Classically defined modes of genetic mutation are known to occur in osteosarcoma. For example, point mutations are likely the result of errors in DNA replication and subsequent proof reading, whereas aneuploidy is the result of errors in chromosomal segregation during cell division.126–128 In addition to these well-defined modes of genetic mutation, a novel mechanism of mutation acquisition known as chromothripsis has recently been identified. This term describes a phenomenon by which tens to hundreds of genomic rearrangements occur during cancer development in a one-off cellular crisis. This occurs through reciprocal exchange of genetic material within or between chromosomes. In contrast to the gradual mode of accumulated genetic aberrations in cancer cells acquired through singular mutational events and subsequent Darwinian clonal selection, this model posits “punctuated equilibrium” as the primary mode of tumor evolution. In their landmark paper, Stephens et al.6 demonstrated that chromothripsis occurs in at least 2–3% of all cancers and approximately 33% of osteosarcoma tumors.
IV. EPIGENETICS
While cancer has been classically defined as a disease resulting from genetic mutations, a vast and ever-growing body of literature, largely published within the past 15 years, has demonstrated that epigenetic mechanisms are near ubiquitous drivers of tumor development and progression. In this article we refer to epigenetics as the study of regulatory mechanisms affecting the expression of DNA templates without altering the sequence of the templates themselves. The most well-described epigenetic mechanisms involved in cancer biology include DNA methylation, histone modification, nucleosome remodeling, and RNA-mediated events.129 Importantly, many of these epigenetic processes can be affected by alterations in DNA sequence and vice versa, such that genetic and epigenetic regulation of cellular behavior are inextricable.129,130 The knowledge and technical approaches resulting from the development of this relatively new area of biologic inquiry have provided a new lens through which the biology of cancers, including osteosarcoma, may be investigated.
A. DNA Methylation
Methylation of the 5-carbon on cytosine in cytosine– phosphate–guanine (CpG) dinucleotides represents a key mechanism of epigenetic gene silencing. Roughly 70% of gene promoters contain dense CpG clusters known as CpG islands. Levels of methylation at promoter CpG islands are inversely correlated with gene expression. Gene silencing through promoter hypermethylation represents the most well-described mechanism of epigenetic dysregulation in cancer.131,132
1. Tumor Suppressors
a. Rb Pathway
While Rb has been shown to be epigenetically silenced by promoter hypermethylation in other cancers,133 there is no strong evidence that epigenetic silencing is a direct mechanism of lost Rb function in osteosarcoma.134 However, other members of the Rb pathway are subject to this mode of dysregulation. Specifically, promoter hypermethylation and consequent reduced gene expression have been demonstrated at the p16INK4a locus.135,136 Promoter hypermethylation also was detected at the CREG1 locus, which encodes a gene that acts to enhance p16INK4a-induced cellular senesce in osteosarcoma. Intriguingly, CREG1 expression was restored through treatment with the DNA demethylating agent 5-aza-2’-deoxycytidine (decitabine).137
b. p53 Pathway
Similar to Rb, the primary mechanism of direct p53 loss of function in osteosarcoma seems to be genetic.134 However, expression levels of multiple members of the p53 pathway are subject to promoter hypermethylation. p14ARF has been shown to be silenced through promoter hypermethylation in both osteosarcoma cell lines and 47% of tumor specimens.136,138 A cell line study suggested that p14ARF expression may be restored by treatment with decitabine.138 The same study demonstrated that expression of p21, a gene regulated by p53 that blocks G1-to-S cell cycle progression, was restored by treatment with decitabine, suggesting that this gene also is silenced by DNA hypermethylation.139 Another gene target of p53, GADD45, which plays a critical role in apoptosis induction in response to DNA damage, possesses a hypermethylated promoter in osteosarcoma cell lines and xenografts.139 GADD45 also has been shown to elicit DNA demethylation in response to numerous stimuli, indicating that epigenetic silencing of this gene is likely to have pleotropic epigenetic consequences in osteosarcoma.11,140–142 In addition, studies combining decitabine treatment with targeted GADD45 knockdown by small interfering RNA have demonstrated that this gene is responsible for apoptosis induction in response to decitabine treatment.107 HIC1 is a key modulator of p53-dependent responses to DNA damage.143 Various studies have shown that 17–77% of osteosarcomas possess hypermethylated HIC1 promoters and concomitant reductions in HIC1 expression.144,145 Inactivation of HIC1 alleviates repression of SIRT1, which subsequently inactivates p53.143
These studies indicate that the key Rb and p53 tumor suppressor pathways are subject to both genetic and epigenetic dysregulation in osteosarcoma. In addition, the activity of the p53 pathway affects the epigenomic DNA methylation status of osteosarcoma through the activity of GADD45, demonstrating a connection between DNA damage and altered epigenetic regulation in these cells.
c. Other Tumor Suppressors
Numerous tumor suppressors have been shown to be downregulated through somatic promoter hypermethylation in osteosarcoma cell lines and/or tumor samples. These include RASSF1A,146–148 TIMP3,148 MGMT,148 DAPK1,148 and WIF-1.149,150 A subset of these genes has been causally linked to tumor progression in osteosarcoma. Levels of promoter methylation of RASSF1A, a gene involved in cell cycle arrest, microtubule stabilization, and apoptosis,151 have been associated with clinical outcomes for patients with osteosarcoma.152 Like many other genes silenced by promoter hypermethylation, RASSF1A levels can be restored by decitabine treatment.147
2. Oncogenes
There exists to date little to no evidence to suggest that DNA methylation plays a role in oncogene overexpression in osteosarcoma.134 There is evidence, however, that oncogene activation may result in downstream epigenetic dysregulation. Activated Ras has been shown to downregulate Fas though DNA methyltransferase (DNMT)–mediated DNA methylation.153,154 Downregulation of Fas has been linked to the apoptotic escape of metastatic osteosarcoma cells within the lung microenvironment.155–159 Treatment of osteosarcoma cells with ibandronate was shown to inhibit activated Ras and downregulate DNMT expression, resulting in increased levels of Fas expression in vitro.160
B. Histone Modification
The essential units of chromatin structure are nucleosomes, consisting of DNA wrapped around core histone proteins. From each of these proteins protrude amino acid tails that are covalently modified at specific residues to dictate the functional modality of associated DNA elements, both coding and noncoding.161 The patterns of histone tail modifications and resultant genomic functions have been called the “histone code.” As one might predict, there are a large number of “writers,” “readers,” and “erasers” of the histone code, each with a specific role in regulating the biology of the eukaryotic genome.162 While much about the syntax of this code has been learned in the past several years,161,163,164 many of the specifics are still being worked out through experimentation.165 In parallel, cancer researchers are delineating how alterations in histone modifications contribute to tumor biology.166 Perhaps the most well-described influence of histone modifications on cell biology is their effect on the transcriptional output of cells traversing differentiation lineages.167,168
As stated above, the cell of origin and differentiation state of osteosarcoma have yet to be definitively characterized.7 It stands to reason that investigations of histone modifications in osteosarcoma cells may yield new insights into this question. While there are limited studies in this area to date, those published have been highly provocative. A number of studies (reviewed by Talluri and Dick169) demonstrated that Rb plays a role in regulating acetylation and methylation of histones through interactions with specific histone-deacetylating and -methylating complexes. Furthermore, Rb has been shown to broadly regulate chromatin structure though the generation of “senescence-associated heterochromatic foci.”170 Dysregulation of Rb in osteosarcoma is therefore expected to influence these processes. Interestingly, a recently published study demonstrated that osteosarcoma cells deficient in both Rb and p53 resembled mesenchymal stem cells and were able to differentiate down osteoblastic and adipogenic lineages, whereas osteosarcoma cells deficient in only p53 showed expression patterns concordant with the preosteoblastic state and were unable to differentiate into bone or fat cells.171 These findings suggest that the Rb mutational status of osteosarcoma is a critical determinate of differentiation potential and that this potential is likely influenced by Rb-mediated changes to chromatin.
Another recent study showed that lysine-specific demethylase 1 (LSD1), a histone demethylase, is overexpressed in osteosarcoma tumors and that treatment of osteosarcoma cells with the LSD1 inhibitor tranylcypromine reduced cell growth.172 LSD1 is known to regulate key processes during cellular development/differentiation,173 and its function has been shown to be partially dependent on p53.174
Finally, it was recently demonstrated that osteosarcoma cells show promoter hypermethylation and concomitant downregulation of a set of 384 genes that also are downregulated in human embryonic stem cells as a result of specific “bivalent” histone modifications linking these genes to pluripotency in human embryonic stem cells.175 The authors went on to demonstrate that osteosarcoma cells lack histone bivalency at these genes, suggesting that promoter hypermethylation may be a tumor-specific means of dedifferentiation, leading to a transcriptional output that more closely resembles that of pluripotent stem cells than the normal tissue from which the cancer developed.175
C. Noncoding RNAs
Studies over the past 50 years have demonstrated that a large proportion of genomic DNA is transcribed but not translated into protein (i.e., noncoding).176–180 Indeed, a recently published study demonstrated that as much as 75% of the human genome is capable of being transcribed,181 whereas only 1–2% of the genome contains protein-coding genes.182 Noncoding RNAs are broadly categorized based on their size. Transcripts whose length exceeds 200 base pairs are considered “long,” whereas transcripts below this length threshold are considered “small.” Long noncoding RNAs (lncR-NAs) are defined as any non–protein-coding transcripts over 200 base pairs in length. Many recently published studies have shown lncRNAs to be key regulators of a number of critical biological processes (summarized by Bergmann and Spector183). These studies demonstrated that one primary function of lncRNAs is to regulate gene expression through the formation of ribonucleoprotein complexes that critically influence the chromatin state of the cell.184 Small noncoding RNAs consist of a large number of subclasses, many of which are key regulators of gene expression. Perhaps the most well studied of these are microRNAs (miRNAs), which act to fine-tune gene expression by binding to messenger RNA transcripts to inhibit translation or induce degradation. The mechanisms of miRNA gene regulation are well described and extensively reviewed elsewhere.185 Both lncRNAs and miRNAs are critical in tuning gene expression during normal developmental processes. Like other forms of eukaryotic gene control, aberrations in gene regulation by non-coding RNAs can contribute to tumor development and progression.186,187 Osteosarcoma is no exception to this rule, and a number of published studies demonstrated the contribution of aberrant noncoding RNAs to osteosarcomagenesis and tumor progression.
1. Long Noncoding RNAs
A recent study91 showed that chr3q13.31, a region containing the lncRNAs LOC285194 and BC040587, was associated with copy number alteration in 80% of osteosarcomas. The most common mutation type was a deletion that resulted in reduced expression of the associated lncRNAs across osteosarcoma samples. It was further demonstrated that downregulation of LOC285194 promoted the proliferation of normal osteoblasts, supporting the role of this lncRNA as a tumor suppressor in osteosarcoma. LOC285194 knockdown also affected the expression of genes involved in the apoptotic and cell cycle progression pathways, as well as vascular endothelial growth factor receptor 1.91 A subsequent study188 showed that LOC285194 is a p53 target and that ectopic expression of LOC285194 inhibits tumor cell growth in vitro and in vivo. Furthermore, this study demonstrated that LOC285194 effects are mediated, at least in part, through its interaction with miR-211. miR-211 expression levels are inversely correlated with those of LOC285194, and miR-211 was shown to promote cell growth.189 Another study looked at global lncRNA expression changes in 9 osteosarcoma samples and adjacent normal tissue. The authors identified 403 lncRNAs that are consistently upregulated in osteosarcoma and 798 lncR-NAs that are consistently downregulated.189 The biological consequence of altered expression of these lncRNAs remains to be determined.
2. MicroRNAs
miRNAs are a key class of epigenetic regulators that act to post-transcriptionally silence large numbers of genes.185 Target genes of miRNAs expressed in osteosarcoma include members of signaling pathways that are key to osteosarcoma pathogenesis, including Ras, Wnt, mitogen-activated protein kinase, and Notch.134 In addition, miRNAs are capable of broadly affecting epigenetic regulation by silencing DNMTs and, like other transcripts, are themselves subject to epigenetic mechanisms of regulation.190 A number of studies have demonstrated that dysregulated miRNA expression results in aberrant expression of osteosarcoma genes that play critical roles in tumorigenesis and progression.191–195
miRNAs downregulated in osteosarcoma include miR-16, miR-34, miR-133a, miR-143, miR-199a-3p, miR-335, and miR-340.10,189,196–200 Many of these miRNAs are downregulated through epigenetic events in osteosarcoma specifically or in other tumor types.196,201,202 miRNAs upregulated in osteosarcoma include miR-20a, miR-29a, miR-140, and miR-181.193,195,203–205 In addition to targeting key genes in osteosarcoma signaling pathways, as described above, a number of miRNAs upregulated in osteosarcoma affect epigenetic regulatory genes. miR-140 has been shown to target histone deacetylase 4.205 miR-20a is thought to be responsible for epigenetically mediated downregulation of Fas.203 Like other oncogenes described above, there exists little to no evidence demonstrating epigenetic changes as a key driver of miRNA overexpression.134 However, this is likely explained by the fact that most epigenetic studies of osteosarcoma completed to date have focused on DNA methylation, which provides only the opportunity to detect hypomethylation as a means of gene activation and would not detect other well-described mechanisms of epigenetically mediated overexpression.
V. POTENTIAL NEW TARGETS FOR THERAPY
In osteosarcoma and other solid tumors with high rates of metastases, therapeutic targets that may most improve patient outcomes have been recognized to be those that target metastatic progression and, as such, may not have substantial activity on measurable primary tumors. Furthermore, the fact that osteosarcoma lesions are associated with a rich bone stroma that may not immediately regress concurrent with a tumor response to therapy complicates the conventional use of tumor response to identify agents that may be active in osteosarcoma. For both of these reasons, drugs with potential clinical efficacy may reasonably fail to show activity in standard phase II clinical trials, which rely on shrinkage of the primary tumor as the key metric of therapeutic response. In light of this, the osteosarcoma drug development community has recently outlined the types of preclinical data that should be prioritized as a novel therapeutic agent is considered for inclusion in the treatment of patients with osteosarcoma as a means to prevent metastatic progression, realizing that response data in human patients may not be available.206 To help assess the potential clinical utility of novel therapeutic agents, an important model/tool is provided by pet dogs that develop osteosarcoma. Indeed, studies of dogs with osteosarcoma are now underway to best define the activity of agents with the greatest promise to improve outcomes for patients. An additional resource available to the osteosarcoma preclinical/clinical research community is the data generated by the Pediatric Preclinical Testing Program, a program to systematically evaluate new agents against childhood leukemia and solid-tumor models (including osteosarcoma). Pediatric Preclinical Testing Program data are publicly available online (http://pptp.nchresearch.org/).
Table 1 presents a list of therapeutic agents that may be reasonably considered to improve treatment outcomes for patients with osteosarcoma. These agents were selected for inclusion based on their specificity for targeting the genetic and epigenetic alterations identified in osteosarcoma and presented in this article, for targeting other key osteosarcoma pathways, or for their promise in preclinical and clinical studies. For each agent listed, a subjective measure of the strength of evidence (based on our assessment) is included.
TABLE 1.
Candidate Osteosarcoma Therapeutic Agents
| Agent | Target(s) | Mechanism of Action |
Preclinical/Clinical Rationale |
Strength of Evidence* |
|---|---|---|---|---|
|
Chemotherapeutics
and small-molecule inhibitors |
||||
| Gemcitabine, aerosolized |
Chemotherapeutic agent; Fas |
Pyrimidine antimetabolite; upregulates Fas expression |
Inhibited metastasis in osteosarcoma xenograft models207; effect abolished in FasL-deficient mice157 |
Medium- high |
| RG7388 | MDM2 | Small-molecule inhibitor of p53–MDM2 interaction |
Evidence for dysregulation of p53/Mdm2 in most osteosarcoma (see text); inhibited osteosarcoma tumor growth in xenograft models208 |
Medium- high |
| PF-2341066 | Met | Small-molecule, ATP-competitive Met inhibitor |
Evidence for overexpression in osteosarcoma tissues; overexpression linked to metastasis biology; reduced primary tumor growth and metastasis in xenograft models209 |
Medium- high |
| NSC305787; NSC668394 |
Ezrin | Protein–protein interaction inhibitors, specific kinase inhibitors |
Expression associated with a less favorable outcome210; knockdown inhibited metastasis in xenograft models210; small-molecule inhibitor reduced invasive phenotype in vitro211 |
Medium |
| Vismodegib (GDC-0449) |
Hedgehog (HH) pathway |
Smoothened receptor (SMO) antagonist |
Known role in stem cell differentiation during normal bone development; known role in metastasis in other cancers212; pathway inhibition inhibits tumor growth in xenograft model213; FDA approved for treatment of other cancers214 |
Medium |
| Saracatinib (AZD0530) |
Src | Selective Src kinase inhibitor |
Reduced cell motility in vitro, no reduction of metastasis in mouse models215; phase II.5 clinical trial underway (www.clincaltrials.gov identifier NCT00752206) |
Medium- low |
| Rapamycin | mTOR | Small-molecule inhibitor |
Signaling pathway active in osteosarcoma tissues216; expression correlated with metastasis and survival216; currently in clinical trials in dogs (COTC020); prevented metastasis in xenograft models217 |
Medium |
|
Immune
modulators and antibody conjugates |
||||
| hu14.18K322A | GD2 | Humanized anti-GD2 antibody |
Ubiquitously expressed in osteosarcoma cell lines and tissues218; currently being tested in phase I clinical trials in osteosarcoma (www.clinicaltrials.gov identifier NCT00743496) |
Medium- low |
| ADXS31-164 | Her2/neu | Vaccine | Expression in osteosarcoma tumors is associated with poor survival outcomes219,220; targeted immunotherapy reduced tumor-initiating cells221; currently in dog clinical trials (www.petcancerinformation.com) |
Medium |
| Glembatumumab vedotin (CDX-011) |
GPNMB | Antibody–auristatin conjugate |
Variably expressed on surface of osteosarcoma xenografts222; significant improvement in event-free survival in osteosarcoma xenograft models222 |
Medium |
|
Epigenetic
modulators |
||||
| 5-aza-CdR (decitabine) |
CREG1, p14ARF, p21, RASSF1 |
DNMTi | Numerous genes associated with promoter hypermethylation in osteosarcoma (see text); phase I clinical trials completed222 and ongoing (www.clinicaltrials.gov identifier NCT01241162) |
Medium- low |
| Ibandronate | Ras, DNMT, Fas | Bisphosphonate; upregulates Fas |
Inhibited Ras function and downregulated DNMT, leading to increased Fas expression160; induced apoptosis in vitro160 |
Low |
| Zolendronate | Small GTPases | Bisphosphonate; downregulates VEGF |
Suppressed lung metastasis and prolonged overall survival in mouse models224,225; phase I clinical trial completed226 |
Medium- high |
| Tranylcypromine | LSD1 | Forms adduct with inactive region of LSD1 |
Expressed in osteosarcoma tissues172; reduced osteosarcoma growth in vitro172 |
Low |
| Pracinotat (SB939) | HDAC | HDACi | Phase I trial completed227 | Medium- low |
| Erinostat (MS-275) | HDAC, Fas | HDACi; Fas upregulation |
Upregulates Fas expression in Fas- metastatic osteosarcoma cells228; caused regression of metastasis in xenograft models through upregulation Fas expression in Fas- cells228 |
Medium |
| Valproic acid | HDAC | HDACi | Inhibited growth in vitro and in xenograph metastasis models in combination with doxorubicin229; phase I clinical trials ongoing (www.clinicaltrials.gov identifiers NCT01106872; NCT01010958) |
Medium- high |
Strength of evidence assessment represents a subjective measure on a scale of “low” to “high,” as judged by the authors based on efficacy in preventing or reversing metastatic progression in preclinical models as well as safety and therapeutic efficacy in dog and human clinical trials.
ACKNOWLEDGMENTS
The authors thank Drs. Lee Helman and Carol Thiele for their critical reading of the manuscript and helpful discussion.
ABBREVIATIONS
- CIN
chromosomal instability
- CpG
cytosine-phosphate-guanine
- DNMT
DNA methyltransferase
- DNMTi
DNA methyltransferase inhibitor
- HDAC
histone deacetylase
- HDACi
histone deacetylase inhibitor
- hESCs
human embryonic stem cells
- lncRNA
long non-coding RNA
- LSD1
lysine-specific demethylase 1
- miRNA
microRNA
- mTOR
mammalian target of rapamycin
- Rb
retinoblastoma
- SNP
single nucleotide polymorphism
- VEGF
vascular endothelial growth factor
REFERENCES
- 1.Klein MJ, Siegal GP. Osteosarcoma: anatomic and histologic variants. Am J Clin Pathol. 2006;125(4):555–81. doi: 10.1309/UC6K-QHLD-9LV2-KENN. [DOI] [PubMed] [Google Scholar]
- 2.What are the key statistics about osteosarcoma–Atlanta: American Cancer Society. [©2015; cited 2015 March 3]. Available at http://www.cancer.org/cancer/osteosarcoma/detailedguide/osteosarcoma-key-statistics.
- 3.Geller DS, Gorlick R. Osteosarcoma: a review of diagnosis, management, and treatment strategies. Clin Adv Hematol Oncol. 2010;8(10):705–18. [PubMed] [Google Scholar]
- 4.Tang N, Song WX, Luo J, Haydon RC, He TC. Osteosarcoma development and stem cell differentiation. Clin Orthop Relat Res. 2008;466(9):2114–30. doi: 10.1007/s11999-008-0335-z. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 5.Martin JW, Squire JA, Zielenska M. The genetics of osteosarcoma. Sarcoma. 2012;2012:627254. doi: 10.1155/2012/627254. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 6.Stephens PJ, Greenman CD, Fu B, Yang F, Bignell GR, Mudie LJ, Pleasance ED, Lau KW, Beare D, Stebbings LA, McLaren S, Lin ML, McBride DJ, Varela I, Nik-Zainal S, Leroy C, Jia M, Menzies A, Butler AP, Teague JW, Quail MA, Burton J, Swerdlow H, Carter NP, Morsberger LA, Iacobuzio-Donahue C, Follows GA, Green AR, Flanagan AM, Stratton MR, Futreal PA, Campbell PJ. Massive genomic rearrangement acquired in a single catastrophic event during cancer development. Cell. 2011;144(1):27–40. doi: 10.1016/j.cell.2010.11.055. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 7.Mutsaers AJ, Walkley CR. 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. [DOI] [PubMed] [Google Scholar]
- 8.Berman SD, Calo E, Landman AS, Danielian PS, Miller ES, West JC, Fonhoue BD, Caron A, Bronson R, Bouxsein ML, Mukherjee S, Lees JA. Metastatic osteosarcoma induced by inactivation of Rb and p53 in the osteoblast lineage. Proc Natl Acad Sci U S A. 2008;105(33):11851–6. doi: 10.1073/pnas.0805462105. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 9.Walkley CR, Qudsi R, Sankaran VG, Perry JA, Gostissa M, Roth SI, Rodda SJ, Snay E, Dunning P, Fahey FH, Alt FW, McMahon AP, Orkin SH. Conditional mouse osteosarcoma, dependent on p53 loss and potentiated by loss of Rb, mimics the human disease. Genes Dev. 2008;22(12):1662–76. doi: 10.1101/gad.1656808. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 10.Mutsaers AJ, Ng AJ, Baker EK, Russell MR, Chalk AM, Wall M, Liddicoat BJ, Ho PW, Slavin JL, Goradia A, Martin TJ, Purton LE, Dickins RA, Walkley CR. Modeling distinct osteosarcoma subtypes in vivo using Cre:lox and lineage-restricted transgenic shRNA. Bone. 2013;55(1):166–78. doi: 10.1016/j.bone.2013.02.016. [DOI] [PubMed] [Google Scholar]
- 11.Lin PP, Pandey MK, Jin F, Raymond AK, 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):1789–95. doi: 10.1093/carcin/bgp180. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 12.Shimizu T, Ishikawa T, Sugihara E, Kuninaka S, Miyamoto T, Mabuchi Y, Matsuzaki Y, Tsunoda T, Miya F, Morioka H, Nakayama R, Kobayashi E, Toyama Y, Kawai A, Ichikawa H, Hasegawa T, Okada S, Ito T, Ikeda Y, Suda T, Saya H. c-MYC overexpression with loss of Ink4a/Arf transforms bone marrow stromal cells into osteosarcoma accompanied by loss of adipogenesis. Oncogene. 2010;29(42):5687–99. doi: 10.1038/onc.2010.312. [DOI] [PubMed] [Google Scholar]
- 13.Leonard P, Sharp T, Henderson S, Hewitt D, Pringle J, Sandison A, Goodship A, Whelan J, Boshoff C. Gene expression array profile of human osteosarcoma. Br J Cancer. 2003;89(12):2284–8. doi: 10.1038/sj.bjc.6601389. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 14.Man TK, Chintagumpala M, Visvanathan J, Shen J, Perlaky L, Hicks J, Johnson M, Davino N, Murray J, Helman L, Meyer W, Triche T, Wong KK, Lau CC. Expression profiles of osteosarcoma that can predict response to chemotherapy. Cancer Res. 2005;65(18):8142–50. doi: 10.1158/0008-5472.CAN-05-0985. [DOI] [PubMed] [Google Scholar]
- 15.Baird K, Davis S, Antonescu CR, Harper UL, Walker RL, Chen Y, Glatfelter AA, Duray PH, Meltzer PS. Gene expression profiling of human sarcomas: insights into sarcoma biology. Cancer Res. 2005;65(20):9226–35. doi: 10.1158/0008-5472.CAN-05-1699. [DOI] [PubMed] [Google Scholar]
- 16.Mirabello L, Troisi RJ, Savage SA. International osteosarcoma incidence patterns in children and adolescents, middle ages and elderly persons. Int J Cancer. 2009;125(1):229–34. doi: 10.1002/ijc.24320. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 17.Borinstein SC, Barkauskas DA, Bernstein M, Goorin A, Gorlick R, Krailo M, Schwartz CL, Wexler LH, Toretsky JA. Analysis of serum insulin growth factor-1 concentrations in localized osteosarcoma: a children’s oncology group study. Pediatr Blood Cancer. 2014;61(4):749–52. doi: 10.1002/pbc.24778. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 18.Valastyan S, Weinberg RA. Tumor metastasis: molecular insights and evolving paradigms. Cell. 2011;147(2):275–92. doi: 10.1016/j.cell.2011.09.024. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 19.Koop S, MacDonald IC, Luzzi K, Schmidt EE, Morris VL, Grattan M, Khokha R, Chambers AF, Groom AC. Fate of melanoma cells entering the microcirculation: over 80% survive and extravasate. Cancer Res. 1995;55(12):2520–3. [PubMed] [Google Scholar]
- 20.Luzzi KJ, MacDonald IC, Schmidt EE, Kerkvliet N, Morris VL, Chambers AF, Groom AC. Multistep nature of metastatic inefficiency: dormancy of solitary cells after successful extravasation and limited survival of early micrometastases. Am J Pathol. 1998;153(3):865–73. doi: 10.1016/S0002-9440(10)65628-3. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 21.Cameron MD, Schmidt EE, Kerkvliet N, Nadkarni KV, Morris VL, Groom AC, Chambers AF, MacDonald IC. Temporal progression of metastasis in lung: cell survival, dormancy, and location dependence of metastatic inefficiency. Cancer Res. 2000;60(9):2541–6. [PubMed] [Google Scholar]
- 22.Weiss L. Metastatic inefficiency. Adv Cancer Res. 1990;54:159–211. doi: 10.1016/s0065-230x(08)60811-8. [DOI] [PubMed] [Google Scholar]
- 23.Hong SH, Ren L, Mendoza A, Eleswarapu A, Khanna C. Apoptosis resistance and PKC signaling: distinguishing features of high and low metastatic cells. Neoplasia. 2012;14(3):249–58. doi: 10.1593/neo.111498. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 24.Oskarsson T, Batlle E, Massague J. Metastatic stem cells: sources, niches, and vital pathways. Cell Stem Cell. 2014;14(3):306–21. doi: 10.1016/j.stem.2014.02.002. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 25.Talmadge JE, Fidler IJ. AACR centennial series: the biology of cancer metastasis: historical perspective. Cancer Res. 2010;70(14):5649–69. doi: 10.1158/0008-5472.CAN-10-1040. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 26.Bacci G, Mercuri M, Briccoli A, Ferrari S, Bertoni F, Donati D, Monti C, Zanoni A, Forni C, Manfrini M. Osteogenic sarcoma of the extremity with detectable lung metastases at presentation. Results of treatment of 23 patients with chemotherapy followed by simultaneous resection of primary and metastatic lesions. Cancer. 1997;79(2):245–54. [PubMed] [Google Scholar]
- 27.Limb-sparing treatment of adult soft-tissue sarcomas and osteosarcomas. National Institutes of Health Consensus Development Conference Statement; Natl Inst Health Consens Dev Conf Consens Statement; 1985. p. 18. [PubMed] [Google Scholar]
- 28.Bielack SS, Kempf-Bielack B, Delling G, Exner GU, Flege S, Helmke K, Kotz R, Salzer-Kuntschik M, Werner M, Winkelmann W, Zoubek A, Jurgens H, Winkler K. Prognostic factors in high-grade osteosarcoma of the extremities or trunk: an analysis of 1,702 patients treated on neoadjuvant cooperative osteosarcoma study group protocols. J Clin Oncol. 2002;20(3):776–90. doi: 10.1200/JCO.2002.20.3.776. [DOI] [PubMed] [Google Scholar]
- 29.Kempf-Bielack B, Bielack SS, Jurgens H, Branscheid D, Berdel WE, Exner GU, Gobel U, Helmke K, Jundt G, Kabisch H, Kevric M, Klingebiel T, Kotz R, Maas R, Schwarz R, Semik M, Treuner J, Zoubek A, Winkler K. Osteosarcoma relapse after combined modality therapy: an analysis of unselected patients in the Cooperative Osteosarcoma Study Group (COSS) J Clin Oncol. 2005;23(3):559–68. doi: 10.1200/JCO.2005.04.063. [DOI] [PubMed] [Google Scholar]
- 30.McCarthy MI, Abecasis GR, Cardon LR, Goldstein DB, Little J, Ioannidis JP, Hirschhorn JN. Genome-wide association studies for complex traits: consensus, uncertainty and challenges. Nat Rev Genet. 2008;9(5):356–69. doi: 10.1038/nrg2344. [DOI] [PubMed] [Google Scholar]
- 31.Savage SA, Mirabello L, Wang Z, Gastier-Foster JM, Gorlick R, Khanna C, Flanagan AM, Tirabosco R, Andrulis IL, Wunder JS, Gokgoz N, Patino-Garcia A, Sierrasesumaga L, Lecanda F, Kurucu N, Ilhan IE, Sari N, Serra M, Hattinger C, Picci P, Spector LG, Barkauskas DA, Marina N, de Toledo SR, Petrilli AS, Amary MF, Halai D, Thomas DM, Douglass C, Meltzer PS, Jacobs K, Chung CC, Berndt SI, Purdue MP, Caporaso NE, Tucker M, Rothman N, Landi MT, Silverman DT, Kraft P, Hunter DJ, Malats N, Kogevinas M, Wacholder S, Troisi R, Helman L, Fraumeni JF, Jr., Yeager M, Hoover RN, Chanock SJ. Genome-wide association study identifies two susceptibility loci for osteosarcoma. Nat Genet. 2013;45(7):799–803. doi: 10.1038/ng.2645. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 32.Karlsson EK, Sigurdsson S, Ivansson E, Thomas R, Elvers I, Wright J, Howald C, Tonomura N, Perloski M, Swofford R, Biagi T, Fryc S, Anderson N, Courtay-Cahen C, Youell L, Ricketts SL, Mandlebaum S, Rivera P, von Euler H, Kisseberth WC, London CA, Lander ES, Couto G, Comstock K, Starkey MP, Modiano JF, Breen M, Lindblad-Toh K. Genome-wide analyses implicate 33 loci in heritable dog osteosarcoma, including regulatory variants near CDKN2A/B. Genome Biol. 2013;14(12):R132. doi: 10.1186/gb-2013-14-12-r132. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 33.Patino-Garcia A, Sotillo-Pieiro E, Modesto C, Sierrases-Maga L. Analysis of the human tumour necrosis factor-alpha (TNFalpha) gene promoter polymorphisms in children with bone cancer. J Med Genet. 2000;37(10):789–92. doi: 10.1136/jmg.37.10.789. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 34.Savage SA, Burdett L, Troisi R, Douglass C, Hoover RN, Chanock SJ. Germ-line genetic variation of TP53 in osteosarcoma. Pediatr Blood Cancer. 2007;49(1):28–33. doi: 10.1002/pbc.21077. [DOI] [PubMed] [Google Scholar]
- 35.Koshkina NV, Kleinerman ES, Li G, Zhao CC, Wei Q, Sturgis EM. Exploratory analysis of Fas gene polymorphisms in pediatric osteosarcoma patients. J Pediatr Hematol Oncol. 2007;29(12):815–21. doi: 10.1097/MPH.0b013e3181581506. [DOI] [PubMed] [Google Scholar]
- 36.Toffoli G, Biason P, Russo A, De Mattia E, Cecchin E, Hattinger CM, Pasello M, Alberghini M, Ferrari C, Scotlandi K, Picci P, Serra M. Effect of TP53 Arg72Pro and MDM2 SNP309 polymorphisms on the risk of high-grade osteosarcoma development and survival. Clin Cancer Res. 2009;15(10):3550–6. doi: 10.1158/1078-0432.CCR-08-2249. [DOI] [PubMed] [Google Scholar]
- 37.Hu YS, Pan Y, Li WH, Zhang Y, Li J, Ma BA. Association between TGFBR1*6A and osteosarcoma: a Chinese case-control study. BMC Cancer. 2010;10:169. doi: 10.1186/1471-2407-10-169. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 38.Mirabello L, Berndt SI, Seratti GF, Burdett L, Yeager M, Chowdhury S, Teshome K, Uzoka A, Douglass C, Hayes RB, Hoover RN, Savage SA. Genetic variation at chromosome 8q24 in osteosarcoma cases and controls. Carcinogenesis. 2010;31(8):1400–4. doi: 10.1093/carcin/bgq117. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 39.Mirabello L, Yu K, Berndt SI, Burdett L, Wang Z, Chowdhury S, Teshome K, Uzoka A, Hutchinson A, Grotmol T, Douglass C, Hayes RB, Hoover RN, Savage SA. A comprehensive candidate gene approach identifies genetic variation associated with osteosarcoma. BMC Cancer. 2011;11:209. doi: 10.1186/1471-2407-11-209. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 40.Molyneux SD, Di Grappa MA, Beristain AG, McKee TD, Wai DH, Paderova J, Kashyap M, Hu P, Maiuri T, Narala SR, Stambolic V, Squire J, Penninger J, Sanchez O, Triche TJ, Wood GA, Kirschner LS, Khokha R. Prkar1a is an osteosarcoma tumor suppressor that defines a molecular subclass in mice. J Clin Invest. 2010;120(9):3310–25. doi: 10.1172/JCI42391. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 41.Griffin KJ, Kirschner LS, Matyakhina L, Stergiopoulos SG, Robinson-White A, Lenherr SM, Weinberg FD, Claflin ES, Batista D, Bourdeau I, Voutetakis A, Sandrini F, Meoli EM, Bauer AJ, Cho-Chung YS, Bornstein SR, Carney JA, Stratakis CA. A transgenic mouse bearing an antisense construct of regulatory subunit type 1A of protein kinase A develops endocrine and other tumours: comparison with Carney complex and other PRKAR1A induced lesions. J Med Genet. 2004;41(12):923–31. doi: 10.1136/jmg.2004.028043. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 42.Mueller F, Fuchs B, Kaser-Hotz B. Comparative biology of human and canine osteosarcoma. Anticancer Res. 2007;27(1A):155–64. [PubMed] [Google Scholar]
- 43.Hauben EI, Arends J, Vandenbroucke JP, van Asperen CJ, Van Marck E, Hogendoorn PC. Multiple primary malignancies in osteosarcoma patients. Incidence and predictive value of osteosarcoma subtype for cancer syndromes related with osteosarcoma. Eur J Hum Genet. 2003;11(8):611–8. doi: 10.1038/sj.ejhg.5201012. [DOI] [PubMed] [Google Scholar]
- 44.Selvarajah S, Yoshimoto M, Maire G, Paderova J, Bayani J, Squire JA, Zielenska M. Identification of cryptic microaberrations in osteosarcoma by high-definition oligonucleotide array comparative genomic hybridization. Cancer Genet Cytogenet. 2007;179(1):52–61. doi: 10.1016/j.cancergencyto.2007.08.003. [DOI] [PubMed] [Google Scholar]
- 45.Selvarajah S, Yoshimoto M, Park PC, Maire G, Paderova J, Bayani J, Lim G, Al-Romaih K, Squire JA, Zielenska M. The breakage-fusion-bridge (BFB) cycle as a mechanism for generating genetic heterogeneity in osteosarcoma. Chromosoma. 2006;115(6):459–67. doi: 10.1007/s00412-006-0074-4. [DOI] [PubMed] [Google Scholar]
- 46.Geigl JB, Obenauf AC, Schwarzbraun T, Speicher MR. Defining ’chromosomal instability’. Trends Genet. 2008;24(2):64–9. doi: 10.1016/j.tig.2007.11.006. [DOI] [PubMed] [Google Scholar]
- 47.Lengauer C, Kinzler KW, Vogelstein B. Genetic instabilities in human cancers. Nature. 1998;396(6712):643–9. doi: 10.1038/25292. [DOI] [PubMed] [Google Scholar]
- 48.van Harn T, Foijer F, van Vugt M, Banerjee R, Yang F, Oostra A, Joenje H, te Riele H. Loss of Rb proteins causes genomic instability in the absence of mitogenic signaling. Genes Dev. 2010;24(13):1377–88. doi: 10.1101/gad.580710. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 49.Weiss MB, Vitolo MI, Mohseni M, Rosen DM, Denmeade SR, Park BH, Weber DJ, Bachman KE. Deletion of p53 in human mammary epithelial cells causes chromosomal instability and altered therapeutic response. Oncogene. 2010;29(33):4715–24. doi: 10.1038/onc.2010.220. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 50.Scheel C, Schaefer KL, Jauch A, Keller M, Wai D, Brinkschmidt C, van Valen F, Boecker W, Dockhorn-Dworniczak B, Poremba C. Alternative lengthening of telomeres is associated with chromosomal instability in osteosarcomas. Oncogene. 2001;20(29):3835–44. doi: 10.1038/sj.onc.1204493. [DOI] [PubMed] [Google Scholar]
- 51.Henson JD, Hannay JA, McCarthy SW, Royds JA, Yeager TR, Robinson RA, Wharton SB, Jellinek DA, Arbuckle SM, Yoo J, Robinson BG, Learoyd DL, Stalley PD, Bonar SF, Yu D, Pollock RE, Reddel RR. A robust assay for alternative lengthening of telomeres in tumors shows the significance of alternative lengthening of telomeres in sarcomas and astrocytomas. Clin Cancer Res. 2005;11(1):217–25. [PubMed] [Google Scholar]
- 52.Sandberg AA, Bridge JA. Updates on the cytogenetics and molecular genetics of bone and soft tissue tumors: osteosarcoma and related tumors. Cancer Genet Cytogenet. 2003;145(1):1–30. [PubMed] [Google Scholar]
- 53.Helman LJ, Meltzer P. Mechanisms of sarcoma development. Nat Rev Cancer. 2003;3(9):685–94. doi: 10.1038/nrc1168. [DOI] [PubMed] [Google Scholar]
- 54.Cormier JN, Pollock RE. Soft tissue sarcomas. CA Cancer J Clin. 2004;54(2):94–109. doi: 10.3322/canjclin.54.2.94. [DOI] [PubMed] [Google Scholar]
- 55.Hayden JB, Hoang BH. Osteosarcoma: basic science and clinical implications. Orthop Clin North Am. 2006;37(1):1–7. doi: 10.1016/j.ocl.2005.06.004. [DOI] [PubMed] [Google Scholar]
- 56.Haydon RC, Luu HH, He TC. Osteosarcoma and osteoblastic differentiation: a new perspective on oncogenesis. Clin Orthop Relat Res. 2007;454:237–46. doi: 10.1097/BLO.0b013e31802b683c. [DOI] [PubMed] [Google Scholar]
- 57.Kansara M, Thomas DM. Molecular pathogenesis of osteosarcoma. DNA Cell Biol. 2007;26(1):1–18. doi: 10.1089/dna.2006.0505. [DOI] [PubMed] [Google Scholar]
- 58.Marina N, Gebhardt M, Teot L, Gorlick R. Biology and therapeutic advances for pediatric osteosarcoma. Oncologist. 2004;9(4):422–41. doi: 10.1634/theoncologist.9-4-422. [DOI] [PubMed] [Google Scholar]
- 59.Nevins JR. The Rb/E2F pathway and cancer. Hum Mol Genet. 2001;10(7):699–703. doi: 10.1093/hmg/10.7.699. [DOI] [PubMed] [Google Scholar]
- 60.Quelle DE, Zindy F, Ashmun RA, Sherr CJ. Alternative reading frames of the INK4a tumor suppressor gene encode two unrelated proteins capable of inducing cell cycle arrest. Cell. 1995;83(6):993–1000. doi: 10.1016/0092-8674(95)90214-7. [DOI] [PubMed] [Google Scholar]
- 61.Miller CW, Aslo A, Won A, Tan M, Lampkin B, Koeffler HP. Alterations of the p53, Rb and MDM2 genes in osteosarcoma. J Cancer Res Clin Oncol. 1996;122(9):559–65. doi: 10.1007/BF01213553. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 62.Belchis DA, Meece CA, Benko FA, Rogan PK, Williams RA, Gocke CD. Loss of heterozygosity and microsatellite instability at the retinoblastoma locus in osteosarcomas. Diagn Mol Pathol. 1996;5(3):214–9. doi: 10.1097/00019606-199609000-00011. [DOI] [PubMed] [Google Scholar]
- 63.Benassi MS, Molendini L, Gamberi G, Ragazzini P, Sollazzo MR, Merli M, Asp J, Magagnoli G, Balladelli A, Bertoni F, Picci P. Alteration of pRb/p16/cdk4 regulation in human osteosarcoma. Int J Cancer. 1999;84(5):489–93. doi: 10.1002/(sici)1097-0215(19991022)84:5<489::aid-ijc7>3.0.co;2-d. [DOI] [PubMed] [Google Scholar]
- 64.Alonso J, Garcia-Miguel P, Abelairas J, Mendiola M, Pestana A. A microsatellite fluorescent method for linkage analysis in familial retinoblastoma and deletion detection at the RB1 locus in retinoblastoma and osteosarcoma. Diagn Mol Pathol. 2001;10(1):9–14. doi: 10.1097/00019606-200103000-00003. [DOI] [PubMed] [Google Scholar]
- 65.Yamaguchi T, Toguchida J, Yamamuro T, Kotoura Y, Takada N, Kawaguchi N, Kaneko Y, Nakamura Y, Sasaki MS, Ishizaki K. Allelotype analysis in osteosarcomas: frequent allele loss on 3q, 13q, 17p, and 18q. Cancer Res. 1992;52(9):2419–23. [PubMed] [Google Scholar]
- 66.Feugeas O, Guriec N, Babin-Boilletot A, Marcellin L, Simon P, Babin S, Thyss A, Hofman P, Terrier P, Kalifa C, Brunat-Mentigny M, Patricot LM, Oberling F. Loss of heterozygosity of the RB gene is a poor prognostic factor in patients with osteosarcoma. J Clin Oncol. 1996;14(2):467–72. doi: 10.1200/JCO.1996.14.2.467. [DOI] [PubMed] [Google Scholar]
- 67.Araki N, Uchida A, Kimura T, Yoshikawa H, Aoki Y, Ueda T, Takai S, Miki T, Ono K. Involvement of the retinoblastoma gene in primary osteosarcomas and other bone and soft-tissue tumors. Clin Orthop Relat Res. 1991;(270):271–7. [PubMed] [Google Scholar]
- 68.Scholz RB, Kabisch H, Weber B, Roser K, Delling G, Winkler K. Studies of the RB1 gene and the p53 gene in human osteosarcomas. Pediatr Hematol Oncol. 1992;9(2):125–37. doi: 10.3109/08880019209018328. [DOI] [PubMed] [Google Scholar]
- 69.Wunder JS, Czitrom AA, Kandel R, Andrulis IL. Analysis of alterations in the retinoblastoma gene and tumor grade in bone and soft-tissue sarcomas. J Natl Cancer Inst. 1991;83(3):194–200. doi: 10.1093/jnci/83.3.194. [DOI] [PubMed] [Google Scholar]
- 70.Toguchida J, Ishizaki K, Sasaki MS, Ikenaga M, Sugimoto M, Kotoura Y, Yamamuro T. Chromosomal reorganization for the expression of recessive mutation of retinoblastoma susceptibility gene in the development of osteosarcoma. Cancer Res. 1988;48(14):3939–43. [PubMed] [Google Scholar]
- 71.Wadayama B, Toguchida J, Shimizu T, Ishizaki K, Sasaki MS, Kotoura Y, Yamamuro T. Mutation spectrum of the retinoblastoma gene in osteosarcomas. Cancer Res. 1994;54(11):3042–8. [PubMed] [Google Scholar]
- 72.Miller CW, Aslo A, Campbell MJ, Kawamata N, Lampkin BC, Koeffler HP. Alterations of the p15, p16,and p18 genes in osteosarcoma. Cancer Genet Cytogenet. 1996;86(2):136–42. doi: 10.1016/0165-4608(95)00216-2. [DOI] [PubMed] [Google Scholar]
- 73.Khanna C, Khan J, Nguyen P, Prehn J, Caylor J, Yeung C, Trepel J, Meltzer P, Helman L. Metastasis-associated differences in gene expression in a murine model of osteosarcoma. Cancer Res. 2001;61(9):3750–9. [PubMed] [Google Scholar]
- 74.Harris SL, Levine AJ. The p53 pathway: positive and negative feedback loops. Oncogene. 2005;24(17):2899–908. doi: 10.1038/sj.onc.1208615. [DOI] [PubMed] [Google Scholar]
- 75.Nigro JM, Baker SJ, Preisinger AC, Jessup JM, Hostetter R, Cleary K, Bigner SH, Davidson N, Baylin S, Devilee P, Glover T, Collins FS, Weslon A, Modali R, Harris CC, Vogelstein B. Mutations in the p53 gene occur in diverse human tumour types. Nature. 1989;342(6250):705–8. doi: 10.1038/342705a0. [DOI] [PubMed] [Google Scholar]
- 76.Andreassen A, Oyjord T, Hovig E, Holm R, Florenes VA, Nesland JM, Myklebost O, Hoie J, Bruland OS, Borresen AL, Fodstad Ø . p53 abnormalities in different subtypes of human sarcomas. Cancer Res. 1993;53(3):468–71. [PubMed] [Google Scholar]
- 77.Castresana JS, Rubio MP, Gomez L, Kreicbergs A, Zetterberg A, Barrios C. Detection of TP53 gene mutations in human sarcomas. Eur J Cancer. 1995;31A(5):735–8. doi: 10.1016/0959-8049(95)00121-x. [DOI] [PubMed] [Google Scholar]
- 78.Masuda H, Miller C, Koeffler HP, Battifora H, Cline MJ. Rearrangement of the p53 gene in human osteogenic sarcomas. Proc Natl Acad Sci U S A. 1987;84(21):7716–9. doi: 10.1073/pnas.84.21.7716. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 79.Mulligan LM, Matlashewski GJ, Scrable HJ, Cavenee WK. Mechanisms of p53 loss in human sarcomas. Proc Natl Acad Sci U S A. 1990;87(15):5863–7. doi: 10.1073/pnas.87.15.5863. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 80.Mousses S, McAuley L, Bell RS, Kandel R, Andrulis IL. Molecular and immunohistochemical identification of p53 alterations in bone and soft tissue sarcomas. Mod Pathol. 1996;9(1):1–6. [PubMed] [Google Scholar]
- 81.Sztan M, Papai Z, Szendroi M, Looij M, Olah E. Allelic losses from chromosome 17 in human osteosarcomas. Pathol Oncol Res. 1997;3(2):115–20. doi: 10.1007/BF02907805. [DOI] [PubMed] [Google Scholar]
- 82.Radig K, Schneider-Stock R, Haeckel C, Neumann W, Roessner A. p53 gene mutations in osteosarcomas of low-grade malignancy. Hum Pathol. 1998;29(11):1310–6. doi: 10.1016/s0046-8177(98)90263-5. [DOI] [PubMed] [Google Scholar]
- 83.Toguchida J, Yamaguchi T, Dayton SH, Beauchamp RL, Herrera GE, Ishizaki K, Yamamuro T, Meyers PA, Little JB, Sasaki MS, Weichselbaum RR, Yandell DW. Prevalence and spectrum of germline mutations of the p53 gene among patients with sarcoma. N Engl J Med. 1992;326(20):1301–8. doi: 10.1056/NEJM199205143262001. [DOI] [PubMed] [Google Scholar]
- 84.Mirabello L, et al. Germline TP53 variants and susceptibility to osteosarcoma. JNCI. 2015 doi: 10.1093/jnci/djv101. Online April 20, 2015. PMID: 25896519. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 85.Entz-Werle N, Lavaux T, Metzger N, Stoetzel C, Lasthaus C, Marec P, Kalifa C, Brugieres L, Pacquement H, Schmitt C, Tabone MD, Gentet JC, Lutz P, Babin A, Oudet P, Gaub MP, Perrin-Schmitt F. Involvement of MET/TWIST/APC combination or the potential role of ossification factors in pediatric high-grade osteosarcoma oncogenesis. Neoplasia. 2007;9(8):678–88. doi: 10.1593/neo.07367. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 86.Mendoza S, David H, Gaylord GM, Miller CW. Allelic loss at 10q26 in osteosarcoma in the region of the BUB3 and FGFR2 genes. Cancer Genet Cytogenet. 2005;158(2):142–7. doi: 10.1016/j.cancergencyto.2004.08.035. [DOI] [PubMed] [Google Scholar]
- 87.Patino-Garcia A, Pineiro ES, Diez MZ, Iturriagagoitia LG, Klussmann FA, Ariznabarreta LS. Genetic and epigenetic alterations of the cell cycle regulators and tumor suppressor genes in pediatric osteosarcomas. J Pediatr Hematol Oncol. 2003;25(5):362–7. doi: 10.1097/00043426-200305000-00003. [DOI] [PubMed] [Google Scholar]
- 88.Smida J, Baumhoer D, Rosemann M, Walch A, Bielack S, Poremba C, Remberger K, Korsching E, Scheurlen W, Dierkes C, Burdach S, Jundt G, Atkinson MJ, Nathrath M. Genomic alterations and allelic imbalances are strong prognostic predictors in osteosarcoma. Clin Cancer Res. 2010;16(16):4256–67. doi: 10.1158/1078-0432.CCR-10-0284. [DOI] [PubMed] [Google Scholar]
- 89.Kresse SH, Ohnstad HO, Paulsen EB, Bjerkehagen B, Szuhai K, Serra M, Schaefer KL, Myklebost O, Meza-Zepeda LA. LSAMP, a novel candidate tumor suppressor gene in human osteosarcomas, identified by array comparative genomic hybridization. Genes Chromosomes Cancer. 2009;48(8):679–93. doi: 10.1002/gcc.20675. [DOI] [PubMed] [Google Scholar]
- 90.Ozaki T, Schaefer KL, Wai D, Buerger H, Flege S, Lindner N, Kevric M, Diallo R, Bankfalvi A, Brinkschmidt C, Juergens H, Winkelmann W, Dockhorn-Dworniczak B, Bielack SS, Poremba C. Genetic imbalances revealed by comparative genomic hybridization in osteosarcomas. Int J Cancer. 2002;102(4):355–65. doi: 10.1002/ijc.10709. [DOI] [PubMed] [Google Scholar]
- 91.Pasic I, Shlien A, Durbin AD, Stavropoulos DJ, Baskin B, Ray PN, Novokmet A, Malkin D. Recurrent focal copy-number changes and loss of heterozygosity implicate two noncoding RNAs and one tumor suppressor gene at chromosome 3q13.31 in osteosarcoma. Cancer Res. 2010;70(1):160–71. doi: 10.1158/0008-5472.CAN-09-1902. [DOI] [PubMed] [Google Scholar]
- 92.Tarkkanen M, Karhu R, Kallioniemi A, Elomaa I, Kivioja AH, Nevalainen J, Bohling T, Karaharju E, Hyytinen E, Knuutila S, Kallioniemi OP. Gains and losses of DNA sequences in osteosarcomas by comparative genomic hybridization. Cancer Res. 1995;55(6):1334–8. [PubMed] [Google Scholar]
- 93.Yen CC, Chen WM, Chen TH, Chen WY, Chen PC, Chiou HJ, Hung GY, Wu HT, Wei CJ, Shiau CY, Wu YC, Chao TC, Tzeng CH, Chen PM, Lin CH, Chen YJ, Fletcher JA. Identification of chromosomal aberrations associated with disease progression and a novel 3q13.31 deletion involving LSAMP gene in osteosarcoma. Int J Oncol. 2009;35(4):775–88. doi: 10.3892/ijo_00000390. [DOI] [PubMed] [Google Scholar]
- 94.Zielenska M, Bayani J, Pandita A, Toledo S, Marrano P, Andrade J, Petrilli A, Thorner P, Sorensen P, Squire JA. Comparative genomic hybridization analysis identifies gains of 1p35 approximately p36 and chromosome 19 in osteosarcoma. Cancer Genet Cytogenet. 2001;130(1):14–21. doi: 10.1016/s0165-4608(01)00461-7. [DOI] [PubMed] [Google Scholar]
- 95.Nishijo K, Nakayama T, Aoyama T, Okamoto T, Ishibe T, Yasura K, Shima Y, Shibata KR, Tsuboyama T, Nakamura T, Toguchida J. Mutation analysis of the RECQL4 gene in sporadic osteosarcomas. Int J Cancer. 2004;111(3):367–72. doi: 10.1002/ijc.20269. [DOI] [PubMed] [Google Scholar]
- 96.Maire G, Yoshimoto M, Chilton-MacNeill S, Thorner PS, Zielenska M, Squire JA. Recurrent RECQL4 imbalance and increased gene expression levels are associated with structural chromosomal instability in sporadic osteosarcoma. Neoplasia. 2009;11(3):260–8. doi: 10.1593/neo.81384. 3p following 68. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 97.Yang J, Cogdell D, Yang D, Hu L, Li H, Zheng H, Du X, Pang Y, Trent J, Chen K, Zhang W. Deletion of the WWOX gene and frequent loss of its protein expression in human osteosarcoma. Cancer Lett. 2010;291(1):31–8. doi: 10.1016/j.canlet.2009.09.018. [DOI] [PubMed] [Google Scholar]
- 98.Martin JW, Yoshimoto M, Ludkovski O, Thorner PS, Zielenska M, Squire JA, Nuin PA. Analysis of segmental duplications, mouse genome synteny and recurrent cancer-associated amplicons in human chromosome 6p21-p12. Cytogenet Genome Res. 2010;128(4):199–213. doi: 10.1159/000308353. [DOI] [PubMed] [Google Scholar]
- 99.Mejia-Guerrero S, Quejada M, Gokgoz N, Gill M, Parkes RK, Wunder JS, Andrulis IL. Characterization of the 12q15 MDM2 and 12q13-14 CDK4 amplicons and clinical correlations in osteosarcoma. Genes Chromosomes Cancer. 2010;49(6):518–25. doi: 10.1002/gcc.20761. [DOI] [PubMed] [Google Scholar]
- 100.Lopez-Guerrero JA, Lopez-Gines C, Pellin A, Carda C, Llombart-Bosch A. Deregulation of the G1 to S-phase cell cycle checkpoint is involved in the pathogenesis of human osteosarcoma. Diagn Mol Pathol. 2004;13(2):81–91. doi: 10.1097/00019606-200406000-00004. [DOI] [PubMed] [Google Scholar]
- 101.Duhamel LA, Ye H, Halai D, Idowu BD, Presneau N, Tirabosco R, Flanagan AM. Frequency of mouse double minute 2 (MDM2) and mouse double minute 4 (MDM4) amplification in parosteal and conventional osteosarcoma subtypes. Histopathology. 2012;60(2):357–9. doi: 10.1111/j.1365-2559.2011.04023.x. [DOI] [PubMed] [Google Scholar]
- 102.Lonardo F, Ueda T, Huvos AG, Healey J, Ladanyi M. p53 and MDM2 alterations in osteosarcomas: correlation with clinicopathologic features and proliferative rate. Cancer. 1997;79(8):1541–7. [PubMed] [Google Scholar]
- 103.Bayani J, Zielenska M, Pandita A, Al-Romaih K, Karaskova J, Harrison K, Bridge JA, Sorensen P, Thorner P, Squire JA. Spectral karyotyping identifies recurrent complex rearrangements of chromosomes 8, 17, and 20 in osteosarcomas. Genes Chromosomes Cancer. 2003;36(1):7–16. doi: 10.1002/gcc.10132. [DOI] [PubMed] [Google Scholar]
- 104.Lau CC, Harris CP, Lu XY, Perlaky L, Gogineni S, Chintagumpala M, Hicks J, Johnson ME, Davino NA, Huvos AG, Meyers PA, Healy JH, Gorlick R, Rao PH. Frequent amplification and rearrangement of chromosomal bands 6p12-p21 and 17p11.2 in osteosarcoma. Genes Chromosomes Cancer. 2004;39(1):11–21. doi: 10.1002/gcc.10291. [DOI] [PubMed] [Google Scholar]
- 105.Henriksen J, Aagesen TH, Maelandsmo GM, Lothe RA, Myklebost O, Forus A. Amplification and overexpression of COPS3 in osteosarcomas potentially target TP53 for proteasome-mediated degradation. Oncogene. 2003;22(34):5358–61. doi: 10.1038/sj.onc.1206671. [DOI] [PubMed] [Google Scholar]
- 106.Yan T, Wunder JS, Gokgoz N, Gill M, Eskandarian S, Parkes RK, Bull SB, Bell RS, Andrulis IL. COPS3 amplification and clinical outcome in osteosarcoma. Cancer. 2007;109(9):1870–6. doi: 10.1002/cncr.22595. [DOI] [PubMed] [Google Scholar]
- 107.Niehrs C, Schafer A. Active DNA demethylation by Gadd45 and DNA repair. Trends Cell Biol. 2012;22(4):220–7. doi: 10.1016/j.tcb.2012.01.002. [DOI] [PubMed] [Google Scholar]
- 108.Lin CY, Loven J, Rahl PB, Paranal RM, Burge CB, Bradner JE, Lee TI, Young RA. Transcriptional amplification in tumor cells with elevated c-Myc. Cell. 2012;151(1):56–67. doi: 10.1016/j.cell.2012.08.026. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 109.Meyer N, Penn LZ. Reflecting on 25 years with MYC. Nat Rev Cancer. 2008;8(12):976–90. doi: 10.1038/nrc2231. [DOI] [PubMed] [Google Scholar]
- 110.Ladanyi M, Park CK, Lewis R, Jhanwar SC, Healey JH, Huvos AG. Sporadic amplification of the MYC gene in human osteosarcomas. Diagn Mol Pathol. 1993;2(3):163–7. [PubMed] [Google Scholar]
- 111.Pompetti F, Rizzo P, Simon RM, Freidlin B, Mew DJ, Pass HI, Picci P, Levine AS, Carbone M. Oncogene alterations in primary, recurrent, and metastatic human bone tumors. J Cell Biochem. 1996;63(1):37–50. doi: 10.1002/(SICI)1097-4644(199610)63:1%3C37::AID-JCB3%3E3.0.CO;2-0. [DOI] [PubMed] [Google Scholar]
- 112.Tarkkanen M, Elomaa I, Blomqvist C, Kivioja AH, Kellokumpu-Lehtinen P, Bohling T, Valle J, Knuutila S. DNA sequence copy number increase at 8q: a potential new prognostic marker in high-grade osteosarcoma. Int J Cancer. 1999;84(2):114–21. doi: 10.1002/(sici)1097-0215(19990420)84:2<114::aid-ijc4>3.0.co;2-q. [DOI] [PubMed] [Google Scholar]
- 113.Stock C, Kager L, Fink FM, Gadner H, Ambros PF. Chromosomal regions involved in the pathogenesis of osteosarcomas. Genes Chromosomes Cancer. 2000;28(3):329–36. doi: 10.1002/1098-2264(200007)28:3<329::aid-gcc11>3.0.co;2-f. [DOI] [PubMed] [Google Scholar]
- 114.Squire JA, Pei J, Marrano P, Beheshti B, Bayani J, Lim G, Moldovan L, Zielenska M. High-resolution mapping of amplifications and deletions in pediatric osteosarcoma by use of CGH analysis of cDNA microarrays. Genes Chromosomes Cancer. 2003;38(3):215–25. doi: 10.1002/gcc.10273. [DOI] [PubMed] [Google Scholar]
- 115.Gamberi G, Benassi MS, Bohling T, Ragazzini P, Molendini L, Sollazzo MR, Pompetti F, Merli M, Magagnoli G, Balladelli A, Picci P. C-myc and c-fos in human osteosarcoma: prognostic value of mRNA and protein expression. Oncology. 1998;55(6):556–63. doi: 10.1159/000011912. [DOI] [PubMed] [Google Scholar]
- 116.Sadikovic B, Thorner P, Chilton-Macneill S, Martin JW, Cervigne NK, Squire J, Zielenska M. Expression analysis of genes associated with human osteosarcoma tumors shows correlation of RUNX2 overexpression with poor response to chemotherapy. BMC Cancer. 2010;10:202. doi: 10.1186/1471-2407-10-202. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 117.Yotov WV, Hamel H, Rivard GE, Champagne MA, Russo PA, Leclerc JM, Bernstein ML, Levy E. Amplifications of DNA primase 1 (PRIM1) in human osteosarcoma. Genes Chromosomes Cancer. 1999;26(1):62–9. doi: 10.1002/(sici)1098-2264(199909)26:1<62::aid-gcc9>3.0.co;2-f. [DOI] [PubMed] [Google Scholar]
- 118.Batanian JR, Cavalli LR, Aldosari NM, Ma E, Sotelo-Avila C, Ramos MB, Rone JD, Thorpe CM, Haddad BR. Evaluation of paediatric osteosarcomas by classic cytogenetic and CGH analyses. Mol Pathol. 2002;55(6):389–93. doi: 10.1136/mp.55.6.389. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 119.van Dartel M, Cornelissen PW, Redeker S, Tarkkanen M, Knuutila S, Hogendoorn PC, Westerveld A, Gomes I, Bras J, Hulsebos TJ. Amplification of 17p11.2 approximately p12, including PMP22, TOP3A, and MAPK7, in high-grade osteosarcoma. Cancer Genet Cytogenet. 2002;139(2):91–6. doi: 10.1016/s0165-4608(02)00627-1. [DOI] [PubMed] [Google Scholar]
- 120.Man TK, Lu XY, Jaeweon K, Perlaky L, Harris CP, Shah S, Ladanyi M, Gorlick R, Lau CC, Rao PH. Genome-wide array comparative genomic hybridization analysis reveals distinct amplifications in osteosarcoma. BMC Cancer. 2004;4:45. doi: 10.1186/1471-2407-4-45. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 121.Zielenska M, Marrano P, Thorner P, Pei J, Beheshti B, Ho M, Bayani J, Liu Y, Sun BC, Squire JA, Hao XS. High-resolution cDNA microarray CGH mapping of genomic imbalances in osteosarcoma using formalin-fixed paraffin-embedded tissue. Cytogenet Genome Res. 2004;107(1–2):77–82. doi: 10.1159/000079574. [DOI] [PubMed] [Google Scholar]
- 122.Atiye J, Wolf M, Kaur S, Monni O, Bohling T, Kivioja A, Tas E, Serra M, Tarkkanen M, Knuutila S. Gene amplifications in osteosarcoma-CGH microarray analysis. Genes Chromosomes Cancer. 2005;42(2):158–63. doi: 10.1002/gcc.20120. [DOI] [PubMed] [Google Scholar]
- 123.Lu XY, Lu Y, Zhao YJ, Jaeweon K, Kang J, Xiao-Nan L, Ge G, Meyer R, Perlaky L, Hicks J, Chintagumpala M, Cai WW, Ladanyi M, Gorlick R, Lau CC, Pati D, Sheldon M, Rao PH. Cell cycle regulator gene CDC5L, a potential target for 6p12-p21 amplicon in osteosarcoma. Mol Cancer Res. 2008;6(6):937–46. doi: 10.1158/1541-7786.MCR-07-2115. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 124.Forus A, Weghuis DO, Smeets D, Fodstad O, Myklebost O, Geurts van Kessel A. Comparative genomic hybridization analysis of human sarcomas: II. Identification of novel amplicons at 6p and 17p in osteosarcomas. Genes Chromosomes Cancer. 1995;14(1):15–21. doi: 10.1002/gcc.2870140104. [DOI] [PubMed] [Google Scholar]
- 125.Yang J, Yang D, Sun Y, Sun B, Wang G, Trent JC, Araujo DM, Chen K, Zhang W. Genetic amplification of the vascular endothelial growth factor (VEGF) pathway genes, including VEGFA, in human osteosarcoma. Cancer. 2011;117(21):4925–38. doi: 10.1002/cncr.26116. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 126.Hoeijmakers JH. Genome maintenance mechanisms for preventing cancer. Nature. 2001;411(6835):366–74. doi: 10.1038/35077232. [DOI] [PubMed] [Google Scholar]
- 127.Kops GJ, Weaver BA, Cleveland DW. On the road to cancer: aneuploidy and the mitotic checkpoint. Nat Rev Cancer. 2005;5(10):773–85. doi: 10.1038/nrc1714. [DOI] [PubMed] [Google Scholar]
- 128.Compton DA. Mechanisms of aneuploidy. Curr Opin Cell Biol. 2011;23(1):109–13. doi: 10.1016/j.ceb.2010.08.007. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 129.Dawson MA, Kouzarides T. Cancer epigenetics: from mechanism to therapy. Cell. 2012;150(1):12–27. doi: 10.1016/j.cell.2012.06.013. [DOI] [PubMed] [Google Scholar]
- 130.Jones PA, Baylin SB. The fundamental role of epigenetic events in cancer. Nat Rev Genet. 2002;3(6):415–28. doi: 10.1038/nrg816. [DOI] [PubMed] [Google Scholar]
- 131.Baylin SB, Jones PA. A decade of exploring the cancer epigenome - biological and translational implications. Nat Rev Cancer. 2011;11(10):726–34. doi: 10.1038/nrc3130. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 132.Robertson KD. DNA methylation and human disease. Nat Rev Genet. 2005;6(8):597–610. doi: 10.1038/nrg1655. [DOI] [PubMed] [Google Scholar]
- 133.Sakai T, Toguchida J, Ohtani N, Yandell DW, Rapaport JM, Dryja TP. Allele-specific hypermethylation of the retinoblastoma tumor-suppressor gene. Am J Hum Genet. 1991;48(5):880–8. [PMC free article] [PubMed] [Google Scholar]
- 134.Li B, Ye Z. Epigenetic alterations in osteosarcoma: promising targets. Mol Biol Rep. 2014;41(5):3303–15. doi: 10.1007/s11033-014-3193-7. [DOI] [PubMed] [Google Scholar]
- 135.Park YB, Park MJ, Kimura K, Shimizu K, Lee SH, Yokota J. Alterations in the INK4a/ARF locus and their effects on the growth of human osteosarcoma cell lines. Cancer Genet Cytogenet. 2002;133(2):105–11. doi: 10.1016/s0165-4608(01)00575-1. [DOI] [PubMed] [Google Scholar]
- 136.Oh JH, Kim HS, Kim HH, Kim WH, Lee SH. Aberrant methylation of p14ARF gene correlates with poor survival in osteosarcoma. Clin Orthop Relat Res. 2006;442:216–22. doi: 10.1097/01.blo.0000188063.56091.69. [DOI] [PubMed] [Google Scholar]
- 137.Moolmuang B, Tainsky MA. CREG1 enhances p16(INK4a) -induced cellular senescence. Cell Cycle. 2011;10(3):518–30. doi: 10.4161/cc.10.3.14756. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 138.Badal V, Menendez S, Coomber D, Lane DP. Regulation of the p14ARF promoter by DNA methylation. Cell Cycle. 2008;7(1):112–9. doi: 10.4161/cc.7.1.5137. [DOI] [PubMed] [Google Scholar]
- 139.Al-Romaih K, Somers GR, Bayani J, Hughes S, Prasad M, Cutz JC, Xue H, Zielenska M, Wang Y, Squire JA. Modulation by decitabine of gene expression and growth of osteosarcoma U2OS cells in vitro and in xenografts: identification of apoptotic genes as targets for demethylation. Cancer Cell Int. 2007;7:14. doi: 10.1186/1475-2867-7-14. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 140.Le May N, Mota-Fernandes D, Velez-Cruz R, Iltis I, Biard D, Egly JM. NER factors are recruited to active promoters and facilitate chromatin modification for transcription in the absence of exogenous genotoxic attack. Mol Cell. 2010;38(1):54–66. doi: 10.1016/j.molcel.2010.03.004. [DOI] [PubMed] [Google Scholar]
- 141.Sen GL, Reuter JA, Webster DE, Zhu L, Khavari PA. DNMT1 maintains progenitor function in self-renewing somatic tissue. Nature. 2010;463(7280):563–7. doi: 10.1038/nature08683. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 142.Guo JU, Ma DK, Mo H, Ball MP, Jang MH, Bonaguidi MA, Balazer JA, Eaves HL, Xie B, Ford E, Zhang K, Ming GL, Gao Y, Song H. Neuronal activity modifies the DNA methylation landscape in the adult brain. Nat Neurosci. 2011;14(10):1345–51. doi: 10.1038/nn.2900. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 143.Chen WY, Wang DH, Yen RC, Luo J, Gu W, Baylin SB. Tumor suppressor HIC1 directly regulates SIRT1 to modulate p53-dependent DNA-damage responses. Cell. 2005;123(3):437–48. doi: 10.1016/j.cell.2005.08.011. [DOI] [PubMed] [Google Scholar]
- 144.Rathi A, Virmani AK, Harada K, Timmons CF, Miyajima K, Hay RJ, Mastrangelo D, Maitra A, Tomlinson GE, Gazdar AF. Aberrant methylation of the HIC1 promoter is a frequent event in specific pediatric neoplasms. Clin Cancer Res. 2003;9(10):3674–8. Pt 1. [PubMed] [Google Scholar]
- 145.Chen W, Cooper TK, Zahnow CA, Overholtzer M, Zhao Z, Ladanyi M, Karp JE, Gokgoz N, Wunder JS, Andrulis IL, Levine AJ, Mankowski JL, Baylin SB. Epigenetic and genetic loss of Hic1 function accentuates the role of p53 in tumorigenesis. Cancer Cell. 2004;6(4):387–98. doi: 10.1016/j.ccr.2004.08.030. [DOI] [PubMed] [Google Scholar]
- 146.Harada K, Toyooka S, Maitra A, Maruyama R, Toyooka KO, Timmons CF, Tomlinson GE, Mastrangelo D, Hay RJ, Minna JD, Gazdar AF. Aberrant promoter methylation and silencing of the RASSF1A gene in pediatric tumors and cell lines. Oncogene. 2002;21(27):4345–9. doi: 10.1038/sj.onc.1205446. [DOI] [PubMed] [Google Scholar]
- 147.Lim S, Yang MH, Park JH, Nojima T, Hashimoto H, Unni KK, Park YK. Inactivation of the RASSF1A in osteosarcoma. Oncol Rep. 2003;10(4):897–901. [PubMed] [Google Scholar]
- 148.Hou P, Ji M, Yang B, Chen Z, Qiu J, Shi X, Lu Z. Quantitative analysis of promoter hypermethylation in multiple genes in osteosarcoma. Cancer. 2006;106(7):1602–9. doi: 10.1002/cncr.21762. [DOI] [PubMed] [Google Scholar]
- 149.Kansara M, Tsang M, Kodjabachian L, Sims NA, Trivett MK, Ehrich M, Dobrovic A, Slavin J, Choong PF, Simmons PJ, Dawid IB, Thomas DM. Wnt inhibitory factor 1 is epigenetically silenced in human osteosarcoma, and targeted disruption accelerates osteosarcomagenesis in mice. J Clin Invest. 2009;119(4):837–51. doi: 10.1172/JCI37175. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 150.Rubin EM, Guo Y, Tu K, Xie J, Zi X, Hoang BH. Wnt inhibitory factor 1 decreases tumorigenesis and metastasis in osteosarcoma. Mol Cancer Ther. 2010;9(3):731–41. doi: 10.1158/1535-7163.MCT-09-0147. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 151.Donninger H, Vos MD, Clark GJ. The RASSF1A tumor suppressor. J Cell Sci. 2007;120:3163–72. doi: 10.1242/jcs.010389. Pt 18. [DOI] [PubMed] [Google Scholar]
- 152.Yang Q, Zage P, Kagan D, Tian Y, Seshadri R, Salwen HR, Liu S, Chlenski A, Cohn SL. Association of epigenetic inactivation of RASSF1A with poor outcome in human neuroblastoma. Clin Cancer Res. 2004;10(24):8493–500. doi: 10.1158/1078-0432.CCR-04-1331. [DOI] [PubMed] [Google Scholar]
- 153.Patra SK, Szyf M. DNA methylation-mediated nucleosome dynamics and oncogenic Ras signaling: insights from FAS, FAS ligand and RASSF1A. FEBS J. 2008;275(21):5217–35. doi: 10.1111/j.1742-4658.2008.06658.x. [DOI] [PubMed] [Google Scholar]
- 154.Thaler R, Agsten M, Spitzer S, Paschalis EP, Karlic H, Klaushofer K, Varga F. Homocysteine suppresses the expression of the collagen cross-linker lysyl oxidase involving IL-6, Fli1, and epigenetic DNA methylation. J Biol Chem. 2011;286(7):5578–88. doi: 10.1074/jbc.M110.166181. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 155.Worth LL, Lafleur EA, Jia SF, Kleinerman ES. Fas expression inversely correlates with metastatic potential in osteosarcoma cells. Oncol Rep. 2002;9(4):823–7. [PubMed] [Google Scholar]
- 156.Jia SF, Worth LL, Densmore CL, Xu B, Zhou Z, Kleinerman ES. Eradication of osteosarcoma lung metastases following intranasal interleukin-12 gene therapy using a nonviral polyethylenimine vector. Cancer Gene Ther. 2002;9(3):260–6. doi: 10.1038/sj.cgt.7700432. [DOI] [PubMed] [Google Scholar]
- 157.Gordon N, Kleinerman ES. Aerosol therapy for the treatment of osteosarcoma lung metastases: targeting the Fas/FasL pathway and rationale for the use of gemcitabine. J Aerosol Med Pulm Drug Deliv. 2010;23(4):189–96. doi: 10.1089/jamp.2009.0812. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 158.Gordon N, Koshkina NV, Jia SF, Khanna C, Mendoza A, Worth LL, Kleinerman ES. Corruption of the Fas pathway delays the pulmonary clearance of murine osteosarcoma cells, enhances their metastatic potential, and reduces the effect of aerosol gemcitabine. Clin Cancer Res. 2007;13(15):4503–10. doi: 10.1158/1078-0432.CCR-07-0313. Pt 1. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 159.Jia SF, Worth LL, Densmore CL, Xu B, Duan X, Kleinerman ES. Aerosol gene therapy with PEI: IL-12 eradicates osteosarcoma lung metastases. Clin Cancer Res. 2003;9(9):3462–8. [PubMed] [Google Scholar]
- 160.Thaler R, Spitzer S, Karlic H, Berger C, Klaushofer K, Varga F. Ibandronate increases the expression of the proapoptotic gene FAS by epigenetic mechanisms in tumor cells. Biochem Pharmacol. 2013;85(2):173–85. doi: 10.1016/j.bcp.2012.10.016. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 161.Bernstein BE, Birney E, Dunham I, Green ED, Gunter C, Snyder M. An integrated encyclopedia of DNA elements in the human genome. Nature. 2012;489(7414):57–74. doi: 10.1038/nature11247. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 162.Jakovcevski M, Akbarian S. Epigenetic mechanisms in neurological disease. Nat Med. 2012;18(8):1194–204. doi: 10.1038/nm.2828. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 163.Hon GC, Hawkins RD, Ren B. Predictive chromatin signatures in the mammalian genome. Hum Mol Genet. 2009;18(R2):R195–201. doi: 10.1093/hmg/ddp409. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 164.Zentner GE, Tesar PJ, Scacheri PC. Epigenetic signatures distinguish multiple classes of enhancers with distinct cellular functions. Genome Res. 2011;21(8):1273–83. doi: 10.1101/gr.122382.111. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 165.Cieslik M, Bekiranov S. Combinatorial epigenetic patterns as quantitative predictors of chromatin biology. BMC Genomics. 2014;15:76. doi: 10.1186/1471-2164-15-76. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 166.Rodriguez-Paredes M, Esteller M. Cancer epigenetics reaches mainstream oncology. Nat Med. 2011;17(3):330–9. doi: 10.1038/nm.2305. [DOI] [PubMed] [Google Scholar]
- 167.Levine M. Transcriptional enhancers in animal development and evolution. Curr Biol. 2010;20(17):R754–63. doi: 10.1016/j.cub.2010.06.070. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 168.Cantone I, Fisher AG. Epigenetic programming and reprogramming during development. Nat Struct Mol Biol. 2013;20(3):282–9. doi: 10.1038/nsmb.2489. [DOI] [PubMed] [Google Scholar]
- 169.Talluri S, Dick FA. Regulation of transcription and chromatin structure by pRB: here, there and everywhere. Cell Cycle. 2012;11(17):3189–98. doi: 10.4161/cc.21263. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 170.Narita M, Nunez S, Heard E, Lin AW, Hearn SA, Spector DL, Hannon GJ, Lowe SW. Rb-mediated heterochromatin formation and silencing of E2F target genes during cellular senescence. Cell. 2003;113(6):703–16. doi: 10.1016/s0092-8674(03)00401-x. [DOI] [PubMed] [Google Scholar]
- 171.Calo E, Quintero-Estades JA, Danielian PS, Nedelcu S, Berman SD, Lees JA. Rb regulates fate choice and lineage commitment in vivo. Nature. 2010;466(7310):1110–4. doi: 10.1038/nature09264. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 172.Bennani-Baiti IM, Machado I, Llombart-Bosch A, Kovar H. Lysine-specific demethylase 1 (LSD1/KDM1A/AOF2/BHC110) is expressed and is an epigenetic drug target in chondrosarcoma, Ewing’s sarcoma, osteosarcoma, and rhabdomyosarcoma. Hum Pathol. 2012;43(8):1300–7. doi: 10.1016/j.humpath.2011.10.010. [DOI] [PubMed] [Google Scholar]
- 173.Shi Y. Histone lysine demethylases: emerging roles in development, physiology and disease. Nat Rev Genet. 2007;8(11):829–33. doi: 10.1038/nrg2218. [DOI] [PubMed] [Google Scholar]
- 174.Tsai WW, Nguyen TT, Shi Y, Barton MC. p53-targeted LSD1 functions in repression of chromatin structure and transcription in vivo. Mol Cell Biol. 2008;28(17):5139–46. doi: 10.1128/MCB.00287-08. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 175.Easwaran H, Johnstone SE, Van Neste L, Ohm J, Mosbruger T, Wang Q, Aryee MJ, Joyce P, Ahuja N, Weisenberger D, Collisson E, Zhu J, Yegnasubramanian S, Matsui W, Baylin SB. A DNA hypermethylation module for the stem/progenitor cell signature of cancer. Genome Res. 2012;22(5):837–49. doi: 10.1101/gr.131169.111. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 176.Weinberg RA, Penman S. Small molecular weight monodisperse nuclear RNA. J Mol Biol. 1968;38(3):289–304. doi: 10.1016/0022-2836(68)90387-2. [DOI] [PubMed] [Google Scholar]
- 177.Paul J, Duerksen JD. Chromatin-associated RNA content of heterochromatin and euchromatin. Mol Cell Biochem. 1975;9(1):9–16. doi: 10.1007/BF01731728. [DOI] [PubMed] [Google Scholar]
- 178.Salditt-Georgieff M, Harpold MM, Wilson MC, Darnell JE., Jr. Large heterogeneous nuclear ribonucleic acid has three times as many 5–caps as polyadenylic acid segments, and most caps do not enter polyribosomes. Mol Cell Biol. 1981;1(2):179–87. doi: 10.1128/mcb.1.2.179. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 179.Salditt-Georgieff M, Darnell JE., Jr. Further evidence that the majority of primary nuclear RNA transcripts in mammalian cells do not contribute to mRNA. Mol Cell Biol. 1982;2(6):701–7. doi: 10.1128/mcb.2.6.701. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 180.Nickerson JA, Krochmalnic G, Wan KM, Penman S. Chromatin architecture and nuclear RNA. Proc Natl Acad Sci U S A. 1989;86(1):177–81. doi: 10.1073/pnas.86.1.177. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 181.Djebali S, Davis CA, Merkel A, Dobin A, Lassmann T, Mortazavi A, Tanzer A, Lagarde J, Lin W, Schlesinger F, Xue C, Marinov GK, Khatun J, Williams BA, Zaleski C, Rozowsky J, Roder M, Kokocinski F, Abdelhamid RF, Alioto T, Antoshechkin I, Baer MT, Bar NS, Batut P, Bell K, Bell I, Chakrabortty S, Chen X, Chrast J, Curado J, Derrien T, Drenkow J, Dumais E, Dumais J, Duttagupta R, Falconnet E, Fastuca M, Fejes-Toth K, Ferreira P, Foissac S, Fullwood MJ, Gao H, Gonzalez D, Gordon A, Gunawardena H, Howald C, Jha S, Johnson R, Kapranov P, King B, Kingswood C, Luo OJ, Park E, Persaud K, Preall JB, Ribeca P, Risk B, Robyr D, Sammeth M, Schaffer L, See LH, Shahab A, Skancke J, Suzuki AM, Takahashi H, Tilgner H, Trout D, Walters N, Wang H, Wrobel J, Yu Y, Ruan X, Hayashizaki Y, Harrow J, Gerstein M, Hubbard T, Reymond A, Antonarakis SE, Hannon G, Giddings MC, Ruan Y, Wold B, Carninci P, Guigo R, Gingeras TR. Landscape of transcription in human cells. Nature. 2012;489(7414):101–8. doi: 10.1038/nature11233. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 182.Bergmann JH, Spector DL. Long non-coding RNAs: modulators of nuclear structure and function. Curr Opin Cell Biol. 2014;26:10–18. doi: 10.1016/j.ceb.2013.08.005. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 183.Lander ES, Linton LM, Birren B, Nusbaum C, Zody MC, Baldwin J, Devon K, Dewar K, Doyle M, FitzHugh W, Funke R, Gage D, Harris K, Heaford A, Howland J, Kann L, Lehoczky J, LeVine R, McEwan P, McKernan K, Meldrim J, Mesirov JP, Miranda C, Morris W, Naylor J, Raymond C, Rosetti M, Santos R, Sheridan A, Sougnez C, Stange-Thomann N, Stojanovic N, Subramanian A, Wyman D, Rogers J, Sulston J, Ainscough R, Beck S, Bentley D, Burton J, Clee C, Carter N, Coulson A, Deadman R, Deloukas P, Dunham A, Dunham I, Durbin R, French L, Grafham D, Gregory S, Hubbard T, Humphray S, Hunt A, Jones M, Lloyd C, McMurray A, Matthews L, Mercer S, Milne S, Mullikin JC, Mungall A, Plumb R, Ross M, Shownkeen R, Sims S, Waterston RH, Wilson RK, Hillier LW, McPherson JD, Marra MA, Mardis ER, Fulton LA, Chinwalla AT, Pepin KH, Gish WR, Chissoe SL, Wendl MC, Delehaunty KD, Miner TL, Delehaunty A, Kramer JB, Cook LL, Fulton RS, Johnson DL, Minx PJ, Clifton SW, Hawkins T, Branscomb E, Predki P, Richardson P, Wenning S, Slezak T, Doggett N, Cheng JF, Olsen A, Lucas S, Elkin C, Uberbacher E, Frazier M, Gibbs RA, Muzny DM, Scherer SE, Bouck JB, Sodergren EJ, Worley KC, Rives CM, Gorrell JH, Metzker ML, Naylor SL, Kucherlapati RS, Nelson DL, Weinstock GM, Sakaki Y, Fujiyama A, Hattori M, Yada T, Toyoda A, Itoh T, Kawagoe C, Watanabe H, Totoki Y, Taylor T, Weissenbach J, Heilig R, Saurin W, Artiguenave F, Brottier P, Bruls T, Pelletier E, Robert C, Wincker P, Smith DR, Doucette-Stamm L, Rubenfield M, Weinstock K, Lee HM, Dubois J, Rosenthal A, Platzer M, Nyakatura G, Taudien S, Rump A, Yang H, Yu J, Wang J, Huang G, Gu J, Hood L, Rowen L, Madan A, Qin S, Davis RW, Federspiel NA, Abola AP, Proctor MJ, Myers RM, Schmutz J, Dickson M, Grimwood J, Cox DR, Olson MV, Kaul R, Shimizu N, Kawasaki K, Minoshima S, Evans GA, Athanasiou M, Schultz R, Roe BA, Chen F, Pan H, Ramser J, Lehrach H, Reinhardt R, McCombie WR, de la Bastide M, Dedhia N, Blocker H, Hornischer K, Nordsiek G, Agarwala R, Aravind L, Bailey JA, Bateman A, Batzoglou S, Birney E, Bork P, Brown DG, Burge CB, Cerutti L, Chen HC, Church D, Clamp M, Copley RR, Doerks T, Eddy SR, Eichler EE, Furey TS, Galagan J, Gilbert JG, Harmon C, Hayashizaki Y, Haussler D, Hermjakob H, Hokamp K, Jang W, Johnson LS, Jones TA, Kasif S, Kaspryzk A, Kennedy S, Kent WJ, Kitts P, Koonin EV, Korf I, Kulp D, Lancet D, Lowe TM, McLysaght A, Mikkelsen T, Moran JV, Mulder N, Pollara VJ, Ponting CP, Schuler G, Schultz J, Slater G, Smit AF, Stupka E, Szustakowski J, Thierry-Mieg D, Thierry-Mieg J, Wagner L, Wallis J, Wheeler R, Williams A, Wolf YI, Wolfe KH, Yang SP, Yeh RF, Collins F, Guyer MS, Peterson J, Felsenfeld A, Wetterstrand KA, Patrinos A, Morgan MJ, de Jong P, Catanese JJ, Osoegawa K, Shizuya H, Choi S, Chen YJ. Initial sequencing and analysis of the human genome. Nature. 2001;409(6822):860–921. doi: 10.1038/35057062. [DOI] [PubMed] [Google Scholar]
- 184.Rinn JL, Chang HY. Genome regulation by long noncoding RNAs. Annu Rev Biochem. 2012;81:145–66. doi: 10.1146/annurev-biochem-051410-092902. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 185.Ameres SL, Zamore PD. Diversifying microRNA sequence and function. Nat Rev Mol Cell Biol. 2013;14(8):475–88. doi: 10.1038/nrm3611. [DOI] [PubMed] [Google Scholar]
- 186.Cheetham SW, Gruhl F, Mattick JS, Dinger ME. Long noncoding RNAs and the genetics of cancer. Br J Cancer. 2013;108(12):2419–25. doi: 10.1038/bjc.2013.233. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 187.Lu J, Getz G, Miska EA, Alvarez-Saavedra E, Lamb J, Peck D, Sweet-Cordero A, Ebert BL, Mak RH, Ferrando AA, Downing JR, Jacks T, Horvitz HR, Golub TR. MicroRNA expression profiles classify human cancers. Nature. 2005;435(7043):834–8. doi: 10.1038/nature03702. [DOI] [PubMed] [Google Scholar]
- 188.Liu Q, Huang J, Zhou N, Zhang Z, Zhang A, Zhaohui L, Wu F, Mo Y-Y. LncRNA loc285194 is a p53-regulated tumor suppressor. Nucleic Acids Res. 2013;41(9):4976–87. doi: 10.1093/nar/gkt182. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 189.Li JP1, Liu LH, Li J, Chen Y, Jiang XW, Ouyang YR, Liu YQ, Zhong H, Li H, Xiao T. Microarray expression profile of long noncoding RNAs in human osteosarcoma. Biochem Biophys Res Commun. 2013;433(2):200–6. doi: 10.1016/j.bbrc.2013.02.083. [DOI] [PubMed] [Google Scholar]
- 190.Kelly TK, De Carvalho DD, Jones PA. Epigenetic modifications as therapeutic targets. Nat Biotechnol. 2010;28(10):1069–78. doi: 10.1038/nbt.1678. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 191.Kobayashi E, Hornicek FJ, Duan Z. MicroRNA involvement in osteosarcoma. Sarcoma. 2012;2012:359739. doi: 10.1155/2012/359739. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 192.Maire G, Martin JW, Yoshimoto M, Chilton-MacNeill S, Zielenska M, Squire JA. Analysis of miRNA-gene expression-genomic profiles reveals complex mechanisms of microRNA deregulation in osteosarcoma. Cancer Genet. 2011;204(3):138–46. doi: 10.1016/j.cancergen.2010.12.012. [DOI] [PubMed] [Google Scholar]
- 193.Jones KB, Salah Z, Del Mare S, Galasso M, Gaudio E, Nuovo GJ, Lovat F, LeBlanc K, Palatini J, Randall RL, Volinia S, Stein GS, Croce CM, Lian JB, Aqeilan RI. miRNA signatures associate with pathogenesis and progression of osteosarcoma. Cancer Res. 2012;72(7):1865–77. doi: 10.1158/0008-5472.CAN-11-2663. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 194.Poos K, Smida J, Nathrath M, Maugg D, Baumhoer D, Korsching E. How microRNA and transcription factor co-regulatory networks affect osteosarcoma cell proliferation. PLoS Comput Biol. 2013;9(8):e1003210. doi: 10.1371/journal.pcbi.1003210. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 195.Namlos HM, Meza-Zepeda LA, Baroy T, Ostensen IH, Kresse SH, Kuijjer ML, Serra M, Burger H, Cleton-Jansen AM, Myklebost O. Modulation of the osteosarcoma expression phenotype by microRNAs. PLoS One. 2012;7(10):e48086. doi: 10.1371/journal.pone.0048086. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 196.He C, Xiong J, Xu X, Lu W, Liu L, Xiao D, Wang D. Functional elucidation of MiR-34 in osteosarcoma cells and primary tumor samples. Biochem Biophys Res Commun. 2009;388(1):35–40. doi: 10.1016/j.bbrc.2009.07.101. [DOI] [PubMed] [Google Scholar]
- 197.Ji F, Zhang H, Wang Y, Li M, Xu W, Kang Y, Wang Z, Cheng P, Tong D, Li C, Tang H. MicroRNA-133a, down-regulated in osteosarcoma, suppresses proliferation and promotes apoptosis by targeting Bcl-xL and Mcl-1. Bone. 2013;56(1):220–6. doi: 10.1016/j.bone.2013.05.020. [DOI] [PubMed] [Google Scholar]
- 198.Duan Z, Choy E, Harmon D, Liu X, Susa M, Mankin H, Hornicek F. MicroRNA-199a-3p is downregulated in human osteosarcoma and regulates cell proliferation and migration. Mol Cancer Ther. 2011;10(8):1337–45. doi: 10.1158/1535-7163.MCT-11-0096. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 199.Wang Y, Zhao W, Fu Q. miR-335 suppresses migration and invasion by targeting ROCK1 in osteosarcoma cells. Mol Cell Biochem. 2013;384(1–2):105–11. doi: 10.1007/s11010-013-1786-4. [DOI] [PubMed] [Google Scholar]
- 200.Zhou X, Wei M, Wang W. MicroRNA-340 suppresses osteosarcoma tumor growth and metastasis by directly targeting ROCK1. Biochem Biophys Res Commun. 2013;437(4):653–8. doi: 10.1016/j.bbrc.2013.07.033. [DOI] [PubMed] [Google Scholar]
- 201.Png KJ, Yoshida M, Zhang XH, Shu W, Lee H, Rimner A, Chan TA, Comen E, Andrade VP, Kim SW, King TA, Hudis CA, Norton L, Hicks J, Massague J, Tavazoie SF. MicroRNA-335 inhibits tumor reinitiation and is silenced through genetic and epigenetic mechanisms in human breast cancer. Genes Dev. 2011;25(3):226–31. doi: 10.1101/gad.1974211. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 202.Das S, Bryan K, Buckley PG, Piskareva O, Bray IM, Foley N, Ryan J, Lynch J, Creevey L, Fay J, Prenter S, Koster J, van Sluis P, Versteeg R, Eggert A, Schulte JH, Schramm A, Mestdagh P, Vandesompele J, Speleman F, Stallings RL. Modulation of neuroblastoma disease pathogenesis by an extensive network of epigenetically regulated microRNAs. Oncogene. 2013;32(24):2927–36. doi: 10.1038/onc.2012.311. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 203.Huang G, Nishimoto K, Zhou Z, Hughes D, Kleinerman ES. miR-20a encoded by the miR-17-92 cluster increases the metastatic potential of osteosarcoma cells by regulating Fas expression. Cancer Res. 2012;72(4):908–16. doi: 10.1158/0008-5472.CAN-11-1460. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 204.Kovalchuk O, Zemp FJ, Filkowski JN, Altamirano AM, Dickey JS, Jenkins-Baker G, Marino SA, Brenner DJ, Bonner WM, Sedelnikova OA. microRNAome changes in bystander three-dimensional human tissue models suggest priming of apoptotic pathways. Carcinogenesis. 2010;31(10):1882–8. doi: 10.1093/carcin/bgq119. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 205.Song B, Wang Y, Xi Y, Kudo K, Bruheim S, Botchkina GI, Gavin E, Wan Y, Formentini A, Kornmann M, Fodstad O, Ju J. Mechanism of chemoresistance mediated by miR-140 in human osteosarcoma and colon cancer cells. Oncogene. 2009;28(46):4065–74. doi: 10.1038/onc.2009.274. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 206.Khanna C, Fan TM, Gorlick R, Helman LJ, Kleinerman ES, Adamson PC, Houghton PJ, Tap WD, Welch DR, Steeg PS, Merlino G, Sorensen PH, Meltzer P, Kirsch DG, Janeway KA, Weigel B, Randall L, Withrow SJ, Paoloni M, Kaplan R, Teicher BA, Seibel NL, Uren A, Patel SR, Trent J, Savage SA, Mirabello L, Reinke D, Barkauskas DA, Krailo M, Bernstein M. Towards a drug development path that targets metastatic progression in osteosarcoma. Clin Cancer Res. 2014;20(16):4200–9. doi: 10.1158/1078-0432.CCR-13-2574. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 207.Koshkina NV, Kleinerman ES. Aerosol gemcitabine inhibits the growth of primary osteosarcoma and osteosarcoma lung metastases. Int J Cancer. 2005;116(3):458–63. doi: 10.1002/ijc.21011. [DOI] [PubMed] [Google Scholar]
- 208.Ding Q, Zhang Z, Liu JJ, Jiang N, Zhang J, Ross TM, Chu XJ, Bartkovitz D, Podlaski F, Janson C, Tovar C, Filipovic ZM, Higgins B, Glenn K, Packman K, Vassilev LT, Graves B. Discovery of RG7388, a potent and selective p53-MDM2 inhibitor in clinical development. J Med Chem. 2013;56(14):5979–83. doi: 10.1021/jm400487c. [DOI] [PubMed] [Google Scholar]
- 209.Sampson ER, Martin BA, Morris AE, Xie C, Schwarz EM, O’Keefe RJ, Rosier RN. The orally bioavailable met inhibitor PF-2341066 inhibits osteosarcoma growth and osteolysis/matrix production in a xenograft model. J Bone Miner Res. 2011;26(6):1283–94. doi: 10.1002/jbmr.336. [DOI] [PubMed] [Google Scholar]
- 210.Khanna C, Wan X, Bose S, Cassaday R, Olomu O, Mendoza A, Yeung C, Gorlick R, Hewitt SM, Helman LJ. The membrane-cytoskeleton linker ezrin is necessary for osteosarcoma metastasis. Nat Med. 2004;10(2):182–6. doi: 10.1038/nm982. [DOI] [PubMed] [Google Scholar]
- 211.Bulut G, Hong SH, Chen K, Beauchamp EM, Rahim S, Kosturko GW, Glasgow E, Dakshanamurthy S, Lee HS, Daar I, Toretsky JA, Khanna C, Uren A. Small molecule inhibitors of ezrin inhibit the invasive phenotype of osteosarcoma cells. Oncogene. 2012;31(3):269–81. doi: 10.1038/onc.2011.245. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 212.Rudin CM. Vismodegib. Clin Cancer Res. 2012;18(12):3218–22. doi: 10.1158/1078-0432.CCR-12-0568. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 213.Lo WW, Wunder JS, Dickson BC, Campbell V, McGovern K, Alman BA, Andrulis IL. Involvement and targeted intervention of dysregulated Hedgehog signaling in osteosarcoma. Cancer. 2014;120(4):537–47. doi: 10.1002/cncr.28439. [DOI] [PubMed] [Google Scholar]
- 214.Dubey AK, Dubey S, Handu SS, Qazi MA. Vismodegib: the first drug approved for advanced and metastatic basal cell carcinoma. J Postgrad Med. 2013;59(1):48–50. doi: 10.4103/0022-3859.109494. [DOI] [PubMed] [Google Scholar]
- 215.Hingorani P, Zhang W, Gorlick R, Kolb EA. Inhibition of Src phosphorylation alters meta-static potential of osteosarcoma in vitro but not in vivo. Clin Cancer Res. 2009;15(10):3416–22. doi: 10.1158/1078-0432.CCR-08-1657. [DOI] [PubMed] [Google Scholar]
- 216.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(4):1239–45. doi: 10.1007/s12032-009-9365-y. [DOI] [PubMed] [Google Scholar]
- 217.Wan X, Mendoza A, Khanna C, Helman LJ. Rapamycin inhibits ezrin-mediated metastatic behavior in a murine model of osteosarcoma. Cancer Res. 2005;65(6):2406–11. doi: 10.1158/0008-5472.CAN-04-3135. [DOI] [PubMed] [Google Scholar]
- 218.Roth M, Linkowski M, Tarim J, Piperdi S, Sowers R, Geller D, Gill J, Gorlick R. Ganglioside GD2 as a therapeutic target for antibody-mediated therapy in patients with osteosarcoma. Cancer. 2014;120(4):548–54. doi: 10.1002/cncr.28461. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 219.Gorlick R, Huvos AG, Heller G, Aledo A, Beardsley GP, Healey JH, Meyers PA. Expression of HER2/erbB-2 correlates with survival in osteosarcoma. J Clin Oncol. 1999;17(9):2781–8. doi: 10.1200/JCO.1999.17.9.2781. [DOI] [PubMed] [Google Scholar]
- 220.Zhou H, Randall RL, Brothman AR, Maxwell T, Coffin CM, Goldsby RE. Her-2/neu expression in osteosarcoma increases risk of lung metastasis and can be associated with gene amplification. J Pediatr Hematol Oncol. 2003;25(1):27–32. doi: 10.1097/00043426-200301000-00007. [DOI] [PubMed] [Google Scholar]
- 221.Rainusso N, Brawley VS, Ghazi A, Hicks MJ, Gottschalk S, Rosen JM, Ahmed N. Immunotherapy targeting HER2 with genetically modified T cells eliminates tumor-initiating cells in osteosarcoma. Cancer Gene Ther. 2012;19(3):212–7. doi: 10.1038/cgt.2011.83. [DOI] [PubMed] [Google Scholar]
- 222.Kolb EA, Gorlick R, Billups CA, Hawthorne T, Kurmasheva RT, Houghton PJ, Smith MA. Initial testing (stage 1) of glembatumumab vedotin (CDX-011) by the pediatric preclinical testing program. Pediatr Blood Cancer. 2014;61(10):1816–21. doi: 10.1002/pbc.25099. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 223.Appleton K, Mackay HJ, Judson I, Plumb JA, McCormick C, Strathdee G, Lee C, Barrett S, Reade S, Jadayel D, Tang A, Bellenger K, Mackay L, Setanoians A, Schatzlein A, Twelves C, Kaye SB, Brown R. Phase I and pharmacodynamic trial of the DNA methyltransferase inhibitor decitabine and carboplatin in solid tumors. J Clin Oncol. 2007;25(29):4603–9. doi: 10.1200/JCO.2007.10.8688. [DOI] [PubMed] [Google Scholar]
- 224.Ory B, Heymann MF, Kamijo A, Gouin F, Heymann D, Redini F. Zoledronic acid suppresses lung metastases and prolongs overall survival of osteosarcoma-bearing mice. Cancer. 2005;104(11):2522–9. doi: 10.1002/cncr.21530. [DOI] [PubMed] [Google Scholar]
- 225.Koto K, Horie N, Kimura S, Murata H, Sakabe T, Matsui T, Watanabe M, Adachi S, Maekawa T, Fushiki S, Kubo T. Clinically relevant dose of zoledronic acid inhibits spontaneous lung metastasis in a murine osteosarcoma model. Cancer Lett. 2009;274(2):271–8. doi: 10.1016/j.canlet.2008.09.026. [DOI] [PubMed] [Google Scholar]
- 226.Goldsby RE, Fan TM, Villaluna D, Wagner LM, Isakoff MS, Meyer J, Randall RL, Lee S, Kim G, Bernstein M, Gorlick R, Krailo M, Marina N. Feasibility and dose discovery analysis of zoledronic acid with concurrent chemotherapy in the treatment of newly diagnosed metastatic osteosarcoma: a report from the Children’s Oncology Group. Eur J Cancer. 2013;49(10):2384–91. doi: 10.1016/j.ejca.2013.03.018. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 227.Zorzi AP, Bernstein M, Samson Y, Wall DA, Desai S, Nicksy D, Wainman N, Eisenhauer E, Baruchel S. A phase I study of histone deacetylase inhibitor, pracinostat (SB939), in pediatric patients with refractory solid tumors: IND203 a trial of the NCIC IND program/C17 pediatric phase I consortium. Pediatr Blood Cancer. 2013;60(11):1868–74. doi: 10.1002/pbc.24694. [DOI] [PubMed] [Google Scholar]
- 228.Rao-Bindal K, Zhou Z, Kleinerman ES. MS-275 sensitizes osteosarcoma cells to Fas ligand-induced cell death by increasing the localization of Fas in membrane lipid rafts. Cell Death Dis. 2012;3:e369. doi: 10.1038/cddis.2012.101. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 229.Wittenburg LA, Bisson L, Rose BJ, Korch C, Thamm DH. The histone deacetylase inhibitor valproic acid sensitizes human and canine osteosarcoma to doxorubicin. Cancer Chemother Pharmacol. 2011;67(1):83–92. doi: 10.1007/s00280-010-1287-z. [DOI] [PMC free article] [PubMed] [Google Scholar]

