Advertisement
Regular article|Articles in Press

Transcriptional Profiling Supports the Notochordal Origin of Chordoma and Its Dependence on a TGFΒ1-TBXT Network

Open AccessPublished:February 16, 2023DOI:https://doi.org/10.1016/j.ajpath.2023.01.014
      Chordoma is a rare malignant tumor demonstrating notochordal differentiation. It is dependent on brachyury (TBXT), a hallmark notochordal gene and transcription factor, and shares histologic features and the same anatomic location as the notochord. In this study, we perform a molecular comparison of chordoma and notochord to identify dysregulated cellular pathways. The lack of a molecular reference from appropriate control tissue limits our understanding of chordoma and its relationship to notochord. Accordingly, we conducted an unbiased comparison of chordoma, human notochord, and an atlas of normal and cancerous tissue using gene expression profiling to clarify the chordoma/notochord relationship and potentially identify novel drug targets. We found striking consistency in gene expression profiles between chordoma and notochord, supporting the hypothesis that chordoma develops from notochordal remnants. We identified a 12-gene diagnostic chordoma signature and found that the TBXT/transforming growth factor (TGF)-β/SOX6/SOX9 pathway is hyperactivated in the tumor, suggesting that pathways associated with chondrogenesis are a central driver of chordoma development. Experimental validation in chordoma cells confirms these findings and emphasizes the dependence of chordoma proliferation and survival on TGF-β. Our computational and experimental evidence provides the first molecular connection between notochord and chordoma and identifies core members of a chordoma regulatory pathway involving TBXT. This pathway provides new therapeutic targets for this unique malignant neoplasm and highlights TGF-β as a prime druggable candidate.
      Chordoma is an indolent malignant tumor of the axial skeleton. It is uncommon and accounts for <4% of all primary bone tumors.
      • Scott D.
      • Pedlow F.
      • Hecht A.
      • Hornicek F.
      Chapter 11; tumors: primary benign and malignant extradural spine tumors.
      The yearly incidence is 0.5 cases per million citizens in the United States.
      • Scott D.
      • Pedlow F.
      • Hecht A.
      • Hornicek F.
      Chapter 11; tumors: primary benign and malignant extradural spine tumors.
      It usually arises in adults aged >50 years, although occasional pediatric cases are observed.
      • Scott D.
      • Pedlow F.
      • Hecht A.
      • Hornicek F.
      Chapter 11; tumors: primary benign and malignant extradural spine tumors.
      Chordoma typically develops in the axial skeleton, and approximately 50% arise in the sacrum.
      • Scott D.
      • Pedlow F.
      • Hecht A.
      • Hornicek F.
      Chapter 11; tumors: primary benign and malignant extradural spine tumors.
      The 5-year survival rate of patients with chordoma is 67%, and the median survival time is 6.3 years.
      • McMaster M.L.
      • Goldstein A.M.
      • Bromley C.M.
      • Ishibe N.
      • Parry D.M.
      Chordoma: incidence and survival patterns in the United States, 1973-1995.
      Surgery is the primary treatment option, with radiotherapy used as an adjuvant.
      • DeLaney T.F.
      • Liebsch N.J.
      • Pedlow F.X.
      • Adams J.
      • Dean S.
      • Yeap B.Y.
      • McManus P.
      • Rosenberg A.E.
      • Nielsen G.P.
      • Harmon D.C.
      • Spiro I.J.
      • Raskin K.A.
      • Suit H.D.
      • Yoon S.S.
      • Hornicek F.J.
      Phase II study of high-dose photon/proton radiotherapy in the management of spine sarcomas.
      Clinical trials using cytotoxic chemotherapy have shown little benefit for the treatment of chordoma; however, initial tests of targeted therapies have shown some promising results in small cohorts of patients.
      • Hofheinz R.D.
      • Kubicka S.
      • Wollert J.
      • Arnold D.
      • Hochhaus A.
      Gefitinib in combination with 5-fluorouracil (5-FU)/folinic acid and irinotecan in patients with 5-FU/oxaliplatin-refractory colorectal cancer: a phase I/II study of the Arbeitsgemeinschaft fur Internistische Onkologie (AIO).
      • Stacchiotti S.
      • Longhi A.
      • Ferraresi V.
      • Grignani G.
      • Comandone A.
      • Stupp R.
      • Bertuzzi A.
      • Tamborini E.
      • Pilotti S.
      • Messina A.
      • Spreafico C.
      • Gronchi A.
      • Amore P.
      • Vinaccia V.
      • Casali P.G.
      Phase II study of imatinib in advanced chordoma.
      • Meng T.
      • Jin J.
      • Jiang C.
      • Huang R.
      • Yin H.
      • Song D.
      • Cheng L.
      Molecular targeted therapy in the treatment of chordoma: a systematic review.
      The molecular mechanisms underlying chordoma are poorly understood; therefore, clinical trials based on genetic mechanisms are limited. Our study attempts to expand our understanding of genetically driven cellular pathways to provide new therapeutic targets to explore.
      Current hypotheses about chordoma development rely mostly on histologic and immunohistochemical studies that show similarities between chordoma and embryonic human notochord. During higher vertebrate development, most fetal notochordal cells regress during embryogenesis; however, remnants within the adult vertebral disk and bone can occur. Benign notochordal cell tumors are relatively common lesions in the adult vertebral bodies, and their etiology is unclear.
      • Picci P.
      • Vanel D.
      • Alberghini M.
      • Mirra J.M.
      • Errani C.
      • Staals E.L.
      • Mercuri M.
      Giant notochordal rests misdiagnosed and treated as chordomas: a retrospective clinical, radiological and histologic study of four cases.
      ,
      • Yamaguchi T.
      • Iwata J.
      • Sugihara S.
      • McCarthy E.F.
      • Karita M.
      • Murakami H.
      • Kawahara N.
      • Tsuchiya H.
      • Tomita K.
      Distinguishing benign notochordal cell tumors from vertebral chordoma.
      In some cases of chordoma, benign notochordal cell tumors have been found within the affected vertebral body,
      • Arain A.
      • Hornicek F.J.
      • Schwab J.H.
      • Chebib I.
      • Damron T.A.
      Chordoma arising from benign multifocal notochordal tumors.
      suggesting that benign notochordal cell tumors may be a precursor lesion.
      Limited molecular examinations support the connection between chordoma and embryonic notochord vestiges. The transcription factor brachyury (TBXT), encoded by the gene TBXT (previously named T), has been previously identified as being important in the development of the normal notochord
      • Kispert A.
      • Koschorz B.
      • Herrmann B.G.
      The T protein encoded by Brachyury is a tissue-specific transcription factor.
      and appears essential for chordoma survival.
      • Presneau N.
      • Shalaby A.
      • Ye H.
      • Pillay N.
      • Halai D.
      • Idowu B.
      • Tirabosco R.
      • Whitwell D.
      • Jacques T.S.
      • Kindblom L.-G.
      • Brüderlein S.
      • Möller P.
      • Leithner A.
      • Liegl B.
      • Amary F.M.
      • Athanasou N.N.
      • Hogendoorn P.C.
      • Mertens F.
      • Szuhai K.
      • Flanagan A.M.
      Role of the transcription factor T (brachyury) in the pathogenesis of sporadic chordoma: a genetic and functional-based study.
      The TBXT gene has been found mutated in some chordomas, and copy number variants have been identified in familial cases.
      • Yang X.R.
      • Ng D.
      • Alcorta D.A.
      • Liebsch N.J.
      • Sheridan E.
      • Li S.
      • Goldstein A.M.
      • Parry D.M.
      • Kelley M.J.
      T (brachyury) gene duplication confers major susceptibility to familial chordoma.
      ,
      • Pillay N.
      • Plagnol V.
      • Tarpey P.S.
      • Lobo S.B.
      • Presneau N.
      • Szuhai K.
      • Halai D.
      • Berisha F.
      • Cannon S.R.
      • Mead S.
      • Kasperaviciute D.
      • Palmen J.
      • Talmud P.J.
      • Kindblom L.G.
      • Amary M.F.
      • Tirabosco R.
      • Flanagan A.M.
      A common single-nucleotide variant in T is strongly associated with chordoma.
      Chordoma cell lines have been shown to depend on TBXT, where its decreased expression leads to cell cycle arrest.
      • Presneau N.
      • Shalaby A.
      • Ye H.
      • Pillay N.
      • Halai D.
      • Idowu B.
      • Tirabosco R.
      • Whitwell D.
      • Jacques T.S.
      • Kindblom L.-G.
      • Brüderlein S.
      • Möller P.
      • Leithner A.
      • Liegl B.
      • Amary F.M.
      • Athanasou N.N.
      • Hogendoorn P.C.
      • Mertens F.
      • Szuhai K.
      • Flanagan A.M.
      Role of the transcription factor T (brachyury) in the pathogenesis of sporadic chordoma: a genetic and functional-based study.
      However, the mechanism that brachyury plays in the context of the disease requires further investigation. TBXT is required for most chordoma tumors to survive, yet it does not have a role in most normal adult tissues, making it an ideal drug target. However, brachyury is a transcription factor, a challenging category of molecules.
      • Chen A.
      • Koehler A.N.
      Transcription factor inhibition: lessons learned and emerging targets.
      Currently, there are no US Food and Drug Administration–approved drugs targeting a transcription factor, but it is an active area of research.
      • Henley M.J.
      • Koehler A.N.
      Advances in targeting “undruggable” transcription factors with small molecules.
      A better understanding of the molecular mechanisms active in chordoma pathogenesis is necessary to provide additional targets for therapeutic development.
      Standard molecular profiling approaches require a control sample for comparison against a disease state, and they generally highlight differences but not similarities between the samples. Chordoma is problematic as its normal tissue counterpart is the notochord, which is only present in the human during the early stages of embryogenesis. To overcome these challenges, we conducted gene expression profiling of chordoma and human notochord and compared it with many non-chordoma tissues to identify the best control tissue and find key pathways associated with tumor pathogenesis. We identified a chordoma diagnostic gene signature, overactive pathways within chordoma cells, and a core gene interaction network required for chordoma survival and proliferation. Finally, shRNA knockdowns and chemical inhibitor experiments support the relevance of the gene network and highlight the importance of transforming growth factor (TGF)-β in chordoma cell survival. We conclude that the pathways related to chondrogenesis are a vital driver of chordoma progression and a promising candidate for therapeutic disruption. Our findings indicate that TGF-β is central to chordoma and is an important druggable target.

      Materials and Methods

      Ethics Statement

      This study used discarded tumor tissues from patients with chordoma as well as notochord cells from discarded human embryonic tissues. The Institutional Review Board at Massachusetts General Hospital (Boston, MA) approved the chordoma study protocol (2009A052093) and the notochord study protocol (2007P-002239). As both protocols used discarded material, the Institutional Review Board waived the need for written informed consent.
      Cell lines U-CH1, U-CH2, U-CH12, HEK293, and 293T were purchased from ATCC (Manassas, VA), and CH22 cells were provided by F.J.H. Cells were cultured in a biosafety level 2 environment according to ATCC guidelines.

      Tumor Sample Processing

      Fresh chordoma specimens from five patients were obtained immediately on surgical resection. The specimens were low in cellularity with abundant extracellular matrix, presenting difficulties in capturing high-quality and adequate amounts of RNA. Tumor samples were cut into 0.5- to 1-cm–diameter pieces. Each sample was either stored in 10 mL of RNAlater (Qiagen, Germantown, MD) or frozen immediately in liquid nitrogen. Homogenization and RNA extraction were performed in the same step. RNA was extracted by placing each tissue sample in 4 mL of Trizol along with 25% (by volume) silicon carbide beads (1 mm; BioSpec, Bartlesville, OK). The tube was placed in a bead beater (BioSpec Mini-Beadbeater-96) and shaken vigorously for 1 to 5 minutes in 30-second intervals. The duration of the homogenization was determined by examining the solution for the disappearance of large masses of tissue. It was not uncommon for tumor samples to contain bone fragments embedded within the tumor mass. In such cases, the tissue was homogenized until only the bone fragments remained intact. The Trizol/tissue homogenate was then transferred to a fresh tube, to which 5 mL of Trizol was added. The tubes were centrifuged for 5 minutes at 12,000 × g to pellet the debris and silicon carbide dust. The homogenate was then transferred to a fresh tube, and 1.6 mL of chloroform was added to the extracted aqueous phase. To precipitate the RNA, 4 mL of isopropanol was immediately added. The pellet was washed in 80% ethanol, dried, and dissolved in 100 μL of water. In cases in which there was still significant visible insoluble material (protein), additional RNA isolation with Trizol was performed on the purified RNA. The RNA solution was then further purified using an RNeasy silica column (Qiagen), according to the manufacturer's protocol. The RNA quality was assessed with a Bioanalyzer RNA Pico kit (Agilent, Santa Clara, CA), and mRNA was purified from samples that passed the quality control step.

      Notochord Laser Microdissection

      Human embryonic notochord specimens were obtained from discarded tissues. To obtain sufficient high-quality mRNA, it was necessary to laser microdissect the cellular vestige of the notochord from the fetal spine. Fetal spinal columns were grossly dissected and frozen to −20°C. The specimens were then mounted on a cryostat chuck using M-1 embedding matrix (ThermoFisher Shandon, Waltham, MA) and sectioned every 25 μm until the central region of the spine was approximated. At this point, the cryomicrotomy blades were changed to a new blade, and 12-μm thin cryosections were mounted on microscopy slides (Gold Seal Rite-On Micro Slides; VWR, Radnor, PA) and immediately processed for laser capture microdissection.
      For the cytoarchitectural visualization of the notochord cells, each tissue section was fixed in 70% ethanol for 30 seconds. The sections were then rinsed with RNase-free distilled water and incubated in HistoGene staining solution (Arcturus; MDS Analytical Technologies, Sunnyvale, CA) for 1 minute, followed by dehydration in increasingly concentrated ethanol (75% to 100%) into xylene and subsequent laser capture microdissection. All incubations and washes were performed at room temperature. Cells representative of the degenerating notochord were clearly visible residing in central lacunae of the nucleus pulposus, and approximately 2000 of these notochord cells from each fetus were captured onto separate polyethylene collecting caps (Macro Cap, Arcturus; MDS Analytical Technologies).
      RNA isolation was performed using the PicoPure RNA isolation system (Arcturus; MDS Analytical Technologies). Plastic laser capture microdissection collecting caps were incubated at 42°C for 30 minutes in 20 μL of the GITC-containing extraction buffer, centrifuged briefly to collect the extracted solution, and then frozen at −80°C. Genomic DNA was removed via RNase-free DNase (Qiagen) digestion on the columns. Total cellular RNA from each column was eluted in a two-step process with 6 μL per step for a total of 12 μL of elution buffer. Isolated RNA was then stored at −80°C until further analysis. The quality of the RNA preparations at various stages was measured using an RNA 6000 Pico chip (Agilent).

      RNA Amplification and Sample Preparation for Microarrays and RNA-Seq

      Given the generally low RNA yield from the chordoma samples, the mRNA was amplified to produce cDNA in sufficient quantity for both microarray and RNA sequencing (RNA-Seq). Single-primer isothermal amplification (Nugen, San Carlos, CA) was used to linearly amplify the purified RNA. For the initial amplification, an Ovation Pico WTA kit (Nugen) was used. Samples destined for microarray were further processed with the Encore Biotin Module. RNA-Seq samples were also processed with the WT-Ovation Exon Module (Nugen) to synthesize the cDNA second strand. High-throughput sequencing libraries for the GAII Illumina platform were constructed using an Illumina kit (Illumina, San Diego, CA). An RNA Amplification System V2 (Ovation) was also used, which includes a complete solution and was not available at the start of the project. In all cases, the manufacturer's directions were followed.

      Microarray Analysis

      Primary normal tissue and cell type expression files were obtained from Gene Expression Omnibus
      • Barrett T.
      • Suzek T.O.
      • Troup D.B.
      • Wilhite S.E.
      • Ngau W.C.
      • Ledoux P.
      • Rudnev D.
      • Lash A.E.
      • Fujibuchi W.
      • Edgar R.
      NCBI GEO: mining millions of expression profiles--database and tools.
      and ArrayExpress.
      • Parkinson H.
      • Kapushesky M.
      • Shojatalab M.
      • Abeygunawardena N.
      • Coulson R.
      • Farne A.
      • Holloway E.
      • Kolesnykov N.
      • Lilja P.
      • Lukk M.
      • Mani R.
      • Rayner T.
      • Sharma A.
      • William E.
      • Sarkans U.
      • Brazma A.
      ArrayExpress--a public database of microarray experiments and gene expression profiles.
      All samples obtained for analysis were profiled on the Affymetrix U133plus2 platform. The authors used version 14 of the custom transcript definition files provided by Brainarray.
      • Dai M.
      • Wang P.
      • Boyd A.D.
      • Kostov G.
      • Athey B.
      • Jones E.G.
      • Bunney W.E.
      • Myers R.M.
      • Speed T.P.
      • Akil H.
      • Watson S.J.
      • Meng F.
      Evolving gene/transcript definitions significantly alter the interpretation of GeneChip data.
      These files redefine Affymetrix probes by remapping individual probes to the human genome and adjusting them to the most up-to-date annotation. The data files were then normalized using the GCRMA module of the Bioconductor software library version 2.22.0,
      • Wu J.
      • Irizarry R.
      • MacDonald J.
      • Gentry J.
      gcrma: Background Adjustment Using Sequence Information.
      and present/absent calls were calculated for each probe using the MAS5 module.
      • Gautier L.
      • Cope L.
      • Bolstad B.M.
      • Irizarry R.A.
      affy—analysis of Affymetrix GeneChip data at the probe level.
      All probes with no present calls were removed and, from the remaining probes, at least one sample was required to have an expression value larger than log2(100). Tumor expression profiles were obtained from Gene Expression Omnibus (https://www.ncbi.nlm.nih.gov/geo/query/acc.cgi?acc=gse2109; accession number GSE2109), and cell line profiles were obtained from the Cancer Cell Line Encyclopedia.
      • Barretina J.
      • Caponigro G.
      • Stransky N.
      • Venkatesan K.
      • Margolin A.A.
      • Kim S.
      • et al.
      The Cancer Cell Line Encyclopedia enables predictive modelling of anticancer drug sensitivity.

      Gene Enrichment Profiles

      The gene enrichment algorithm has previously been described
      • Benita Y.
      • Cao Z.
      • Giallourakis C.
      • Li C.
      • Gardet A.
      • Xavier R.J.
      Gene enrichment profiles reveal T-cell development, differentiation, and lineage-specific transcription factors including ZBTB25 as a novel NF-AT repressor.
      (source code can be found at (http://xavierlab2.mgh.harvard.edu/chordoma/code.html, last accessed November 7, 2022). Briefly, all normal primary cell types and tissues contained replicates, and each such group was compared with each of the other normal tissues using the Limma module of Bioconductor version 3.6.9.
      • Smyth G.K.
      • Michaud J.
      • Scott H.S.
      Use of within-array replicate spots for assessing differential expression in microarray experiments.
      Limma uses linear models and Bayes methods to assess differential expression. For each group, a linear model coefficient was obtained, which is a measure of differences between two cell types. The enrichment score for each probe was defined as the sum of all statistically significant (Bonferroni-adjusted P < 0.05) coefficients. In this study, the authors extended this method further by calculating the gene enrichment in cell lines and tumors. To avoid a bias due to the large number of tumors and cell lines, each individual sample was compared with all normal cell types. Thus, the enrichment value in each tumor or cell line sample reflects enrichment with respect to a body atlas of all normal cell types.

      Assembly of BA0

      To assemble BA0, the authors identified and downloaded 1499 publicly available Affymetrix gene expression microarray data files representing 127 normal human cell types and tissues. Next, the authors filtered and normalized the chordoma and notochord gene expression microarrays together with those of the assembled body atlas as a single data set (BA0). To identify enriched genes, the authors compared chordoma samples pairwise with each primary cell type using a linear model. This comparison yielded 126 linear models, each with a coefficient and an associated P value for each gene. The coefficient is a measure of differences between two samples, with larger coefficient values associated with larger differences. In each comparison, the enrichment score for each gene was defined as the sum of all individual coefficients with an adjusted P < 0.05. Thus, a gene that was highly expressed in one cell type would produce a high enrichment score, as a result of its 126 large and statistically significant coefficient values. This score is comparable between genes within a sample, enabling ranking to identify the specific genes that define chordoma or notochord.

      Assembly of BA1 and Identification of Diagnostic Signature

      The authors obtained 2158 tumor profiles (ExpO data set) and 810 cell lines from the public domain and normalized and processed all samples (including BA0) as a single data set containing 4475 microarrays (BA1). The authors derived enrichment scores for each tumor and cell line by comparing it with all normal cell types, as described above, one sample at a time.
      The authors used two parameters to identify genes that differentiate chordoma from other tumors. First, the authors identified genes enriched within each chordoma sample with a z-score ≥5; this stringent cutoff was used to increase specificity. Next, using the normalized gene expression data, the authors calculated a z-score for each gene across all the tumor samples (expression z-score). The authors deemed genes with a chordoma enrichment z-score ≥5, as well as a chordoma expression z-score ≥5, to be both specific and highly expressed in chordoma.

      RNA-Seq Analysis

      Next-generation sequencing data were first mapped to ribosomal RNA (12s, 16s, 18s, 28s, and 5.8s) using Bowtie version 0.12.7.
      • Langmead B.
      • Trapnell C.
      • Pop M.
      • Salzberg S.L.
      Ultrafast and memory-efficient alignment of short DNA sequences to the human genome.
      All matching reads were discarded. The human genome sequence version HG19 and associated gene models (knownGenes) were obtained from the University of California, Santa Cruz, genome browser.
      • Karolchik D.
      • Kuhn R.M.
      • Baertsch R.
      • Barber G.P.
      • Clawson H.
      • Diekhans M.
      • Giardine B.
      • Harte R.A.
      • Hinrichs A.S.
      • Hsu F.
      • Kober K.M.
      • Miller W.
      • Pedersen J.S.
      • Pohl A.
      • Raney B.J.
      • Rhead B.
      • Rosenbloom K.R.
      • Smith K.E.
      • Stanke M.
      • Thakkapallayil A.
      • Trumbower H.
      • Wang T.
      • Zweig A.S.
      • Haussler D.
      • Kent W.J.
      The UCSC genome browser database: 2008 update.
      Short reads were mapped to the genome using TopHat version 1.4.1.
      • Trapnell C.
      • Pachter L.
      • Salzberg S.L.
      TopHat: discovering splice junctions with RNA-Seq.
      Only reads mapping at least partially to defined exons were retained and were aggregated for a single gene locus. The number of reads per sample was normalized to reads per million; and for each gene, the expression value was calculated as the normalized number of reads/observed gene length. The observed length of a gene was defined as the number of bases detected by sequencing. This adjustment was necessary because the Nugen kit applied to the samples was based on nonrandom probes and resulted in inconsistent coverage of genes with typical 3′-end enrichment (Supplemental Figure S1 provides a representative chordoma/notochord gene coverage plot). rRNA-depleted data, as well as count-based expression profiles generated using STAR version 2.7.10a,
      • Dobin A.
      • Davis C.A.
      • Schlesinger F.
      • Drenkow J.
      • Zaleski C.
      • Jha S.
      • Batut P.
      • Chaisson M.
      • Gingeras T.R.
      STAR: ultrafast universal RNA-seq aligner.
      have been uploaded to Gene Expression Omnibus (https://www.ncbi.nlm.nih.gov/geo/query/acc.cgi?acc=GSE205458; accession number GSE205458).

      Functional Enrichment Analysis

      Pathway data were obtained from the following data sources: MetaBase from GeneGo version 6.3 (Clarivate, Chandler, AZ), Ingenuity Pathway Analysis from Ingenuity Systems (Redwood City, CA), Reactome Pathway,
      • Vastrik I.
      • D'Eustachio P.
      • Schmidt E.
      • Joshi-Tope G.
      • Gopinath G.
      • Croft D.
      • de Bono B.
      • Gillespie M.
      • Jassal B.
      • Lewis S.
      • Matthews L.
      • Wu G.
      • Birney E.
      • Stein L.
      Reactome: a knowledge base of biologic pathways and processes.
      and Gene Ontology.
      • Ashburner M.
      • Ball C.A.
      • Blake J.A.
      • Botstein D.
      • Butler H.
      • Cherry J.M.
      • Davis A.P.
      • Dolinski K.
      • Dwight S.S.
      • Eppig J.T.
      • Harris M.A.
      • Hill D.P.
      • Issel-Tarver L.
      • Kasarskis A.
      • Lewis S.
      • Matese J.C.
      • Richardson J.E.
      • Ringwald M.
      • Rubin G.M.
      • Sherlock G.
      The Gene Ontology Consortium: Gene ontology: tool for the unification of biology..
      To identify pathways and processes that were enriched in a gene list, a hypergeometric-based enrichment analysis was implemented. The hypergeometric P value was calculated using the R program version 3.5.0 (https://www.R-project.org, last accessed November 7, 2022) with the following command: phyper(x − 1, m, n-m, k and lower.tail = FALSE), where x is the number of genes from the gene list that are members of the pathway, m is the number of genes in the pathway, n is the total number of unique genes in all pathways, and k is the number of genes from the list that were present in at least one pathway. The resulting P value is indicative of the likelihood of enrichment for a specific pathway by chance given the size of the gene list.
      This approach typically results in multiple significant pathways because of redundancy. To control for redundancy, this approach was employed iteratively. In each iteration, the most significantly enriched pathway and associated genes from the list were set aside, shortening the gene list. The process was repeated until no pathway was significant (P < 0.05).

      Protein Interactions and Protein Networks

      Protein interaction data were obtained from National Center for Biotechnology Information GeneRIF, MetaBase from GeneGo version 6.3, Ingenuity Pathway Analysis from Ingenuity Systems, and NetPro from Molecular Connections (Bangalore, India). Nonhuman protein interactions were mapped to human homologs using National Center for Biotechnology Information HomoloGene.
      • Wheeler D.L.
      • Barrett T.
      • Benson D.A.
      • Bryant S.H.
      • Canese K.
      • Church D.M.
      • DiCuccio M.
      • Edgar R.
      • Federhen S.
      • Helmberg W.
      • Kenton D.L.
      • Khovayko O.
      • Lipman D.J.
      • Madden T.L.
      • Maglott D.R.
      • Ostell J.
      • JU Pontius
      • Pruitt K.D.
      • Schuler G.D.
      • Schriml L.M.
      • Sequeira E.
      • Sherry S.T.
      • Sirotkin K.
      • Starchenko G.
      • Suzek T.O.
      • Tatusov R.
      • Tatusova T.A.
      • Wagner L.
      • Yaschenko E.
      Database resources of the National Center for Biotechnology Information.
      Data were integrated on the basis of PubMed identifier references supporting each interaction. PubMed identifiers associated with >10 interactions were discarded, and from the remaining interactions, at least two PubMed identifiers were required for the interaction to be retained. Pathway-associated interactions were defined as any of the above interactions that were present in one of the GeneGo pathway maps.
      Interaction networks were simplified by including only interactions that were present in the manually curated GeneGo pathway maps, producing a high-confidence network. Networks were further subjected to functional Reactome, GeneGo, Gene Ontology, and Ingenuity pathway analyses. To eliminate database redundancy, genes were allowed to be associated with only the single-most significant pathway per database.
      All graphical networks were constructed using the OmniGraffle implementation of the GraphViz Dot language (The Omni Group, Seattle, WA).
      All raw and processed profiling data and source code are available for download from a supplementary web page (http://xavierlab2.mgh.harvard.edu/chordoma, last accessed November 7, 2022).

      Lentivirus Production

      shRNA clones were provided by the Broad (within either their pLKO.1 or their pLKO_TRC005 vector) along with packaging (pCMV-dR8.91) and envelope (pMD2.G) plasmids (https://www.broadinstitute.org/rnai-consortium/rnai-consortium-shrna-library, last accessed November 7, 2022). Lentivirus-expressing shRNA against genes of interest was produced as follows: 293T packaging cells were seeded at 2.2 × 105 cells/mL in complete Dulbecco's modified Eagle’s medium in 6-well tissue culture plates. The cells were incubated for about 24 hours in a humidified incubator at 37°C with 5% CO2. When the cells reached 80% confluency, each well was transfected as follows: a mixture of DNA was prepared with 500 ng of packaging plasmid, 50 ng of envelope plasmid, and 500 ng of pooled shRNA vectors targeting the gene of interest (sequences in Supplemental Table S1). In addition, a reagent mix was prepared containing 3 μL of TransIT-LT1 (Mirus Bio; catalog number MIR 2305) and 15 μL of Opti-Mem. Both solutions were mixed after a 5-minute incubation at room temperature, and then the reagent mix was added, dropwise, to the DNA mix. This final mix was incubated at room temperature for 30 minutes and then transferred, dropwise, to the cells. The cells were incubated for 18 hours; then, the medium was removed and replaced with a high-growth medium (Dulbecco's modified Eagle’s medium with 20% serum). The cells were incubated for a further 24 hours, at which point the medium was harvested, stored at 4°C, and replaced with fresh high-growth medium. After another 24 hours, the medium was harvested and pooled with that from the first harvest. The cell plate was then discarded. The pooled medium was spun at 500 × g for 5 minutes to pellet any carryover packaging cells, and then the supernatant was aliquoted into 1-mL aliquots and stored at −80°C. Lenti-X GoStix (Takara, San Jose, CA) was used to verify that the viral concentrations were >5 × 105 IFU/mL.

      Gene Expression Knockdown

      The cells of interest were plated in T25 tissue culture flasks and grown to 80% confluency using the supplier's recommended protocol for each cell line. For each flask, the following mixture was used for the transduction: 3 mL of media, 8 μL of polybrene (4 mg/mL stock), and a 1-mL virus aliquot. The transduction mix was added to the cells and incubated at 37°C. After 24 hours, the medium was changed. After a further 24 hours, the medium was changed to complete medium with 0.5 μg/mL puromycin to start the selection process. After 3 days, the cells were switched to 1 μg/mL puromycin, and split 1:2 if they were approaching confluency. The cells were monitored daily, and split at 80% confluency; at each split, the puromycin concentration was increased by 0.5 μg/mL until the cells were stable at 2 μg/mL puromycin. The cells were eventually transferred to a T75 flask and subsequently subjected to RNA isolation, bromodeoxyuridine (BrdU) treatment, and cytometric profiling.

      BrdU Proliferation Assay

      A BD Biosciences APC BrdU Flow kit (BD Biosciences, Franklin Lakes, NJ; catalog number 552598) was used to measure cell proliferation. BrdU was diluted in culture medium to a 10 μmol/L final concentration and added to 50% confluent cells. The cells were incubated at 37°C/5% CO2 until control cells reached near confluency (up to 1 week for chordoma cell lines), at which point the cells were lifted off the plate using TrypLE Express (ThermoFisher; 12605093). The cells were then fixed and permeabilized, then stained with anti-BrdU antibody in accordance with the manufacturer's recommendations, and analyzed on a Sony SH800 cell sorter (Sony, Tokyo, Japan).

      Annexin and Apoptosis Assay

      Cells were lifted off tissue culture plates with TrypLE Express (ThermoFisher; 12605093) and split into two aliquots. One aliquot was stained with a Molecular Probes Vybrant FAM Poly Caspases Assay Kit (ThermoFisher; V35117) followed by Hoechst 33342 (ThermoFisher; 62249), according to manufacturer's directions. The second aliquot of cells was stained for annexin V as follows: cells were resuspended in 150 μL of 1× Annexin V binding buffer (BD Biosciences; 51-66121E); 5 μL of fluorescein isothiocyanate–annexin V (BioLegend, San Diego, CA; 640905) and 1.2 μL of propidium iodide (BioLegend; 79997) were added; and cells were incubated at room temperature for 20 minutes. Both aliquots of cells (stained for annexin and FAM caspases) were run on a NovoCyte flow cytometer (Agilent) for quantitation and analysis.

      Western Blot Analysis

      To isolate protein for Western blot analysis, cell lines were grown to 80% confluency. Cells were washed with phosphate-buffered saline (PBS), then lifted with TrypLE Express (ThermoFisher; 12605093), spun down at 500 × g for 5 minutes, washed in 13 mL PBS, spun down again, resuspended in 5 mL PBS, and then a 10-μL aliquot was removed and stained with AO/propidium iodide dye (Logos Biosystems, South Korea; F23011) and counted with a Luna FX7 (Logos Biosystems) in fluorescent mode. Cells were spun down once more, supernatant was aspirated, and the pellets were stored at −80°C. Lysis buffer was formulated as follows: PBS; 0.4% Triton X-100 (Sigma, St. Louis, MO; X100-100 ML); 1% Benzonase (Sigma; E1014-5KU); and manufacturer-recommended concentration of PhosSTOP (Roche, Basel, Switzerland; 04 906 837 001) and EDTA-free EASYpack (Roche; 04 693 159 001). Cell pellets were thawed and resuspended in 100 μL lysis buffer per 1 million cells, then incubated at room temperature for 15 minutes. Lysis mixture was spun down at 4°C for 10 minutes at maximum centrifuge speed, and then supernatant was removed to a new tube.
      Electrophoresis was performed using precast NuPAGE gels (Thermo; NP0322BOX). Protein lysate was mixed with concentrated PAGE buffer (with TCEP) to a final 1× PAGE buffer solution, then heated to 95°C for 3 minutes. The 1× SDS running buffer was prepared from a 20× stock (Thermo; NP0002). Gel running box was assembled as per manufacturer's directions, and 20 μL of sample was loaded in each well. A total of 5 μL of ladder (Bio-Rad, Hercules, CA; 1610374) was loaded in end lanes. Electrophoresis was performed at 140 V for 90 minutes.
      Protein was transferred to an LF polyvinylidene difluoride membrane using a Trans-Blot Turbo (Bio-Rad), according to manufacturer's directions (Mixed MW MIDID Run protocol was used, at 2.5 A and 25 V for 7 minutes). Membranes were placed into blocking solution (5% milk in PBS) and incubated at room temperature for 1 hour on a gently shaking platform. Primary antibody solutions were made with 1% milk in PBS with antibody diluted at these ratios: actin (Cell Signaling Technology, Danvers, MA; 8H10D10) 1:1000; ACAN (Thermo; MA3-16888) 1:750; TBXT (Thermo; MA5-17185) 1:1000; WWP2 (Thermo; A302-935A) 1:500; KCNK2 (Thermo; PA1-16981) 1:1000; XYLB (Thermo; 26541-1-AP) 1:500; ENPP1 (Thermo; BS-4913R) 1:250; and SPDYE1 (Thermo; PA5-62470) 1:125. Blocking solution was removed, and membranes were placed into primary antibody solution and left overnight at 4°C on a gently shaking platform.
      The next day, membranes were washed using PBS with 0.1% Tween by rinsing the membrane twice, then performing three washes for 5 minutes each on a gently shaking platform at room temperature. Secondary antibody solutions were made with PBS with 0.1% Tween and these dilutions of antibody: goat anti-mouse (Li-Cor, Lincoln, NE; 926-32210) 1:5000; donkey anti-rabbit (Li-Cor; 926-68073) 1:5000; goat anti-rabbit (Li-Cor; 926-32211) 1:5000; and donkey anti-mouse (Li-Cor; 926-68072) 1:10,000. Membranes were placed into 10 mL of secondary staining mix, and then wrapped in aluminum foil to protect from light. Membranes were incubated on a gently shaking platform at room temperature for 1 hour. Secondary antibody solution was washed off by rinsing the membrane twice in PBS with 0.1% Tween, and then performing three washes for 5 minutes each on a gently shaking platform at room temperature. Membranes were imaged using an Odyssey DLx (Li-Cor).

      Quantitative PCR and Empirical Network Construction

      For laboratory-grown cells, RNA was isolated using a Qiagen RNeasy micro kit (Qiagen; catalog number 74004). The RNA was reverse transcribed into cDNA using an iScript cDNA synthesis kit (Bio-Rad; catalog number 1708891). This cDNA was diluted 1:5.5 with nuclease-free water (final volume, 110 μL), and real-time quantitative PCR was then performed with 5 μL of diluted cDNA, 0.3 μmol/L forward/reverse primer, and 10 μL of SybrGreen Supermix (Bio-Rad; catalog number 1725121) in a 20-μL reaction. The real-time quantitative PCR was thermocycled on a Bio-Rad CFX384 Touch with the following program: 95°C for 3 minutes, followed by 46 cycles of 95°C for 10 seconds and 60°C for 30 seconds. The fluorescence measurements were performed every cycle, and the Ct measurements were calculated with the Bio-Rad software. All downstream calculations were performed on the average of three technical replicates. Measurements for each sample were normalized to those of glyceraldehyde-3-phosphate dehydrogenase; then, the fold change was calculated on the basis of cycle-count differences between the test and the shGFP control samples. The empirical network was generated using the R iGraph package version 1.2.4.
      • Csardi G.
      • Nepusz T.
      The igraph software package for complex network research. InterJournal.
      In the network graph, nodes represent the target shRNA, and edges (arrows) point to the regulated gene. Only relationships with a fold change >1.5 or <1/1.5 were kept.

      Drug Treatment Viability Assays

      HEK293 or U-CH2 cells were plated into 96-well plates at a density of 15,000 or 10,000 cells per well, respectively. Wells containing U-CH2 were previously coated with collagen and fibronectin. The plates used were black tissue culture plates with clear bottoms (Corning, Corning, NY; 3904). The reagents from the RealTime-Glo assay kit (Promega, Madison, WI) were added at the time of plating, according to the manufacturer's recommendations. Luciferase measurements were made using a Microbeta2 (Perkin Elmer, Waltham, MA) at 1 hour predosing, and the following post-dose time points: 1, 3, 19.5, 24, 27.5, 44, and 48 hours. The drug LDN-212854 (Selleck Chemicals, Houston, TX) was reconstituted in dimethyl sulfoxide at 100 mmol/L and dosed at concentrations from 1 to 100 μmol/L 24 hours after cell plating. The luciferase values for each well were normalized to the predose measurement, and dose-response curves were calculated using the DRC package version 3.0-1.
      • Ritz C.
      • Baty F.
      • Streibig J.C.
      • Gerhard D.
      Dose-response analysis using R.
      The normalized luciferase values for each time point/cell line combination were fitted with a three-parameter log-logistic function using the LL.3() function call, then EC50 values were calculated using the ED() function, and 95% CIs were calculated using the delta method. Relative potency was calculated using the EDcomp() function, and represents the EC50 of HEK293 divided by the EC50 of U-CH2. CIs were again calculated using the delta method. EC50 values for time points <19.5 hours were excluded from bar plots comparing U-CH2 and HEK293 because these early time points did not exhibit dose-dependent killing of HEK293 cells, and an accurate dose-response curve could not be fitted. The results were plotted using the ggplot2 R package version 3.3.6 (https://ggplot2.tidyverse.org, last accessed November 7, 2022).

      Results

      Chordoma and Notochord Gene Expression Profiles Are Closest Compared with All Other Examined Profiles

      The authors developed a method to extract high-quality RNA from chordomas and profiled five non-clival chordoma tumors using RNA-Seq. The authors also profiled three of the tumor samples using Affymetrix gene expression microarrays. The authors obtained notochordal cells through the laser microdissection of three human fetal notochord tissue samples, and the authors profiled the cells using both RNA-Seq and Affymetrix microarrays (Gene Expression Omnibus study, https://www.ncbi.nlm.nih.gov/geo/query/acc.cgi?acc=GSE205458; accession number GSE205458).
      Comparing chordoma and notochord gene expression identified differences between the two tissues, but not similarities that could be consequential for tumor progression. For instance, TBXT has been previously identified as critical for chordoma proliferation,
      • Presneau N.
      • Shalaby A.
      • Ye H.
      • Pillay N.
      • Halai D.
      • Idowu B.
      • Tirabosco R.
      • Whitwell D.
      • Jacques T.S.
      • Kindblom L.-G.
      • Brüderlein S.
      • Möller P.
      • Leithner A.
      • Liegl B.
      • Amary F.M.
      • Athanasou N.N.
      • Hogendoorn P.C.
      • Mertens F.
      • Szuhai K.
      • Flanagan A.M.
      Role of the transcription factor T (brachyury) in the pathogenesis of sporadic chordoma: a genetic and functional-based study.
      but is highly expressed in both chordoma and notochord tissues; from a differential expression standpoint, it is thus indistinguishable from other housekeeping genes that would also be expressed at similar levels, and would not be identified as a driver gene. Therefore, the authors developed an unbiased method to compare chordoma and notochord tissues, while also identifying similarities with other tissues that might enhance the mechanistic knowledge surrounding chordoma tumorigenesis. The authors used this gene enrichment method to compare the expression levels of each gene against those in a body atlas data set of normal primary tissues. The authors assembled this body atlas, BA0, from 1499 publicly available Affymetrix gene expression microarray data files that span 127 normal human cell types/tissues.
      Within BA0, comparing the 20 genes most enriched in chordoma with the top 20 genes for notochord revealed five genes, including COL2A1 (collagen 2A1), TBXT (brachyury), CA3 (carbonic anhydrase 3), ACAN (aggrecan), and CD109, that were enriched in both tissue types (Figure 1A and Supplemental Figure S2, A and B). In addition, by exploring the scores of individual genes enriched in chordoma, the authors found that TBXT enrichment was strictly limited to chordoma and notochord, whereas COL2A1, ACAN, and KCNK2 were also highly enriched in fetal cartilage, chondrocytes, and notochord (Supplemental Figure S2C). Finally, the authors examined all genes with a z-score ≥4 within chordoma or notochord (Figure 1A). The authors identified 28 genes enriched in both tissues, 45 genes enriched only in chordoma, and 57 genes enriched only in notochord (Supplemental Table S2). This overlap is highly significant because the authors were unable to achieve an overlap of more than six genes in 100,000 iterations when randomly selecting genes from the chordoma and notochord lists.
      Figure thumbnail gr1
      Figure 1Analysis of chordoma-enriched genes. A: Scatterplot of genes enriched in chordoma and notochord selected using a z-score ≥4 within chordoma or notochord. Red dots indicate genes enriched in both samples, blue dots indicate genes enriched only in chordoma, and green dots indicate genes enriched only in notochord. A curated selection of potentially important genes is labeled on the plot; the full list of genes is provided in . B: Twelve genes were identified as being highly enriched in chordoma compared with normal tissues, tumors, and cell lines. The average enrichment score of these genes was used as a chordoma signature score and applied to the entire data set. The top 15 scoring samples are shown with a breakdown of the enrichment score for the individual genes composing the signature. C: Validation of the chordoma signature genes in an independent body atlas data set of publicly available chordoma samples. Eight genes from the chordoma signature were available on this platform and were used to rank chordoma and 126 normal tissues and cell types. Chordoma is the top-scoring tissue, as shown by the bar plot.

      Chordoma Is Characterized by a 12-Gene Diagnostic Signature

      Having identified gene expression signatures specific to notochord and chordoma, the authors next sought to assess whether a gene expression signature could be used to diagnose chordoma. To generate this gene expression signature, the authors compared chordoma-enriched genes with those enriched in other tumors to exclude general cancer-associated genes. To perform this comparison, the authors generated another body atlas, BA1, that includes both normal and tumor samples.
      Consequently, the authors identified a panel of 12 genes that are more highly expressed in all chordoma samples than in normal tissues, tumors, and cell lines, and thus form a chordoma signature. Next, the authors used the mean enrichment score of the 12 genes as a chordoma signature score to rank tumors and cell lines. The authors found that of all the unique samples in the data set (BA0 + ExpO + cell lines), notochord, the chordoma-derived cell line UCH1, fetal and bone cartilage, and chondrocytes scored the highest (Figure 1B).
      To validate the chordoma diagnostic signature, the authors assembled a second independent body atlas (BA2). BA2 included the only chordoma samples available in the public domain at the time and 557 normal samples representing 126 cell types and tissues. All BA2 samples were profiled on the Affymetrix U133A platform. This data set had a different cell type composition compared with BA1, with the most obvious difference being the larger number of different immune cell types (approximately 30% of the panel). The authors normalized and processed the BA2 data set for enrichment as described for BA1. Of the 12 genes identified as the chordoma signature, 8 were available on the U133A platform. Chordoma was the top-scoring group in the BA2 data set, enriched for all eight genes (Figure 1C). For further validation, the authors performed Western blot analysis and validated the expression of a selection of these genes in chordoma cell lines (Supplemental Figure S2D).

      Functional Analysis of Chordoma Genes

      Having defined a core chordoma gene expression signature, the authors next sought to understand the molecular mechanism. First, the authors identified the chordoma signature genes with shared functionality. The authors generated a protein interaction network using the 12 signature genes, allowing for a single intermediate chordoma-expressed gene to serve as a link between each of the signature genes. The resulting network of 7 signature proteins and 13 intermediate proteins is shown in Figure 2A. The network appears centered around TGF-β, SOX9, and the extracellular matrix proteins COL2A1 and ACAN, which are direct transcriptional targets of SOX9.
      Figure thumbnail gr2
      Figure 2Identification of a transforming growth factor (TGF)-β/SOX6/SOX9 signaling hub. A: A protein interaction network of chordoma signature genes and chordoma-expressed genes. One intermediate gene (colored in gray) was allowed to connect signature genes. All interactions were manually curated and are supported by at least two references. Bold edges indicate interactions present in curated pathway maps. B: Gene members of the cartilage development pathway that are enriched in chordoma. Genes are colored by the number of chordoma tumors in which they were found to be enriched (z-score > 4). The circle sizes are relative to the number of non-chordoma tumors in which these genes were found to also be enriched. Edges are shown for the genes that interact directly within the pathway. C: Heat map illustrating the effect of U-CH2 gene expression knockdown on selected chordoma-related genes. U-CH2 cells were transduced with shRNA constructs targeting individual genes listed on the x axis. RNA expression of the genes listed on the y axis is assessed by real-time quantitative PCR (qPCR). The expression levels for each sample were normalized via glyceraldehyde-3-phosphate dehydrogenase (GAPDH) expression, and each sample was subsequently normalized to represent expression fold-change values relative to the expression of the green fluorescent protein (GFP) control sample. D: Network graph visualization of the qPCR results shown in C. The nodes are the probed genes. The edges point from the knocked-down gene to a downstream regulated gene; the size of each edge is scaled by the magnitude of the expression change in the regulated gene; the edges are colored blue to indicate expression down-regulation and are colored red to indicate expression up-regulation. The size of each node is scaled on the basis of the number of shRNAs that caused a change in the expression of the gene. TBXT, brachyury.
      In addition to analyzing the chordoma signature genes, the authors analyzed genes that had enriched expression in chordoma compared with normal cells and tissues. Initially, the authors generated seven chordoma gene lists (genes with a z-score >4) made from the three chordoma samples and four publicly available samples. The authors analyzed each gene list individually for pathway enrichment using GeneGo pathway maps, and found cartilage development/chondrogenesis and cell adhesion ranked as the two most significant pathways (Supplemental Figure S3A). To assess the significance of this finding, the authors repeated the same procedure for each of the 2158 tumor samples in the ExpO data set. From this analysis, the authors identified a similar list of pathways as identified in the chordoma samples (Supplemental Figure S3B). Further exploration of this resemblance revealed that although similar pathways were enriched in chordoma and non-chordoma tumor samples, the genes within the pathways differed. For instance, within the cartilage development pathway, ACAN, COL2A1, SOX9, SOX6, CTGF, FN1, and others were enriched in chordoma tumors but rarely enriched in other tumor types (Figure 2B). In contrast, collagen type I and III genes, enriched in some chordomas, were also frequently enriched in other tumors. In summary, chordoma signature genes participate in pathways common to many other types of cancer, yet frequently represent unique portions of the pathways.
      Although the data set includes only a small number of chordoma tumors, the authors sought to identify additional pathways that might play a role in chordoma and to identify potential therapeutic targets for further validation. To this end, the authors repeated the interaction network analysis above and included all chordoma genes that were enriched in at least two of the three chordoma samples and expressed on average ≥20 reads per million in the RNA-Seq data; the authors identified 62 such genes (Supplemental Table S3). Next, the authors generated a protein interaction network that included the proteins encoded by the 62 chordoma-enriched genes and the chordoma-expressed genes (Supplemental Table S4) that serve as a link between at least two nodes among the 62 genes. The resulting network was dense and revealed an interconnected TGF-β–SOX9 hub, an MDM2-p53 hub, and a RELA, β-catenin, and HIF-1 hub (Supplemental Figure S4A). The authors simplified the network by including only high-confidence interactions (Supplemental Figure S4B) and then enhanced this with pathway analysis. The resulting analysis is illustrated as a network in Supplemental Figure S5, where each edge represents pathway membership, as determined through the analysis. The pathways and processes identified using this method included cartilage development, integrin binding, Wnt signaling, NF-κB activation, and angiogenesis. Taken together, all the functional analyses suggest the presence of a forward-feedback loop where the TGF-β pathway is activated by extracellular matrix via integrin signaling, leading to SOX9 activation and additional extracellular matrix generation. This integrative model is illustrated in Supplemental Figure S6.

      Brachyury Core Network Members Are Highly Interconnected

      The computational analysis identified a core chordoma network that encompasses members with published interactions, so the authors sought to identify a central hub that might present a therapeutic target. The authors knocked down the expression of SOX5, SOX6, SOX9, TBXT, and TGFΒ1 via shRNA and examined the effect of the knockdown on the expression of other genes of interest using real-time quantitative PCR. The expression results were normalized using GAPDH expression; then, the fold induction was calculated relative to that of a control sample expressing shRNA against green fluorescent protein. The results are shown in Figure 2C. Notably, TGFΒ1 expression increased significantly as a result of the knockdown of any other network member (aside from TGFΒ1 itself). To further illustrate the interconnectedness, the authors generated a network diagram representing the relationships empirically identified via real-time quantitative PCR; the authors considered each edge (arrow) to represent a regulatory effect from the sending node (the one probed via shRNA) to the receiving node (gene with altered expression level) (Figure 2D). These experiments supported TGF-β as a central hub, exhibiting both regulation of and by other genes in the network.

      Core Chordoma Network Members Are Required for Chordoma Cell Line Survival

      Having established a core interconnected chordoma network, the authors next aimed to assess the reliance of chordoma cells on the identified network genes. The authors knocked down the expression of SOX6, SOX9, TGFB1, and TBXT using lentiviral shRNA constructs and selected for cells stably expressing the relevant shRNAs. Subsequently, the authors examined cell proliferation in each case by BrdU incorporation in newly synthesized DNA, as measured by flow cytometry with an anti-BrdU antibody. SOX6, SOX9, and TBXT knockdown resulted in severely decreased proliferation, whereas TGFB1 knockdown resulted in only a modest decrease (Figure 3A). However, the BrdU incorporation assay did not demonstrate the extent of visually observed cell death caused by the gene knockdowns (especially in the case of TGFB1); it only measured the proliferation in the surviving cells. To address this gap, the authors collected media from TGFB1 and TBXT knockdown cells and quantitatively measured annexin V binding and propidium iodide staining. Most floating cells were dead, regardless of which gene was knocked down. However, TGFΒ1 knockdown resulted in a far greater number of floating dead cells than TBXT knockdown. Furthermore, the authors assayed the surviving cells from shRNA knockdowns for early and late apoptosis. The authors stained with annexin V and propidium iodide to measure late apoptosis, and found TGFΒ1 knockdown yielded higher proportions of apoptotic cells than either TBXT knockdown or the negative control sample (shRNA against green fluorescent protein) (Figure 3). The authors also used a caspase assay to measure early apoptosis and found TGFΒ1 knockdown cells had the highest proportion of dying cells (Figure 3). To verify whether the TGFB1 knockdown-induced apoptosis is a chordoma-specific phenomenon, the authors performed the same annexin and early apoptosis assay in HEK293 cells expressing shRNAs against either TGFB1 or green fluorescent protein. The authors found HEK293 cells did not exhibit any increased apoptosis after TGFB1 knockdown compared with the green fluorescent protein negative control (Supplemental Figure S7). In summary, all members of the core chordoma network are necessary for cell survival, with some genes exhibiting a dichotomy between proliferative and immediate apoptotic effects. TGFΒ1 knockdown results in moderately decreased proliferation and a strong induction of apoptosis specifically in chordoma cells, whereas TBXT knockdown results in severely reduced proliferation and mildly increased cell death during the experimental time frame.
      Figure thumbnail gr3
      Figure 3TGFB1 knockdown results in apoptotic phenotype. A: Phase-contrast microscopy images demonstrate the hyperapoptotic phenotype after TGFB1 knockdown. Cells were imaged in a time course after selection for integration of lentiviral vector. Day 1 represents the first day after antibiotic selection was finished. By day 9, the sh-GFP control knockdown cells exhibit proliferation consistent with U-CH2 cells, whereas brachyury (TBXT) knockdown cells have not expanded much. In contrast, TGFB1 knockdown results in extensive cell death, with remaining cells lacking the chordoma characteristic physaliferous phenotype and presenting as extremely unhealthy. BD: Cytometric studies characterize the proportion of green fluorescent protein (GFP) knockdown cells (B), TBXT knockdown cells (C), or TGFB1 knockdown cells (D) that are in the early stage of apoptosis, late stage of apoptosis, or actively proliferating. E: Bar plot summarizing the cytometric proportions characterized in B through D. BrdU, bromodeoxyuridine; FITC, fluorescein isothiocyanate.

      TGF-β Pathway Inhibition Limits the Growth of Chordoma Cell Lines

      Computational analysis and experimental data support the hypothesis that TGF-β acts as a hub for chordoma growth. Several chemical inhibitors of the TGF-β pathway are available, and the authors used LDN-212854 (which acts primarily through inhibition of ALK2 and ALK1
      • Mohedas A.H.
      • Xing X.
      • Armstrong K.A.
      • Bullock A.N.
      • Cuny G.D.
      • Yu P.B.
      Development of an ALK2-biased BMP type I receptor kinase inhibitor.
      ) to interrogate the effect of TGF-β pathway inhibition in chordoma cell lines. The authors treated HEK293 (control) and U-CH2 chordoma cells with varying concentrations of LDN-212854 (Selleck Chemicals) and measured viability using a RealTime-Glo viability assay kit (Promega). The authors found LDN-212854 has about a twofold higher relative potency in U-CH2 cells over HEK293, with U-CH2 cells exhibiting strong growth inhibition at lower drug concentrations than HEK293 (Figure 4). U-CH2 cells also exhibited characteristic dose-dependent killing at earlier time points than HEK293 cells (Figure 4). A more comprehensive survey of all TGF-β pathway inhibitors might identify more potent and practical molecules. These findings suggest that TGF-β pathway inhibition is a viable chordoma treatment strategy.
      Figure thumbnail gr4
      Figure 4Chordoma cells respond faster and with higher sensitivity to transforming growth factor (TGF)-β pathway inhibition compared with HEK293 controls. HEK293 or U-CH2 chordoma cells were treated with varying doses of LDN212854, a potent inhibitor of a receptor integral to the TGF-β signaling pathway. Cell viability was measured at multiple time points via a luciferase assay. A dose-response curve was fitted for each time point and colored according to cell type. EC50 values were calculated for each cell line at each time point and plotted as a bar plot. Relative potency was calculated as the EC50 ratio between HEK293 and U-CH2 cells (a solid red line is shown at 1.0, indicating the level of no difference in potency between HEK293 and U-CH2 cells).

      Discussion

      In this study, we utilized molecular profiling approaches to compare human notochord and chordoma to provide insight into alternative and potentially synergistic therapeutic targets in addition to brachyury, an established critical transcription factor.
      • Presneau N.
      • Shalaby A.
      • Ye H.
      • Pillay N.
      • Halai D.
      • Idowu B.
      • Tirabosco R.
      • Whitwell D.
      • Jacques T.S.
      • Kindblom L.-G.
      • Brüderlein S.
      • Möller P.
      • Leithner A.
      • Liegl B.
      • Amary F.M.
      • Athanasou N.N.
      • Hogendoorn P.C.
      • Mertens F.
      • Szuhai K.
      • Flanagan A.M.
      Role of the transcription factor T (brachyury) in the pathogenesis of sporadic chordoma: a genetic and functional-based study.
      The tumors we profiled were all non-clival in origin. Although both clival and sacral tumors are known to express and be dependent on brachyury, clival and sacral tumors have been shown to have differences in other pathway dependencies.
      • Jäger D.
      • Barth T.F.E.
      • Brüderlein S.
      • Scheuerle A.
      • Rinner B.
      • von Witzleben A.
      • Lechel A.
      • Meyer P.
      • Mayer-Steinacker R.
      • Baer Avon
      • Schultheiss M.
      • Wirtz C.R.
      • Möller P.
      • Mellert K.
      HOXA7, HOXA9, and HOXA10 are differentially expressed in clival and sacral chordomas.
      ,
      • Lohberger B.
      • Scheipl S.
      • Heitzer E.
      • Quehenberger F.
      • de Jong D.
      • Szuhai K.
      • Liegl-Atzwanger B.
      • Rinner B.
      Higher cMET dependence of sacral compared to clival chordoma cells: contributing to a better understanding of cMET in chordoma.
      Our results were generated from sacral tumors, and we used sacral cell lines to perform validation experiments—thus, we can only confidently apply the results here to tumors of a sacral origin, which is the predominant anatomic location for chordomas.
      • Bakker S.H.
      • Jacobs W.C.H.
      • Pondaag W.
      • Gelderblom H.
      • Nout R.A.
      • Dijkstra P.D.S.
      • Peul W.C.
      • Vleggeert-Lankamp C.L.A.
      Chordoma: a systematic review of the epidemiology and clinical prognostic factors predicting progression-free and overall survival.
      We found human notochord to have the closest resemblance to chordoma among the hundreds of tissues and cancers we surveyed, providing strong molecular support to the close relationship between notochord and chordoma previously observed via histology and immunohistochemistry. In addition, we identified a TGF-β/SOX6/SOX9/TBXT pathway that appears to play a key role in chordoma development and survival, with TGF-β serving as a critical member. TGF-β presents an attractive therapeutic target, which we support using both genetic (shRNA knockdown) and chemical (TGF-β pathway inhibitor) methods.
      We needed to address two challenging issues that have previously made the molecular profiling and analysis of chordoma difficult. First, the extraction of high-quality RNA from chordoma specimens was addressed, which is notoriously difficult because of the low cellularity and large quantity and composition of the extracellular matrix.
      • Baelde H.J.
      • Cleton-Jansen A.M.
      • van Beerendonk H.
      • Namba M.
      • Bovee J.V.
      • Hogendoorn P.C.
      High quality RNA isolation from tumours with low cellularity and high extracellular matrix component for cDNA microarrays: application to chondrosarcoma.
      ,
      • Scheil S.
      • Bruderlein S.
      • Liehr T.
      • Starke H.
      • Herms J.
      • Schulte M.
      • Moller P.
      Genome-wide analysis of sixteen chordomas by comparative genomic hybridization and cytogenetics of the first human chordoma cell line, U-CH1.
      The second issue addressed problems in identifying both similarities and differences in gene expression between chordoma and notochord. We developed a method that compares gene expression in individual samples with that in a body atlas of normal tissues and tumors, a method we previously applied to the identification of cell- and tissue-specific genes.
      • Benita Y.
      • Cao Z.
      • Giallourakis C.
      • Li C.
      • Gardet A.
      • Xavier R.J.
      Gene enrichment profiles reveal T-cell development, differentiation, and lineage-specific transcription factors including ZBTB25 as a novel NF-AT repressor.
      In the context presented here, this method allowed us to identify genes that define chordoma and notochord independently, eliminating the need to compare the two directly and rely on assumptions about proper control tissues. We considered that the composition of the body atlas might bias the results of this analysis; however, we found that observations made in one body atlas can be validated in a second body atlas with a different sample composition and a different platform. This approach can be extended to other cancer studies and types of data, and it will be useful for cases where control tissues are unavailable, unknown, or difficult to obtain. Examples of such studies include nervous system tumors, sarcomas, other heterogeneous tumors, and tumors with an extreme mutational burden.
      The comparison of chordoma and notochord in the context of our body atlas showed high molecular resemblance. The chordoma gene-expression signature is more similar to that of notochord than that of any other cancer. In addition, although some genes are different between chordoma and notochord, the diagnostic signature separating chordoma from thousands of tumors and normal cells does not distinguish chordoma from notochord (Figure 1, A and B). Of the top 10 chordoma-enriched genes (Figure 1A and Supplemental Figure S2A), 9, including TBXT, are also enriched in notochord. These data provide novel and strong molecular evidence suggesting that chordoma is strongly related to notochord.
      To date, most chordoma studies have primarily focused on brachyury, which is used clinically as a diagnostic marker and has been implicated in chordoma by several genetic studies.
      • Yang X.R.
      • Ng D.
      • Alcorta D.A.
      • Liebsch N.J.
      • Sheridan E.
      • Li S.
      • Goldstein A.M.
      • Parry D.M.
      • Kelley M.J.
      T (brachyury) gene duplication confers major susceptibility to familial chordoma.
      ,
      • Pillay N.
      • Plagnol V.
      • Tarpey P.S.
      • Lobo S.B.
      • Presneau N.
      • Szuhai K.
      • Halai D.
      • Berisha F.
      • Cannon S.R.
      • Mead S.
      • Kasperaviciute D.
      • Palmen J.
      • Talmud P.J.
      • Kindblom L.G.
      • Amary M.F.
      • Tirabosco R.
      • Flanagan A.M.
      A common single-nucleotide variant in T is strongly associated with chordoma.
      Functionally, the role of brachyury in cancer is unknown, although some studies have suggested that it contributes to a mesenchymal phenotype.
      • Fernando R.I.
      • Litzinger M.
      • Trono P.
      • Hamilton D.H.
      • Schlom J.
      • Palena C.
      The T-box transcription factor Brachyury promotes epithelial-mesenchymal transition in human tumor cells.
      • Roselli M.
      • Fernando R.I.
      • Guadagni F.
      • Spila A.
      • Alessandroni J.
      • Palmirotta R.
      • Costarelli L.
      • Litzinger M.
      • Hamilton D.
      • Huang B.
      • Tucker J.
      • Tsang K.Y.
      • Schlom J.
      • Palena C.
      Brachyury, a driver of the epithelial-mesenchymal transition, is overexpressed in human lung tumors: an opportunity for novel interventions against lung cancer.
      • Du R.
      • Wu S.
      • Lv X.
      • Fang H.
      • Wu S.
      • Kang J.
      Overexpression of brachyury contributes to tumor metastasis by inducing epithelial-mesenchymal transition in hepatocellular carcinoma.
      Our data confirm that brachyury is a unique marker of chordoma and notochord, and it is rarely expressed in other normal or cancer tissues. Knockdown experiments further confirm its essential role in underlying chordoma proliferation, although it is unclear whether brachyury's primary role is in tumor onset or maintenance/growth. In addition, our experiments suggest that brachyury participates in the TGF-β pathway. These observations suggest a potential role for brachyury in chordoma that needs further study.
      Our analysis of the molecular mechanism driving chordoma further revealed that the process of chondrogenesis is the most significant and highly enriched cellular pathway. Although this pathway is common to many cancer types,
      • Kang H.S.
      • Ahn J.M.
      • Kang Y.
      Chondrogenic tumors.
      • Huh Y.H.
      • Ryu J.-H.
      • Shin S.
      • Lee D.-U.
      • Yang S.
      • Oh K.-S.
      • Chun C.-H.
      • Choi J.-K.
      • Song W.K.
      • Chun J.-S.
      Esophageal cancer related gene 4 (ECRG4) is a marker of articular chondrocyte differentiation and cartilage destruction.
      • Matsumoto Y.
      • Sato S.
      • Maeda T.
      • Kishino M.
      • Toyosawa S.
      • Usami Y.
      • Iwai S.
      • Nakazawa M.
      • Yura Y.
      • Ogawa Y.
      Transcription factors related to chondrogenesis in pleomorphic adenoma of the salivary gland: a mechanism of mesenchymal tissue formation.
      • Robert A.W.
      • Marcon B.H.
      • Dallagiovanna B.
      • Shigunov P.
      Adipogenesis, osteogenesis, and chondrogenesis of human mesenchymal stem/stromal cells: a comparative transcriptome approach.
      • Murakami S.
      • Lefebvre V.
      • de Crombrugghe B.
      Potent inhibition of the master chondrogenic factor Sox9 gene by interleukin-1 and tumor necrosis factor-α.
      we found that the chordoma-relevant chondrogenesis genes are not typically expressed in other cancers. Furthermore, our protein interaction network analysis revealed that the extracellular matrix proteins COL2A1 and ACAN, integrins, TGF-β, and SOX9 are all key players in the process of chondrogenesis.
      Integrins are adhesion molecules that play an important role in the initiation and progression of cancer,
      • Desgrosellier J.S.
      • Cheresh D.A.
      Integrins in cancer: biological implications and therapeutic opportunities.
      and previous studies have shown that COL2A1 can bind and activate integrins on chondrocytes.
      • Camper L.
      • Hellman U.
      • Lundgren-Akerlund E.
      Isolation, cloning, and sequence analysis of the integrin subunit alpha10, a beta1-associated collagen binding integrin expressed on chondrocytes.
      ,
      • Hynes R.O.
      Integrins: bidirectional, allosteric signaling machines.
      Furthermore, integrins are modulators of the TGF-β pathway,
      • Munger J.S.
      • Huang X.
      • Kawakatsu H.
      • Griffiths M.J.
      • Dalton S.L.
      • Wu J.
      • Pittet J.F.
      • Kaminski N.
      • Garat C.
      • Matthay M.A.
      • Rifkin D.B.
      • Sheppard D.
      The integrin alpha v beta 6 binds and activates latent TGF beta 1: a mechanism for regulating pulmonary inflammation and fibrosis.
      ,
      • Ludbrook S.B.
      • Barry S.T.
      • Delves C.J.
      • Horgan C.M.
      The integrin alphavbeta3 is a receptor for the latency-associated peptides of transforming growth factors beta1 and beta3.
      which is well characterized in cancer as an inducer of epithelial-to-mesenchymal transformation.
      • Wendt M.K.
      • Tian M.
      • Schiemann W.P.
      Deconstructing the mechanisms and consequences of TGF-beta-induced EMT during cancer progression.
      ,
      • Bachman K.E.
      • Park B.H.
      Duel nature of TGF-beta signaling: tumor suppressor vs. tumor promoter.
      TGF-β treatment of mesenchymal stem cells can also induce a chondrogenic phenotype that induces COL2A1 and ACAN.
      • Mehlhorn A.T.
      • Schmal H.
      • Kaiser S.
      • Lepski G.
      • Finkenzeller G.
      • Stark G.B.
      • Sudkamp N.P.
      Mesenchymal stem cells maintain TGF-beta-mediated chondrogenic phenotype in alginate bead culture.
      A link between TGF-β and chordoma has been suggested by a study of copy number alteration in 21 chordoma specimens, in which a deletion of 1p36 was identified in 90% of chordoma samples.
      • Le L.P.
      • Nielsen G.P.
      • Rosenberg A.E.
      • Thomas D.
      • Batten J.M.
      • Deshpande V.
      • Schwab J.
      • Duan Z.
      • Xavier R.J.
      • Hornicek F.J.
      • Iafrate A.J.
      Recurrent chromosomal copy number alterations in sporadic chordomas.
      The tumor suppressor implicated in this region is RUNX3, a known repressor of TGF-β. In addition, a significant number of tumors had genomic amplifications of the TGFΒ1 gene, a key regulator of the pathway.
      • Le L.P.
      • Nielsen G.P.
      • Rosenberg A.E.
      • Thomas D.
      • Batten J.M.
      • Deshpande V.
      • Schwab J.
      • Duan Z.
      • Xavier R.J.
      • Hornicek F.J.
      • Iafrate A.J.
      Recurrent chromosomal copy number alterations in sporadic chordomas.
      Later studies have also found TGF-β signaling
      • Duan W.
      • Zhang B.
      • Li X.
      • Chen W.
      • Jia S.
      • Xin Z.
      • Jian Q.
      • Jian F.
      • Chou D.
      • Chen Z.
      Single-cell transcriptome profiling reveals intra-tumoral heterogeneity in human chordomas.
      and TGFB1 expression
      • Ma J.
      • Tian K.
      • Wang L.
      • Wang K.
      • Du J.
      • Li D.
      • Wu Z.
      • Zhang J.
      High expression of TGF-β1 predicting tumor progression in skull base chordomas.
      associated with tumor progression, whereas down-regulation of TGFB3 has been reported to cause chordomagenesis.
      • Wang L.
      • Guan X.
      • Hu Q.
      • Wu Z.
      • Chen W.
      • Song L.
      • Wang K.
      • Tian K.
      • Cao C.
      • Zhang D.
      • Ma J.
      • Tong X.
      • Zhang B.
      • Zhang J.
      • Zeng C.
      TGFB3 downregulation causing chordomagenesis and its tumor suppression role maintained by Smad7.
      The TGF-β pathway plays a well-established role in chondrogenesis through the activation of SOX9,
      • Lorda-Diez C.I.
      • Montero J.A.
      • Martinez-Cue C.
      • Garcia-Porrero J.A.
      • Hurle J.M.
      Transforming growth factors beta coordinate cartilage and tendon differentiation in the developing limb mesenchyme.
      • Wu M.Y.
      • Hill C.S.
      Tgf-beta superfamily signaling in embryonic development and homeostasis.
      • Furumatsu T.
      • Tsuda M.
      • Taniguchi N.
      • Tajima Y.
      • Asahara H.
      Smad3 induces chondrogenesis through the activation of SOX9 via CREB-binding protein/p300 recruitment.
      a transcription factor that directly induces COL2A1 and ACAN expression.
      • Oh C.D.
      • Maity S.N.
      • Lu J.F.
      • Zhang J.
      • Liang S.
      • Coustry F.
      • de Crombrugghe B.
      • Yasuda H.
      Identification of SOX9 interaction sites in the genome of chondrocytes.
      • Nakamura Y.
      • Yamamoto K.
      • He X.
      • Otsuki B.
      • Kim Y.
      • Murao H.
      • Soeda T.
      • Tsumaki N.
      • Deng J.M.
      • Zhang Z.
      • Behringer R.R.
      • Crombrugghe B.
      • Postlethwait J.H.
      • Warman M.L.
      • Nakamura T.
      • Akiyama H.
      Wwp2 is essential for palatogenesis mediated by the interaction between Sox9 and mediator subunit 25.
      • Kawakami Y.
      • Tsuda M.
      • Takahashi S.
      • Taniguchi N.
      • Esteban C.R.
      • Zemmyo M.
      • Furumatsu T.
      • Lotz M.
      • Izpisua Belmonte J.C.
      • Asahara H.
      Transcriptional coactivator PGC-1alpha regulates chondrogenesis via association with Sox9.
      • Han Y.
      • Lefebvre V.
      L-Sox5 and Sox6 drive expression of the aggrecan gene in cartilage by securing binding of Sox9 to a far-upstream enhancer.
      Interestingly, pathogenic germline variants in COL2A1 have recently been associated with chordoma in a cohort of patients with European or Chinese ancestry.
      • Yepes S.
      • Shah N.N.
      • Bai J.
      • Koka H.
      • Li C.
      • Gui S.
      • McMaster M.L.
      • Xiao Y.
      • Jones K.
      • Wang M.
      • Vogt A.
      • Zhu B.
      • Zhu B.
      • Hutchinson A.
      • Yeager M.
      • Hicks B.
      • Carter B.
      • Freedman N.D.
      • Beane-Freeman L.
      • Chanock S.J.
      • Zhang Y.
      • Parry D.M.
      • Yang X.R.
      • Goldstein A.M.
      Rare germline variants in chordoma-related genes and chordoma susceptibility.
      Additional pathways that have been observed enriched in chordoma, such as NF-κB, can also serve to enhance chondrogenesis.
      • Ushita M.
      • Saito T.
      • Ikeda T.
      • Yano F.
      • Higashikawa A.
      • Ogata N.
      • Chung U.
      • Nakamura K.
      • Kawaguchi H.
      Transcriptional induction of SOX9 by NF-kappaB family member RelA in chondrogenic cells.
      ,
      • Caron M.M.
      • Emans P.J.
      • Surtel D.A.
      • Cremers A.
      • Voncken J.W.
      • Welting T.J.
      • van Rhijn L.W.
      Activation of NF-kappaB/p65 facilitates early chondrogenic differentiation during endochondral ossification.
      Regardless of the role brachyury plays in tumorigenesis or chordoma risk, as a transcription factor it remains a challenging drug target for cancer therapy.
      • Chen A.
      • Koehler A.N.
      Transcription factor inhibition: lessons learned and emerging targets.
      Our analysis presents a new list of potential therapeutic targets, the TGF-β pathway. In vitro shRNA knockdown experiments revealed the core chordoma TGF-β/SOX6/SOX9/TBXT network to be highly interdependent, with TGF-β potentially playing a central role. Further flow cytometry analysis revealed a dichotomous relationship between proliferation and apoptosis associated with knockdown of pathway members. Specifically, during the experimental timeline of about 1 week, TBXT knockdown is associated with vastly decreased proliferation and moderately increased apoptosis, whereas TGFΒ1 knockdown is associated with moderately decreased proliferation and vastly increased apoptosis. Although TBXT knockdown does lead to cell death after a significant delay, TGFΒ1 knockdown resulted in fewer surviving chordoma cells than TBXT knockdown, presenting a therapeutic alternative or supplement to targeting TBXT.
      Experiments with small-molecule inhibitors further pinpoint the hyperactivated areas within the TGF-β pathway. Our tests with chemical inhibitors of various members of the TGF-β pathway identified ALK1 and ALK2 inhibitors as especially potent in chordoma cells (Figure 4), with inhibitors of other TGF-β pathway receptors showing little differential effect in chordoma cells versus HEK293 controls (data not shown). Better therapeutic molecules should be further investigated within the TGF-β pathway. More important, TGF-β inhibition has been proposed as a therapeutic avenue in many other cancers, making the drug development tractable for a rare tumor, such as chordoma. In support of this, there are already multiple clinical trials underway for various inhibitors of the TGF-β pathway.
      • Ciardiello D.
      • Elez E.
      • Tabernero J.
      • Seoane J.
      Clinical development of therapies targeting TGFβ: current knowledge and future perspectives.
      The findings of this study provide a new window into the molecular mechanism of chordoma, which, until now, was primarily centered on brachyury. Current treatment options for patients with chordoma are severely limited, and additional therapeutic targets driven by unbiased data analysis are desperately needed. Our analysis opens a new avenue for chordoma research and therapy, implicating the TGF-β pathway as a promising target.

      Acknowledgments

      We thank Brian Seed and Ramnik J. Xavier for input, support, and discussions about this project; and the Chordoma Foundation for support.

      Supplemental Data

      • Supplemental Figure S1

        Genome browser view of RNA-sequencing profiling of chordoma and notochord samples as well as fetal spine (FE) and muscle (M1) control samples. Controls 1 and 2 are two control libraries of unknown origin provided by Nugen as a positive control of a successful RNA-sequencing run.

      • Supplemental Figure S2

        Gene enrichment in chordoma and notochord. A and B: Top 20 chordoma-enriched (A) and notochord-enriched (B) genes. Black bars indicate genes enriched in both chordoma and notochord. C: Top 10 scoring tissues or cell types for four genes highly enriched in chordoma and notochord. D: A selection of genes from the diagnostic signature panel was validated in chordoma cell lines (U-CH1, U-CH2, U-CH12, and CH22) using Western blot analysis. KCNK2 can be detected as either a high-molecular-weight glycosylated homodimer or a monomeric unit.

      • Supplemental Figure S3

        A functional analysis of genes enriched in chordoma. A: Cumulative P values of a gene enrichment analysis of seven chordoma samples. P values were computed per gene list using the hypergeometric test on the GeneGo pathway maps. B: Cumulative P values of gene enrichment analysis for 2158 tumors in the ExpO data set. P values were computed per gene list using the hypergeometric test on the GeneGo pathway maps. ECM, extracellular matrix; EMT, epithelial-mesenchymal transition; PI3K, phosphatidylinositol 3-kinase.

      • Supplemental Figure S4

        A protein interaction network of manually curated interactions linking chordoma signature genes (yellow), chordoma-enriched genes (red), and expressed genes (gray). Full network is shown (A), and a filtered network with only interactions present in manually curated pathway maps is shown (B).

      • Supplemental Figure S5

        A functional analysis of chordoma enriched/expressed genes, spanning Reactome, GeneGo, Gene Ontology, and Ingenuity pathway databases. Each edge in this network represents the most significant pathway membership in each of the databases. TGF-β, transforming growth factor-β.

      • Supplemental Figure S6

        A proposed molecular model of chordoma. SOX9 and SOX6 are suggested as the main transcription factors driving the production of extracellular matrix. The matrix activates the transforming growth factor (TGF)-β pathway through integrin signaling, which feeds back into SOX9 (orange lines). In addition, the activated NF-κB pathway is also feeding into SOX transcription factors and is potentially activated by MDM2 (green lines).

      • Supplemental Figure S7

        The same assay for annexin or markers of early apoptosis performed in Figure 3 was done on HEK293 cells stably expressing shRNA against either TGFB1 (shTGFB1) or green fluorescent protein (GFP; shGFP). Cell positivity percentages were calculated via cytometric analysis (A) and plotted in a bar plot (B) alongside the corresponding data for U-CH2 that was presented in Figure 3. FITC, fluorescein isothiocyanate.

      References

        • Scott D.
        • Pedlow F.
        • Hecht A.
        • Hornicek F.
        Chapter 11; tumors: primary benign and malignant extradural spine tumors.
        in: The Adult and Pediatric Spine. ed 3. Lippincott Williams & Wilkins, 2004: 191-246
        • McMaster M.L.
        • Goldstein A.M.
        • Bromley C.M.
        • Ishibe N.
        • Parry D.M.
        Chordoma: incidence and survival patterns in the United States, 1973-1995.
        Cancer Causes Control. 2001; 12: 1-11
        • DeLaney T.F.
        • Liebsch N.J.
        • Pedlow F.X.
        • Adams J.
        • Dean S.
        • Yeap B.Y.
        • McManus P.
        • Rosenberg A.E.
        • Nielsen G.P.
        • Harmon D.C.
        • Spiro I.J.
        • Raskin K.A.
        • Suit H.D.
        • Yoon S.S.
        • Hornicek F.J.
        Phase II study of high-dose photon/proton radiotherapy in the management of spine sarcomas.
        Int J Radiat Oncol Biol Phys. 2009; 74: 732-739
        • Hofheinz R.D.
        • Kubicka S.
        • Wollert J.
        • Arnold D.
        • Hochhaus A.
        Gefitinib in combination with 5-fluorouracil (5-FU)/folinic acid and irinotecan in patients with 5-FU/oxaliplatin-refractory colorectal cancer: a phase I/II study of the Arbeitsgemeinschaft fur Internistische Onkologie (AIO).
        Onkologie. 2006; 29: 563-567
        • Stacchiotti S.
        • Longhi A.
        • Ferraresi V.
        • Grignani G.
        • Comandone A.
        • Stupp R.
        • Bertuzzi A.
        • Tamborini E.
        • Pilotti S.
        • Messina A.
        • Spreafico C.
        • Gronchi A.
        • Amore P.
        • Vinaccia V.
        • Casali P.G.
        Phase II study of imatinib in advanced chordoma.
        J Clin Oncol. 2012; 30: 914-920
        • Meng T.
        • Jin J.
        • Jiang C.
        • Huang R.
        • Yin H.
        • Song D.
        • Cheng L.
        Molecular targeted therapy in the treatment of chordoma: a systematic review.
        Front Oncol. 2019; 9: 30
        • Picci P.
        • Vanel D.
        • Alberghini M.
        • Mirra J.M.
        • Errani C.
        • Staals E.L.
        • Mercuri M.
        Giant notochordal rests misdiagnosed and treated as chordomas: a retrospective clinical, radiological and histologic study of four cases.
        JCO. 2008; 26: 21503
        • Yamaguchi T.
        • Iwata J.
        • Sugihara S.
        • McCarthy E.F.
        • Karita M.
        • Murakami H.
        • Kawahara N.
        • Tsuchiya H.
        • Tomita K.
        Distinguishing benign notochordal cell tumors from vertebral chordoma.
        Skeletal Radiol. 2008; 37: 291-299
        • Arain A.
        • Hornicek F.J.
        • Schwab J.H.
        • Chebib I.
        • Damron T.A.
        Chordoma arising from benign multifocal notochordal tumors.
        Skeletal Radiol. 2017; 46: 1745-1752
        • Kispert A.
        • Koschorz B.
        • Herrmann B.G.
        The T protein encoded by Brachyury is a tissue-specific transcription factor.
        EMBO J. 1995; 14: 4763-4772
        • Presneau N.
        • Shalaby A.
        • Ye H.
        • Pillay N.
        • Halai D.
        • Idowu B.
        • Tirabosco R.
        • Whitwell D.
        • Jacques T.S.
        • Kindblom L.-G.
        • Brüderlein S.
        • Möller P.
        • Leithner A.
        • Liegl B.
        • Amary F.M.
        • Athanasou N.N.
        • Hogendoorn P.C.
        • Mertens F.
        • Szuhai K.
        • Flanagan A.M.
        Role of the transcription factor T (brachyury) in the pathogenesis of sporadic chordoma: a genetic and functional-based study.
        J Pathol. 2010; 223: 327-335
        • Yang X.R.
        • Ng D.
        • Alcorta D.A.
        • Liebsch N.J.
        • Sheridan E.
        • Li S.
        • Goldstein A.M.
        • Parry D.M.
        • Kelley M.J.
        T (brachyury) gene duplication confers major susceptibility to familial chordoma.
        Nat Genet. 2009; 41: 1176-1178
        • Pillay N.
        • Plagnol V.
        • Tarpey P.S.
        • Lobo S.B.
        • Presneau N.
        • Szuhai K.
        • Halai D.
        • Berisha F.
        • Cannon S.R.
        • Mead S.
        • Kasperaviciute D.
        • Palmen J.
        • Talmud P.J.
        • Kindblom L.G.
        • Amary M.F.
        • Tirabosco R.
        • Flanagan A.M.
        A common single-nucleotide variant in T is strongly associated with chordoma.
        Nat Genet. 2012; 44: 1185-1187
        • Chen A.
        • Koehler A.N.
        Transcription factor inhibition: lessons learned and emerging targets.
        Trends Mol Med. 2020; 26: 508-518
        • Henley M.J.
        • Koehler A.N.
        Advances in targeting “undruggable” transcription factors with small molecules.
        Nat Rev Drug Discov. 2021; 20: 669-688
        • Barrett T.
        • Suzek T.O.
        • Troup D.B.
        • Wilhite S.E.
        • Ngau W.C.
        • Ledoux P.
        • Rudnev D.
        • Lash A.E.
        • Fujibuchi W.
        • Edgar R.
        NCBI GEO: mining millions of expression profiles--database and tools.
        Nucleic Acids Res. 2005; 33: D562-D566
        • Parkinson H.
        • Kapushesky M.
        • Shojatalab M.
        • Abeygunawardena N.
        • Coulson R.
        • Farne A.
        • Holloway E.
        • Kolesnykov N.
        • Lilja P.
        • Lukk M.
        • Mani R.
        • Rayner T.
        • Sharma A.
        • William E.
        • Sarkans U.
        • Brazma A.
        ArrayExpress--a public database of microarray experiments and gene expression profiles.
        Nucleic Acids Res. 2007; 35: D747-D750
        • Dai M.
        • Wang P.
        • Boyd A.D.
        • Kostov G.
        • Athey B.
        • Jones E.G.
        • Bunney W.E.
        • Myers R.M.
        • Speed T.P.
        • Akil H.
        • Watson S.J.
        • Meng F.
        Evolving gene/transcript definitions significantly alter the interpretation of GeneChip data.
        Nucleic Acids Res. 2005; 33: e175
        • Wu J.
        • Irizarry R.
        • MacDonald J.
        • Gentry J.
        gcrma: Background Adjustment Using Sequence Information.
        2021
        • Gautier L.
        • Cope L.
        • Bolstad B.M.
        • Irizarry R.A.
        affy—analysis of Affymetrix GeneChip data at the probe level.
        Bioinformatics. 2004; 20: 307-315
        • Barretina J.
        • Caponigro G.
        • Stransky N.
        • Venkatesan K.
        • Margolin A.A.
        • Kim S.
        • et al.
        The Cancer Cell Line Encyclopedia enables predictive modelling of anticancer drug sensitivity.
        Nature. 2012; 483: 603-607
        • Benita Y.
        • Cao Z.
        • Giallourakis C.
        • Li C.
        • Gardet A.
        • Xavier R.J.
        Gene enrichment profiles reveal T-cell development, differentiation, and lineage-specific transcription factors including ZBTB25 as a novel NF-AT repressor.
        Blood. 2010; 115: 5376-5384
        • Smyth G.K.
        • Michaud J.
        • Scott H.S.
        Use of within-array replicate spots for assessing differential expression in microarray experiments.
        Bioinformatics. 2005; 21: 2067-2075
        • Langmead B.
        • Trapnell C.
        • Pop M.
        • Salzberg S.L.
        Ultrafast and memory-efficient alignment of short DNA sequences to the human genome.
        Genome Biol. 2009; 10: R25
        • Karolchik D.
        • Kuhn R.M.
        • Baertsch R.
        • Barber G.P.
        • Clawson H.
        • Diekhans M.
        • Giardine B.
        • Harte R.A.
        • Hinrichs A.S.
        • Hsu F.
        • Kober K.M.
        • Miller W.
        • Pedersen J.S.
        • Pohl A.
        • Raney B.J.
        • Rhead B.
        • Rosenbloom K.R.
        • Smith K.E.
        • Stanke M.
        • Thakkapallayil A.
        • Trumbower H.
        • Wang T.
        • Zweig A.S.
        • Haussler D.
        • Kent W.J.
        The UCSC genome browser database: 2008 update.
        Nucleic Acids Res. 2008; 36: D773-D779
        • Trapnell C.
        • Pachter L.
        • Salzberg S.L.
        TopHat: discovering splice junctions with RNA-Seq.
        Bioinformatics. 2009; 25: 1105-1111
        • Dobin A.
        • Davis C.A.
        • Schlesinger F.
        • Drenkow J.
        • Zaleski C.
        • Jha S.
        • Batut P.
        • Chaisson M.
        • Gingeras T.R.
        STAR: ultrafast universal RNA-seq aligner.
        Bioinformatics. 2013; 29: 15-21
        • Vastrik I.
        • D'Eustachio P.
        • Schmidt E.
        • Joshi-Tope G.
        • Gopinath G.
        • Croft D.
        • de Bono B.
        • Gillespie M.
        • Jassal B.
        • Lewis S.
        • Matthews L.
        • Wu G.
        • Birney E.
        • Stein L.
        Reactome: a knowledge base of biologic pathways and processes.
        Genome Biol. 2007; 8: R39
        • Ashburner M.
        • Ball C.A.
        • Blake J.A.
        • Botstein D.
        • Butler H.
        • Cherry J.M.
        • Davis A.P.
        • Dolinski K.
        • Dwight S.S.
        • Eppig J.T.
        • Harris M.A.
        • Hill D.P.
        • Issel-Tarver L.
        • Kasarskis A.
        • Lewis S.
        • Matese J.C.
        • Richardson J.E.
        • Ringwald M.
        • Rubin G.M.
        • Sherlock G.
        The Gene Ontology Consortium: Gene ontology: tool for the unification of biology..
        Nat Genet. 2000; 25: 25-29
        • Wheeler D.L.
        • Barrett T.
        • Benson D.A.
        • Bryant S.H.
        • Canese K.
        • Church D.M.
        • DiCuccio M.
        • Edgar R.
        • Federhen S.
        • Helmberg W.
        • Kenton D.L.
        • Khovayko O.
        • Lipman D.J.
        • Madden T.L.
        • Maglott D.R.
        • Ostell J.
        • JU Pontius
        • Pruitt K.D.
        • Schuler G.D.
        • Schriml L.M.
        • Sequeira E.
        • Sherry S.T.
        • Sirotkin K.
        • Starchenko G.
        • Suzek T.O.
        • Tatusov R.
        • Tatusova T.A.
        • Wagner L.
        • Yaschenko E.
        Database resources of the National Center for Biotechnology Information.
        Nucleic Acids Res. 2005; 33: D39-D45
        • Csardi G.
        • Nepusz T.
        The igraph software package for complex network research. InterJournal.
        2006 (Complex Systems:1695)
        • Ritz C.
        • Baty F.
        • Streibig J.C.
        • Gerhard D.
        Dose-response analysis using R.
        PLoS One. 2015; 10e0146021
        • Mohedas A.H.
        • Xing X.
        • Armstrong K.A.
        • Bullock A.N.
        • Cuny G.D.
        • Yu P.B.
        Development of an ALK2-biased BMP type I receptor kinase inhibitor.
        ACS Chem Biol. 2013; 8: 1291-1302
        • Jäger D.
        • Barth T.F.E.
        • Brüderlein S.
        • Scheuerle A.
        • Rinner B.
        • von Witzleben A.
        • Lechel A.
        • Meyer P.
        • Mayer-Steinacker R.
        • Baer Avon
        • Schultheiss M.
        • Wirtz C.R.
        • Möller P.
        • Mellert K.
        HOXA7, HOXA9, and HOXA10 are differentially expressed in clival and sacral chordomas.
        Sci Rep. 2017; 7: 2032
        • Lohberger B.
        • Scheipl S.
        • Heitzer E.
        • Quehenberger F.
        • de Jong D.
        • Szuhai K.
        • Liegl-Atzwanger B.
        • Rinner B.
        Higher cMET dependence of sacral compared to clival chordoma cells: contributing to a better understanding of cMET in chordoma.
        Sci Rep. 2021; 11: 12466
        • Bakker S.H.
        • Jacobs W.C.H.
        • Pondaag W.
        • Gelderblom H.
        • Nout R.A.
        • Dijkstra P.D.S.
        • Peul W.C.
        • Vleggeert-Lankamp C.L.A.
        Chordoma: a systematic review of the epidemiology and clinical prognostic factors predicting progression-free and overall survival.
        Eur Spine J. 2018; 27: 3043-3058
        • Baelde H.J.
        • Cleton-Jansen A.M.
        • van Beerendonk H.
        • Namba M.
        • Bovee J.V.
        • Hogendoorn P.C.
        High quality RNA isolation from tumours with low cellularity and high extracellular matrix component for cDNA microarrays: application to chondrosarcoma.
        J Clin Pathol. 2001; 54: 778-782
        • Scheil S.
        • Bruderlein S.
        • Liehr T.
        • Starke H.
        • Herms J.
        • Schulte M.
        • Moller P.
        Genome-wide analysis of sixteen chordomas by comparative genomic hybridization and cytogenetics of the first human chordoma cell line, U-CH1.
        Genes Chromosomes Cancer. 2001; 32: 203-211
        • Fernando R.I.
        • Litzinger M.
        • Trono P.
        • Hamilton D.H.
        • Schlom J.
        • Palena C.
        The T-box transcription factor Brachyury promotes epithelial-mesenchymal transition in human tumor cells.
        J Clin Invest. 2010; 120: 533-544
        • Roselli M.
        • Fernando R.I.
        • Guadagni F.
        • Spila A.
        • Alessandroni J.
        • Palmirotta R.
        • Costarelli L.
        • Litzinger M.
        • Hamilton D.
        • Huang B.
        • Tucker J.
        • Tsang K.Y.
        • Schlom J.
        • Palena C.
        Brachyury, a driver of the epithelial-mesenchymal transition, is overexpressed in human lung tumors: an opportunity for novel interventions against lung cancer.
        Clin Cancer Res. 2012; 18: 3868-3879
        • Du R.
        • Wu S.
        • Lv X.
        • Fang H.
        • Wu S.
        • Kang J.
        Overexpression of brachyury contributes to tumor metastasis by inducing epithelial-mesenchymal transition in hepatocellular carcinoma.
        J Exp Clin Cancer Res. 2014; 33: 105
        • Kang H.S.
        • Ahn J.M.
        • Kang Y.
        Chondrogenic tumors.
        in: Kang H.S. Ahn J.M. Kang Y. Oncologic Imaging: Bone Tumors [Internet]. Springer, Singapore2017: 77-129https://doi.org/10.1007/978-981-287-703-1_3
        • Huh Y.H.
        • Ryu J.-H.
        • Shin S.
        • Lee D.-U.
        • Yang S.
        • Oh K.-S.
        • Chun C.-H.
        • Choi J.-K.
        • Song W.K.
        • Chun J.-S.
        Esophageal cancer related gene 4 (ECRG4) is a marker of articular chondrocyte differentiation and cartilage destruction.
        Gene. 2009; 448: 7-15
        • Matsumoto Y.
        • Sato S.
        • Maeda T.
        • Kishino M.
        • Toyosawa S.
        • Usami Y.
        • Iwai S.
        • Nakazawa M.
        • Yura Y.
        • Ogawa Y.
        Transcription factors related to chondrogenesis in pleomorphic adenoma of the salivary gland: a mechanism of mesenchymal tissue formation.
        Lab Invest. 2016; 96: 16-24
        • Robert A.W.
        • Marcon B.H.
        • Dallagiovanna B.
        • Shigunov P.
        Adipogenesis, osteogenesis, and chondrogenesis of human mesenchymal stem/stromal cells: a comparative transcriptome approach.
        Front Cell Dev Biol. 2020; 8: 561
        • Murakami S.
        • Lefebvre V.
        • de Crombrugghe B.
        Potent inhibition of the master chondrogenic factor Sox9 gene by interleukin-1 and tumor necrosis factor-α.
        J Biol Chem. 2000; 275: 3687-3692
        • Desgrosellier J.S.
        • Cheresh D.A.
        Integrins in cancer: biological implications and therapeutic opportunities.
        Nat Rev Cancer. 2010; 10: 9-22
        • Camper L.
        • Hellman U.
        • Lundgren-Akerlund E.
        Isolation, cloning, and sequence analysis of the integrin subunit alpha10, a beta1-associated collagen binding integrin expressed on chondrocytes.
        J Biol Chem. 1998; 273: 20383-20389
        • Hynes R.O.
        Integrins: bidirectional, allosteric signaling machines.
        Cell. 2002; 110: 673-687
        • Munger J.S.
        • Huang X.
        • Kawakatsu H.
        • Griffiths M.J.
        • Dalton S.L.
        • Wu J.
        • Pittet J.F.
        • Kaminski N.
        • Garat C.
        • Matthay M.A.
        • Rifkin D.B.
        • Sheppard D.
        The integrin alpha v beta 6 binds and activates latent TGF beta 1: a mechanism for regulating pulmonary inflammation and fibrosis.
        Cell. 1999; 96: 319-328
        • Ludbrook S.B.
        • Barry S.T.
        • Delves C.J.
        • Horgan C.M.
        The integrin alphavbeta3 is a receptor for the latency-associated peptides of transforming growth factors beta1 and beta3.
        Biochem J. 2003; 369: 311-318
        • Wendt M.K.
        • Tian M.
        • Schiemann W.P.
        Deconstructing the mechanisms and consequences of TGF-beta-induced EMT during cancer progression.
        Cell Tissue Res. 2012; 347: 85-101
        • Bachman K.E.
        • Park B.H.
        Duel nature of TGF-beta signaling: tumor suppressor vs. tumor promoter.
        Curr Opin Oncol. 2005; 17: 49-54
        • Mehlhorn A.T.
        • Schmal H.
        • Kaiser S.
        • Lepski G.
        • Finkenzeller G.
        • Stark G.B.
        • Sudkamp N.P.
        Mesenchymal stem cells maintain TGF-beta-mediated chondrogenic phenotype in alginate bead culture.
        Tissue Eng. 2006; 12: 1393-1403
        • Le L.P.
        • Nielsen G.P.
        • Rosenberg A.E.
        • Thomas D.
        • Batten J.M.
        • Deshpande V.
        • Schwab J.
        • Duan Z.
        • Xavier R.J.
        • Hornicek F.J.
        • Iafrate A.J.
        Recurrent chromosomal copy number alterations in sporadic chordomas.
        PLoS One. 2011; 6: e18846
        • Duan W.
        • Zhang B.
        • Li X.
        • Chen W.
        • Jia S.
        • Xin Z.
        • Jian Q.
        • Jian F.
        • Chou D.
        • Chen Z.
        Single-cell transcriptome profiling reveals intra-tumoral heterogeneity in human chordomas.
        Cancer Immunol Immunother. 2022; 71: 2185-2195
        • Ma J.
        • Tian K.
        • Wang L.
        • Wang K.
        • Du J.
        • Li D.
        • Wu Z.
        • Zhang J.
        High expression of TGF-β1 predicting tumor progression in skull base chordomas.
        World Neurosurg. 2019; 131: e265-e270
        • Wang L.
        • Guan X.
        • Hu Q.
        • Wu Z.
        • Chen W.
        • Song L.
        • Wang K.
        • Tian K.
        • Cao C.
        • Zhang D.
        • Ma J.
        • Tong X.
        • Zhang B.
        • Zhang J.
        • Zeng C.
        TGFB3 downregulation causing chordomagenesis and its tumor suppression role maintained by Smad7.
        Carcinogenesis. 2021; 42: 913-923
        • Lorda-Diez C.I.
        • Montero J.A.
        • Martinez-Cue C.
        • Garcia-Porrero J.A.
        • Hurle J.M.
        Transforming growth factors beta coordinate cartilage and tendon differentiation in the developing limb mesenchyme.
        J Biol Chem. 2009; 284: 29988-29996
        • Wu M.Y.
        • Hill C.S.
        Tgf-beta superfamily signaling in embryonic development and homeostasis.
        Dev Cell. 2009; 16: 329-343
        • Furumatsu T.
        • Tsuda M.
        • Taniguchi N.
        • Tajima Y.
        • Asahara H.
        Smad3 induces chondrogenesis through the activation of SOX9 via CREB-binding protein/p300 recruitment.
        J Biol Chem. 2005; 280: 8343-8350
        • Oh C.D.
        • Maity S.N.
        • Lu J.F.
        • Zhang J.
        • Liang S.
        • Coustry F.
        • de Crombrugghe B.
        • Yasuda H.
        Identification of SOX9 interaction sites in the genome of chondrocytes.
        PLoS One. 2010; 5: e10113
        • Nakamura Y.
        • Yamamoto K.
        • He X.
        • Otsuki B.
        • Kim Y.
        • Murao H.
        • Soeda T.
        • Tsumaki N.
        • Deng J.M.
        • Zhang Z.
        • Behringer R.R.
        • Crombrugghe B.
        • Postlethwait J.H.
        • Warman M.L.
        • Nakamura T.
        • Akiyama H.
        Wwp2 is essential for palatogenesis mediated by the interaction between Sox9 and mediator subunit 25.
        Nat Commun. 2011; 2: 251
        • Kawakami Y.
        • Tsuda M.
        • Takahashi S.
        • Taniguchi N.
        • Esteban C.R.
        • Zemmyo M.
        • Furumatsu T.
        • Lotz M.
        • Izpisua Belmonte J.C.
        • Asahara H.
        Transcriptional coactivator PGC-1alpha regulates chondrogenesis via association with Sox9.
        Proc Natl Acad Sci U S A. 2005; 102: 2414-2419
        • Han Y.
        • Lefebvre V.
        L-Sox5 and Sox6 drive expression of the aggrecan gene in cartilage by securing binding of Sox9 to a far-upstream enhancer.
        Mol Cell Biol. 2008; 28: 4999-5013
        • Yepes S.
        • Shah N.N.
        • Bai J.
        • Koka H.
        • Li C.
        • Gui S.
        • McMaster M.L.
        • Xiao Y.
        • Jones K.
        • Wang M.
        • Vogt A.
        • Zhu B.
        • Zhu B.
        • Hutchinson A.
        • Yeager M.
        • Hicks B.
        • Carter B.
        • Freedman N.D.
        • Beane-Freeman L.
        • Chanock S.J.
        • Zhang Y.
        • Parry D.M.
        • Yang X.R.
        • Goldstein A.M.
        Rare germline variants in chordoma-related genes and chordoma susceptibility.
        Cancers (Basel). 2021; 13: 2704
        • Ushita M.
        • Saito T.
        • Ikeda T.
        • Yano F.
        • Higashikawa A.
        • Ogata N.
        • Chung U.
        • Nakamura K.
        • Kawaguchi H.
        Transcriptional induction of SOX9 by NF-kappaB family member RelA in chondrogenic cells.
        Osteoarthr Cartil. 2009; 17: 1065-1075
        • Caron M.M.
        • Emans P.J.
        • Surtel D.A.
        • Cremers A.
        • Voncken J.W.
        • Welting T.J.
        • van Rhijn L.W.
        Activation of NF-kappaB/p65 facilitates early chondrogenic differentiation during endochondral ossification.
        PLoS One. 2012; 7: e33467
        • Ciardiello D.
        • Elez E.
        • Tabernero J.
        • Seoane J.
        Clinical development of therapies targeting TGFβ: current knowledge and future perspectives.
        Ann Oncol. 2020; 31: 1336-1349