Next Article in Journal
Biomarkers in Hepatobiliary Cancers: What Is Useful in Clinical Practice?
Next Article in Special Issue
Familial Risk and Heritability of Hematologic Malignancies in the Nordic Twin Study of Cancer
Previous Article in Journal
CALD1 Modulates Gliomas Progression via Facilitating Tumor Angiogenesis
Previous Article in Special Issue
A Rare Variant in ERF (rs144812092) Predisposes to Prostate and Bladder Cancers in an Extended Pedigree
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Rare Germline Variants in Chordoma-Related Genes and Chordoma Susceptibility

by
Sally Yepes
1,*,
Nirav N. Shah
1,
Jiwei Bai
2,
Hela Koka
1,
Chuzhong Li
2,
Songbai Gui
2,
Mary Lou McMaster
1,
Yanzi Xiao
1,
Kristine Jones
1,3,
Mingyi Wang
1,3,
Aurelie Vogt
1,3,
Bin Zhu
1,
Bin Zhu
1,3,
Amy Hutchinson
1,3,
Meredith Yeager
1,3,
Belynda Hicks
1,3,
Brian Carter
4,
Neal D. Freedman
1,
Laura Beane-Freeman
1,
Stephen J. Chanock
1,
Yazhuo Zhang
2,
Dilys M. Parry
1,
Xiaohong R. Yang
1,† and
Alisa M. Goldstein
1,*,†
add Show full author list remove Hide full author list
1
Division of Cancer Epidemiology and Genetics, National Cancer Institute, National Institutes of Health, Bethesda, MD 20892, USA
2
Beijing Tiantan Hospital, Beijing 100070, China
3
Cancer Genomics Research Laboratory, Leidos Biomedical Research, Frederick National Laboratory for Cancer Research, Frederick, MD 21702-1201, USA
4
American Cancer Society, Inc, Atlanta, GA 30303, USA
*
Authors to whom correspondence should be addressed.
Co-senior authors.
Cancers 2021, 13(11), 2704; https://doi.org/10.3390/cancers13112704
Submission received: 27 April 2021 / Revised: 20 May 2021 / Accepted: 24 May 2021 / Published: 30 May 2021
(This article belongs to the Special Issue Familial Cancer and Germline Genetics)

Abstract

:

Simple Summary

Chordoma is an extremely rare bone cancer that has not been fully characterized and few risk factors have been identified, highlighting the need for improving our understanding of the disease biology. Our study aims to identify chordoma susceptibility genes by investigating 265 genes involved in chordoma-related signaling pathways and other biological processes on germline DNA of 138 chordoma patients of European ancestry compared to internal control datasets and general population databases. Results were intersected with whole genome sequencing data from 80 skull-base chordoma patients of Chinese ancestry. Several rare loss-of-function and predicted deleterious missense variants were enriched in chordoma cases in both datasets, suggesting a complex model of pathways potentially involved in chordoma development and susceptibility, warranting further investigation in larger studies.

Abstract

Background: Chordoma is a rare bone cancer with an unknown etiology. TBXT is the only chordoma susceptibility gene identified to date; germline single nucleotide variants and copy number variants in TBXT have been associated with chordoma susceptibility in familial and sporadic chordoma. However, the genetic susceptibility of chordoma remains largely unknown. In this study, we investigated rare germline genetic variants in genes involved in TBXT/chordoma-related signaling pathways and other biological processes in chordoma patients from North America and China. Methods: We identified variants that were very rare in general population and internal control datasets and showed evidence for pathogenicity in 265 genes in a whole exome sequencing (WES) dataset of 138 chordoma patients of European ancestry and in a whole genome sequencing (WGS) dataset of 80 Chinese patients with skull base chordoma. Results: Rare and likely pathogenic variants were identified in 32 of 138 European ancestry patients (23%), including genes that are part of notochord development, PI3K/AKT/mTOR, Sonic Hedgehog, SWI/SNF complex and mesoderm development pathways. Rare pathogenic variants in COL2A1, EXT1, PDK1, LRP2, TBXT and TSC2, among others, were also observed in Chinese patients. Conclusion: We identified several rare loss-of-function and predicted deleterious missense variants in germline DNA from patients with chordoma, which may influence chordoma predisposition and reflect a complex susceptibility, warranting further investigation in large studies.

1. Introduction

Chordoma is a rare, malignant bone tumor that occurs in the axial skeleton at cranial, spinal and sacral sites. Chordoma tumors are considered slow growing; however, local recurrences are frequent and treatment options are limited, particularly for those with advanced disease, highlighting the need for improving our knowledge of the disease biology and the discovery of novel druggable targets [1,2,3]. The clinical progression of skull base chordoma is highly variable [4] and there are no validated clinical or molecular prognostic markers.
Chordoma is hypothesized to originate from remnants of the notochord, which is a mesodermal structure in the embryo and signal tissues for organization and differentiation. Genes implicated in chordoma formation include the T-brachyury gene (TBXT), which is required for differentiation of the notochord and formation of mesoderm during posterior development [5].
TBXT is the most relevant susceptibility gene identified in chordoma, with germline TBXT duplication conferring susceptibility to familial chordoma [6] (Yang XR et al. 2009) and single-nucleotide polymorphisms (SNPs) in TBXT associated with chordoma risk in both sporadic and familial chordoma [7,8]. The role of TBXT in disease pathogenesis has also been shown to be critical: TBXT expression is considered as a diagnostic marker for chordoma [9,10] and copy number gains of TBXT have been identified in chordoma tumors [11,12]. Further, chordoma growth in cell models could be inhibited by TBXT silencing [12,13,14].
Despite some progress, etiologic factors and genetic susceptibility of chordoma are not well understood and have not been extensively studied. The goal of this study was to identify and characterize rare and pathogenic germline variants in genes of interest for chordoma pathogenesis through next generation sequencing of 138 chordoma cases of European ancestry. Genes of interest from the initial evaluation were also examined in a set of 80 skull-base chordoma cases from an Asian population for replication. Due to the critical relevance of TBXT in the biological mechanisms of chordoma, we investigated germline variants in genes related to TBXT, as well as other genes involved in mesoderm and notochord development. We also evaluated genes that have been associated with chordoma tumor development, including several members of the SWI/SNF complex, PI3K/AKT/mTOR and Sonic Hedgehog pathways. Using this approach, we identified several potential chordoma susceptibility genes that warrant investigation in future studies.

2. Materials and Methods

2.1. Study Populations

The current study included 138 patients of European ancestry from the United States and Canada, processed using whole exome sequencing (WES). All diagnoses of chordoma were confirmed by reviewing pathologic slides or reports, medical records or death certificates.
We used WES data from 598 healthy Caucasian cancer-free unrelated individuals, from the Prostate, Lung, Colorectal and Ovarian Cancer Screening Trial (PLCO) [15] and Cancer Prevention Study (CPS) as reference/controls for comparison with the European ancestry patient samples and for conducting rare variant burden tests. These controls were sequenced and analyzed using the same sequencing platform and ensemble variant calling pipeline used for the European ancestry chordoma cases.
We also analyzed germline whole genome sequencing (WGS) data from 80 Chinese skull-base chordoma patients to evaluate candidate genes identified from our main analysis of the European ancestry chordoma patients. In the Chinese patient cohort, all patients were diagnosed with skull base chordoma and underwent endoscopic endonasal surgeries at the Neurosurgery Department of Beijing Tiantan Hospital, Capital Medical University, between October 2010 and November 2017. All methods were performed in accordance with the relevant guidelines and regulations. This research has been approved by the National Cancer Institute (NCI) Institutional Review Board (IRB) (Protocols: 10CN188, 04/08/2020; 78C-0039, 01/08/2021) and the ethics committee of the Beijing Tiantan Hospital (IRB code: 2009-47, 20 December 2009).

2.2. Library Construction, Sequencing and Bioinformatics Analysis

For the European ancestry and Chinese cases, germline genomic DNA was extracted from saliva and peripheral blood. WES was performed at the Cancer Genomics Research Laboratory, National Cancer Institute (CGR, NCI). Briefly, SeqCAP EZ Human Exome Library v3.0 (Roche NimbleGen, Madison, WI, USA) was utilized for exome sequence capture. Exome sequencing was performed to a sufficient depth to achieve a minimum coverage of 15 reads in at least 80% of the coding sequence from the UCSC hg19 transcripts database.
After the exclusion of reads containing adapter contamination and low-quality nucleotides, the data were mapped to the reference human genome (UCSC hg19) using the Burrows–Wheeler Aligner software [16]. SAMtools [17], Picard and GATK [18] were used to sort aligned files and perform base quality recalibration, duplicate reads removal and local realignment to generate final BAM files for mutation calling.
Three variant callers, GATK HaplotyperCaller, UnifiedGenotyper and FreeBayes [19], were used to call germline variants. We included all target regions, as well as a 250 bp flanking region on each side. An ensemble variant calling pipeline was then implemented to integrate the analysis results from the above mentioned three callers. Subsequently, the ensemble variant calling pipeline that applies a support-vector machine (SVM) learning algorithm was used to identify an optimal decision boundary based on the variant calling results out of the multiple variant callers to produce a more balanced decision between false positives and true positives.
For the Chinese cases, WGS was carried out by the Novogene Corporation (Beijing, China) on the Illumina HiSeq X platform with an average depth of 41X. Variant calling, QC and filtering steps were conducted by CGR, NCI, using similar approaches as those used for WES of the European ancestry chordoma samples.
Gene prioritization: We assembled a set of 265 genes chosen from the published literature for their role in chordoma physiopathology. The following functional categories were included: TBXT super-enhancer associated genes and chordoma essentiality genes, which are essential for the proliferation and survival of cancer cells in CRISPR-CAS9 screens, notochord related genes, key genes in mesoderm commitment pathway, SWI/SNF complex, PI3K/Akt/mTOR pathway and Sonic Hedgehog pathway. Variants were filtered based on quality control measures, frequency of minor allele in populations and pathogenicity criteria.
Variants quality control: Filtering of WES variants was based on the following criteria: Variants flagged with our pipeline quality control metric (CScorefilter), read depth <10, ABHet <0.2 or >0.8, or called by only one of the three variant callers used were excluded. Only non-synonymous variants, including frameshift, stop/gain, inframe deletion or insertion, missense and splicing site variants, were studied.
Frequency and specificity: Variants were excluded if both of the following criteria were met: (1) they had a minor allele frequency (MAF) of >0.001 in any of the consulted population databases (the 1000 Genomes Project, Exome Sequencing Project (ESP6500) and Exome Aggregation Consortium (ExAC) in European population); (2) they were observed in >2 families from an in-house database (CGR, NCI) of ~2000 exomes in ~1000 cancer-prone families, excluding chordoma cancer families.
Pathogenicity evidence and in silico analysis: Variants passing quality control and frequency filters in both datasets were evaluated for pathogenicity. Variants were studied further if classified as one of the following: (1) High impact variants (frameshift indels, stop gain/loss, or known splice sites). (2) Missense variants with evidence of pathogenicity based on 3 in silico predictions (Meta Likelihood ratio: D, METASVM: D and CADD: ≥ 20). The first two in silico algorithms are ensemble prediction scores that incorporate results from nine algorithms (SIFT, PolyPhen-2, GERP ++, Mutation Taster, Mutation Assessor, FATHMM, LRT, SiPhy and PhyloP) and allele frequency [20]. (3) Variants classified as pathogenic (P) or likely pathogenic (LP) by ClinVar [21]. (4) Disease-causing mutation (DM) by HGMD [22]. Details of the criteria for classification of pathogenicity are shown in Supplementary Table S1 [23].
Similar filtering as described for the European ancestry dataset was performed in the Chinese dataset processed with WGS. For this latter dataset, only protein-coding genes, for which potentially pathogenic variants were observed in the European ancestry dataset, were considered. Variants were excluded based on MAF of >0.001 in any of the databases in any population, including East Asian. Pathogenicity evidence and in silico analysis were performed as described above for the European ancestry dataset.

2.3. Rare Variant Burden Test

We conducted rare variant burden tests for genes with rare and potentially pathogenic variants identified from the European ancestry cohort in 138 chordoma patients and 598 non-cancer control samples from the PLCO and CPS-II studies with European ancestry and processed at CGR, NCI. We used SKAT-O statistics [24], which is a linear combination of the burden test (aimed to test effect size of variants with the same direction in cases and controls) and variance component test (designed to test effect size of variants with different directions in cases and controls).

3. Results

The main analysis included 138 European ancestry patients; the average age at diagnosis was 46.9 years (range 7–78) and the ratio of male to female patients was 0.7. The vast majority of patients (97%) had classic chordoma histology. The chordoma site distribution was 55.7% skull base, 23% spinal and 20.3% sacral. The Chinese cases were all skull base chordoma, with a mean age of 44.7 (7–79) years, the majority being male (62.5%) and having classic chordoma histology (80%).
We assembled a set of 265 genes chosen from the published literature for their role in chordoma physiopathology, including TBXT super-enhancer associated genes and chordoma essentiality genes in CRISPR-CAS9 screens (26), notochord related genes (21), mesoderm commitment pathway (166), SWI/SNF complex (18), PI3K/Akt/mTOR pathway (33) and Sonic Hedgehog pathway (28). Some of these genes are present in more than one category (Supplementary Table S2). Figure 1 shows the stepwise pipeline used for selecting, prioritizing and filtering the genes and variants for the evaluation.
We identified 34 potentially pathogenic variants (7 loss of function and 27 missense variants) in 2 notochord development genes, 4 Sonic Hedgehog pathway genes, 16 mesoderm commitment pathway genes, 4 PI3K/AKT/mTOR pathway genes, 2 SWI/SNF complex pathway genes, 2 TBXT super-enhancer genes and 1 driver gene (Table 1). These variants were identified in 32 of 138 European ancestry patients (23%). Table 1 and Table 2 show patients’ characteristics, the type, location and frequencies of these variants, evidence for pathogenicity and the involved pathways/biologic processes. Supplementary Tables S3 and S4 show details of the variants identified in the European ancestry and Chinese datasets, respectively.
Several variants were classified as DM/P by HGMD/ClinVar, such as GDF3 (rs146973734), AKT1 (rs397514644) and EXT2 (rs138495222). Many of the variants were either extremely rare in the general population or absent from the internal control samples and population databases such as variants identified in TBXT, ATP6V1B2 and COL2A1.
Burden test: To identify genes with a higher genetic burden of rare variants, we conducted a rare variant burden test of the 265 genes by comparing cases to controls of the same ancestry and sequenced using similar analytics and pipelines. Burden tests were performed in several different ways: (1) analyzing all rare variants in a targeted gene; (2) examining only rare pathogenic variants in a targeted gene; (3) examining all genes in a process or pathway. The small number of cases carrying variants in the same genes limited the power to evaluate statistical significance for individual genes. As a result, none of the examined genes showed statistical significance after correction for multiple testing (Supplementary Table S5). However, fourteen rare pathogenic variants were found only in chordoma cases and were absent from all 598 controls (variants in these genes: AKT1, ATP6V1B2, DEPTOR, FOXA2, HHIP, NFE2L2, PAX6, PDK1, PRKACA, PTCH1, SMARCB1, SOX21, SOX9 and TBXT). Evaluation of the biologic processes or pathways showed suggestive evidence for association only for the notochord-related genes (SKATO P = 0.06; Burden P = 0.03).
We investigated the 31 genes (see Table 1) identified in the main analysis in an independent dataset of 80 Chinese chordoma cases. After applying the same filtering criteria as for the main analysis (see methods), we identified rare pathogenic variants in 11 genes in 18 of 80 (22.5%) Chinese patients. Most variants were very rare, with MAFs ranging between 1.74 E–04 and zero in all control databases (Table 2). Although most genes/variants were observed in only a single Chinese chordoma case, multiple potentially pathogenic variants were observed in COL2A1, LRP2, TCF7L1 and TSC2. In particular, the two variants in TSC2 were observed in two (g. 2134572) or three (g. 2121610) different patients, respectively. Further, the TSC2 variant observed in three Chinese chordoma cases was classified as DM by HGMD (Table 2).

4. Discussion

In this study, we investigated the role of rare germline variants in genes involved in chordoma related processes/pathways in sporadic chordoma patients from a European ancestry and a Chinese population. We identified a number of rare loss-of-function and predicted deleterious missense variants that were enriched in chordoma cases, suggesting some of these genes may contribute to chordoma susceptibility.
Genes with rare germline variants enriched in chordoma cases were involved in signaling pathways and physiopathological processes that are affected in chordoma at the somatic level (e.g., notochord development, PI3K/AKT/mTOR and SWI/SNF pathways), which underscores the importance of pathway/process-level analysis of germline alterations to identify potentially novel genes. Increasing evidence suggests that specific germline variants might determine which somatic events and mutations are generated and selected in cancer cells during tumorigenesis, such as mutational signatures, allele-specific copy number changes, or altered signaling pathways [25].
Overall, we report that ~23% of chordoma cases from two independent populations carried rare germline variants that are potentially pathogenic in selected genes. Eleven genes were shared by the two datasets, including COL2A1 (notochord), EXT1 (mesoderm development), PDK1 (PI3K/Akt/mTOR) and LRP2 (Sonic Hedgehog) and thus worthy of additional studies.
Some of the genes identified in this study have previously demonstrated important biological functions relevant to chordoma development. For example, a recent study found that TBXT is associated with a 1.5 Mb region containing “super-enhancers” and is the most highly expressed super-enhancer-associated transcription factor [26]. ATP6V1B2 was among the genes that were considered as both super-enhancer-associated and essential for chordoma cell viability. sgRNA-mediated TBXT repression experiments have been shown to reduce the expression of this gene and the direct binding of TBXT at the ATP6V1B2 locus has been demonstrated [26]. COL2A1 was also shown to be down-regulated by TBXT expression in a clival chordoma cell line (UM-Chor1) [26]. Interestingly, Col2a1-null mice lack intervertebral discs and cannot remove the notochord [27], suggesting the potential importance of this gene in chordoma development. We also identified rare variants in SHH, a member of the Sonic Hedgehog pathway, which is secreted by fetal notochord and participates in vertebrate patterning of the neural tube during embryonic development via binding to its ligand PTCH1. Interestingly, Shh-expressing notochord cells remain in the vertebral column after embryonic development ends, resembling notochord remnants [28] and suggesting that it may cause transformation of this remnant tissue.
The SMARCB1 gene is a critical component of the SWI/SNF chromatin-remodeling complex which antagonizes the histone methyltransferase EZH2, a molecule that is being targeted by several inhibitors in clinical trials [29]. Inactivation of SMARCB1 has been described as a critical event in various tumors, including malignant rhabdoid tumors and epithelioid sarcoma [30]. Loss of SMARCB1 expression in chordoma has also been identified in poorly differentiated chordomas [31]. Recently, other components of SWI/SNF complex have also been implicated in chordoma pathogenesis, such as PBRM1 as a potential driver gene for chordoma [11].
PI3K/Akt/mTOR signaling in the context of chordoma was first suggested after numerous reports of chordoma tumors in individuals with tuberous sclerosis complex (TSC). Chordoma also occurs in association with TSC, an autosomal dominant neurocutaneous syndrome characterized by abnormal tissue growth in multiple organ systems caused by inactivating germline mutations in either TSC1 or TSC2 [32]. None of the patients in our study with TSC variants have clinical manifestations of TSC. Nevertheless, mutations in PI3K/AKT/mTOR genes have been reported in chordoma tumors with potential therapeutic relevance [11].
The strengths of our study include the careful and comprehensive literature review to select chordoma candidate genes, the evaluation of germline rare variants in two independent chordoma sequencing datasets and the inclusion of controls without cancer from the same population ancestry as our European ancestry chordoma cases that were processed using the same platforms and bioinformatics pipeline. However, we cannot rule out the possibility that other genes, including cancer-predisposing genes, may also play a role in susceptibility.
The main limitation of our study is the relatively small number of chordoma cases in each dataset to statistically test case-control differences at an individual gene level, with most variants found only in single chordoma cases. In addition, we were unable to assess family history for most cases. There was also limited representation of histological subtypes for statistical evaluation of association with the variants identified.
In this study, approximately 23% of patients with chordoma from two different population had rare pathogenic/likely pathogenic variants in various biological processes, suggesting a complex model of pathways potentially important for susceptibility and development, including notochord development, PI3K/AKT/mTOR, Sonic Hedgehog and SWI/SNF complex. Future large studies, particularly at the consortium level, are needed to follow up on our findings for a better understanding of genetic susceptibility of this extremely rare cancer.

5. Conclusions

Our study searched for germline variants in signaling pathways and other biological processes previously identified in the disease’s pathogenesis in patients from two independent populations. We identified rare loss-of-function and predicted deleterious missense variants enriched in chordoma cases, suggesting a complex model of pathways potentially involved in chordoma development and susceptibility. Further evaluation of the identified candidate genes is needed to determine their importance in chordoma risk.

Supplementary Materials

The following are available online at https://www.mdpi.com/article/10.3390/cancers13112704/s1, Supplementary Table S1: Details of the criteria for classification of pathogenicity categories, Supplementary Table S2: Genes evaluated in this study and their biological functions, Supplementary Table S3: Details of variants identified in the European ancestry dataset, Supplementary Table S4: Details of potentially pathogenic variants in the Chinese dataset, Supplementary Table S5: Top genes with higher burden of variants.

Author Contributions

Conceptualization, S.Y., J.B., X.R.Y. and A.M.G.; formal analysis, S.Y., N.N.S., H.K., Y.X. and M.W.; investigation, S.Y., J.B., C.L., S.G., M.L.M., K.J., M.W., A.V., B.Z. (Bin Zhu 1), B.Z. (Bin Zhu 2), A.H., M.Y., B.H., B.C., N.D.F., L.B.-F., S.J.C., Y.Z., D.M.P., X.R.Y. and A.M.G.; resources, J.B., C.L., S.G., B.C., N.D.F., L.B.-F., Y.Z., X.R.Y. and A.M.G.; supervision, X.R.Y. and A.M.G.; writing—original draft, S.Y., J.B., X.R.Y. and A.M.G.; writing—review and editing, S.Y., N.N.S., J.B., H.K., C.L., S.G., M.L.M., Y.X., K.J., M.W., A.V., B.Z. (Bin Zhu 1), B.Z. (Bin Zhu 2), A.H., M.Y., B.H., B.C., N.D.F., L.B.-F., S.J.C., Y.Z., D.M.P., X.R.Y. and A.M.G. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the Intramural Research Program of the NIH, NCI, DCEG. Part of this research was supported by Beijing Municipal Science and Technology Commission (Z171100000117002), Research Special Fund for Public Welfare Industry of Health (201402008).

Institutional Review Board Statement

All subjects gave their informed consent for inclusion before they participated in the study. The study was conducted in accordance with the Declaration of Helsinki and the protocol was approved by the Ethics Committee of the National Cancer Institute (NCI) Institutional Review Board (IRB) (Protocols: 10CN188, 04/08/2020; 78C-0039, 01/08/2021) and the ethics committee of the Beijing Tiantan Hospital (IRB code: 2009-47, 20 December 2009).

Informed Consent Statement

Informed consent was obtained from all individual participants included in the study.

Data Availability Statement

According to US NIH policy, the exome sequencing data for the European ancestry dataset will be released to the database of Genotypes and Phenotypes (dbGAP).

Acknowledgments

This work utilized the computational resources of the NIH high performance computational capabilities Biowulf cluster (http://hpc.nih.gov).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Wasserman, J.K.; Gravel, D.; Purgina, B. Chordoma of the Head and Neck: A Review. Head Neck Pathol. 2018, 12, 261–268. [Google Scholar] [CrossRef]
  2. Stacchiotti, S.; Sommer, J. Building a global consensus approach to chordoma: A position paper from the medical and patient community. Lancet Oncol. 2015, 16, e71–e83. [Google Scholar] [CrossRef]
  3. Stacchiotti, S.; Gronchi, A.; Fossati, P.; Akiyama, T.; Alapetite, C.; Baumann, M.; Blay, J.Y.; Bolle, S.; Boriani, S.; Bruzzi, P.; et al. Best practices for the management of local-regional recurrent chordoma: A position paper by the Chordoma Global Consensus Group. Ann. Oncol. 2017, 28, 1230–1242. [Google Scholar] [CrossRef]
  4. Jones, P.S.; Aghi, M.K.; Muzikansky, A.; Shih, H.A.; Barker, F.G.; Curry, W.T. Outcomes and patterns of care in adult skull base chordomas from the Surveillance, Epidemiology, and End Results (SEER) database. J. Clin. Neurosci. 2014, 21, 1490–1496. [Google Scholar] [CrossRef] [PubMed]
  5. 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. [Google Scholar] [CrossRef]
  6. Yang, X.R.; Ng, D.; Alcorta, D.A.; Liebsch, N.J.; Sheridan, E.; Li, S.; Goldstein, A.M.; Parry, D.M.; Kelley, M. T (brachyury) gene duplication confers major susceptibility to familial chordoma. Nat. Genet. 2009, 41, 1176–1178. [Google Scholar] [CrossRef]
  7. Pillay, N.; Plagnol, V.; Tarpey, P.S.; Lobo, S.B.; Presneau, N.; Szuhai, K.; Halai, D.; Berisha, F.; Cannon, S.R.; Mead, S.; et al. A common single-nucleotide variant in T is strongly associated with chordoma. Nat. Genet. 2012, 44, 1185–1187. [Google Scholar] [CrossRef] [PubMed]
  8. Kelley, M.; Shi, J.; Ballew, B.; Hyland, P.L.; Li, W.-Q.; Rotunno, M.; Alcorta, D.A.; Liebsch, N.J.; Mitchell, J.; Bass, S.; et al. Characterization of T gene sequence variants and germline duplications in familial and sporadic chordoma. Qual. Life Res. 2014, 133, 1289–1297. [Google Scholar] [CrossRef]
  9. Vujovic, S.; Henderson, S.; Presneau, N.; Odell, E.; Jacques, T.S.; Tirabosco, R.; Boshoff, C.; Flanagan, A.M. Brachyury, a crucial regulator of notochordal development, is a novel biomarker for chordomas. J. Pathol. 2006, 209, 157–165. [Google Scholar] [CrossRef]
  10. Palena, C.; Polev, D.E.; Tsang, K.Y.; Fernando, R.I.; Litzinger, M.; Krukovskaya, L.L.; Baranova, A.V.; Kozlov, A.P.; Schlom, J. The Human T-Box Mesodermal Transcription Factor Brachyury Is a Candidate Target for T-Cell–Mediated Cancer Immunotherapy. Clin. Cancer Res. 2007, 13, 2471–2478. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  11. Tarpey, P.S.; Behjati, S.; Young, M.D.; Martincorena, I.; Alexandrov, L.B.; Farndon, S.J.; Guzzo, C.; Hardy, C.; Latimer, C.; Butler, A.P.; et al. The driver landscape of sporadic chordoma. Nat. Commun. 2017, 8, 1–6. [Google Scholar] [CrossRef] [Green Version]
  12. Presneau, N.; Shalaby, A.; Ye, H.; Pillay, N.; Halai, D.; Idowu, B.; Tirabosco, R.; Whitwell, D.; Jacques, T.S.; Kindblom, L.G.; et al. Role of the transcription factor T (brachyury) in the pathogenesis of sporadic chordoma: A genetic and functional-based study. J. Pathol. 2011, 223, 327–335. [Google Scholar] [CrossRef]
  13. Hsu, W.; Mohyeldin, A.; Shah, S.R.; Ap Rhys, C.M.; Johnson, L.F.; Sedora-Roman, N.I.; Kosztowski, T.A.; Awad, O.A.; McCarthy, E.F.; Loeb, D.M.; et al. Generation of chordoma cell line JHC7 and the identification of Brachyury as a novel molecular target. J. Neurosurg. 2011, 115, 760–769. [Google Scholar] [CrossRef] [Green Version]
  14. Shah, S.R.; David, J.; Tippens, N.D.; Mohyeldin, A.; Martinez-Gutierrez, J.C.; Ganaha, S.; Schiapparelli, P.; Hamilton, D.H.; Palena, C.; Levchenko, A.; et al. Brachyury-YAP Regulatory Axis Drives Stemness and Growth in Cancer. Cell Rep. 2017, 21, 495–507. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  15. Gohagan, J.K.; Prorok, P.C.; Hayes, R.B.; Kramer, B.S. The Prostate, Lung, Colorectal and Ovarian (PLCO) Cancer Screening Trial of the National Cancer Institute: History, organization, and status. Control. Clin. Trials. 2000, 21, 251S–272S. [Google Scholar] [CrossRef]
  16. Li, H.; Durbin, R. Fast and accurate short read alignment with Burrows-Wheeler transform. Bioinformatics. 2009, 25, 1754–1760. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  17. Li, H.; Handsaker, B.; Wysoker, A.; Fennell, T.; Ruan, J.; Homer, N.; Marth, G.; Abecasis, G.; Durbin, R.; 1000 Genome Project Data Processing Subgroup. The Sequence Alignment/Map Format and SAMtools. Bioinformatics 2009, 25, 2078–2079. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  18. McKenna, A.; Hanna, M.; Banks, E.; Sivachenko, A.; Cibulskis, K.; Kernytsky, A.; Garimella, K.; Altshuler, D.; Gabriel, S.; Daly, M.; et al. The Genome Analysis Toolkit: A MapReduce framework for analyzing next-generation DNA sequencing data. Genome Res. 2010, 20, 1297–1303. [Google Scholar] [CrossRef] [Green Version]
  19. Garrison, E.; Marth, G. Haplotype-based variant detection from short-read sequencing. arXiv 2012, arXiv:1207.3907. [Google Scholar]
  20. Dong, C.; Wei, P.; Jian, X.; Gibbs, R.; Boerwinkle, E.; Wang, K.; Liu, X. Comparison and integration of deleteriousness prediction methods for nonsynonymous SNVs in whole exome sequencing studies. Hum. Mol. Genet. 2015, 24, 2125–2137. [Google Scholar] [CrossRef] [Green Version]
  21. Landrum, M.J.; Chitipiralla, S.; Brown, G.R.; Chen, C.; Gu, B.; Hart, J.; Hoffman, D.; Jang, W.; Kaur, K.; Liu, C.; et al. ClinVar: Improvements to accessing data. Nucleic Acids Res. 2020, 48, D835–D844. [Google Scholar] [CrossRef]
  22. Stenson, P.D.; Mort, M.; Ball, E.V.; Shaw, K.; Phillips, A.D.; Cooper, D.N. The Human Gene Mutation Database: Building a comprehensive mutation repository for clinical and molecular genetics, diagnostic testing and personalized genomic medicine. Qual. Life Res. 2014, 133, 1–9. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  23. Mirabello, L.; Zhu, B.; Koster, R.; Karlins, E.; Dean, M.; Yeager, M.; Gianferante, M.; Spector, L.; Morton, L.M.; Karyadi, D.; et al. Frequency of Pathogenic Germline Variants in Cancer-Susceptibility Genes in Patients With Osteosarcoma. JAMA Oncol. 2020, 6, 724–734. [Google Scholar] [CrossRef]
  24. Lee, S.; Emond, M.J.; Bamshad, M.J.; Barnes, K.C.; Rieder, M.J.; Nickerson, D.A.; Christiani, D.C.; Wurfel, M.M.; Lin, X. Optimal Unified Approach for Rare-Variant Association Testing with Application to Small-Sample Case-Control Whole-Exome Sequencing Studies. Am. J. Hum. Genet. 2012, 91, 224–237. [Google Scholar] [CrossRef] [Green Version]
  25. Ramroop, J.R.; Gerber, M.M.; Toland, A.E. Germline Variants Impact Somatic Events during Tumorigenesis. Trends Genet. 2019, 35, 515–526. [Google Scholar] [CrossRef]
  26. Sharifnia, T.; Wawer, M.J.; Chen, T.; Huang, Q.-Y.; Weir, B.A.; Sizemore, A.; Lawlor, M.A.; Goodale, A.; Cowley, G.S.; Vazquez, F.; et al. Small-molecule targeting of brachyury transcription factor addiction in chordoma. Nat. Med. 2019, 25, 292–300. [Google Scholar] [CrossRef]
  27. Aszódi, A.; Chan, D.; Hunziker, E.; Bateman, J.F.; Fässler, R. Collagen II Is Essential for the Removal of the Notochord and the Formation of Intervertebral Discs. J. Cell Biol. 1998, 143, 1399–1412. [Google Scholar] [CrossRef] [PubMed]
  28. Choi, K.-S.; Cohn, M.J.; Harfe, B.D. Identification of nucleus pulposus precursor cells and notochordal remnants in the mouse: Implications for disk degeneration and chordoma formation. Dev. Dyn. 2008, 237, 3953–3958. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  29. Knutson, S.K.; Warholic, N.M.; Wigle, T.J.; Klaus, C.R.; Allain, C.J.; Raimondi, A.; Scott, M.P.; Chesworth, R.; Moyer, M.P.; Copeland, R.A.; et al. Durable tumor regression in genetically altered malignant rhabdoid tumors by inhibition of methyltransferase EZH2. Proc. Natl. Acad. Sci. USA. 2013, 110, 7922–7927. [Google Scholar] [CrossRef] [Green Version]
  30. Hollmann, T.J.; Hornick, J.L. INI1-deficient tumors: Diagnostic features and molecular genetics. Am. J. Surg. Pathol. 2011, 35, e47–e63. [Google Scholar] [CrossRef]
  31. Yadav, R.; Sharma, M.C.; Malgulwar, P.B.; Pathak, P.; Sigamani, E.; Suri, V.; Sarkar, C.; Kumar, A.; Singh, M.; Sharma, B.S.; et al. Prognostic value of MIB-1, p53, epidermal growth factor receptor, and INI1 in childhood chordomas. Neuro-Oncology 2013, 16, 372–381. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  32. McMaster, M.L.; Goldstein, A.M.; Parry, D.M. Clinical features distinguish childhood chordoma associated with tuberous sclerosis complex (TSC) from chordoma in the general paediatric population. J. Med Genet. 2011, 48, 444–449. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Overview of study samples and prioritization pipeline to evaluate rare and potentially pathogenic variants in genes of interest for chordoma.
Figure 1. Overview of study samples and prioritization pipeline to evaluate rare and potentially pathogenic variants in genes of interest for chordoma.
Cancers 13 02704 g001
Table 1. Information on rare variants identified in the European ancestry patients after the prioritization procedure.
Table 1. Information on rare variants identified in the European ancestry patients after the prioritization procedure.
IDGenderAgeMorphologySiteGeneChrLocationSNP IDSREFVARVariant
Type
Protein
Change
Variant
Impact
Pathogenecity Prediction aHGMD/
ClinVar
MAF in Control DatasetsPathway/Process
METASVMMETALRCADDgnomAD Exomes NFE bgnomAD Genomes NFE c
1004FemaleNAConventional1ATP8B21154305114 GTmissenseArg210LeumoderateDD34 1.06 × 10−5 Mesoderm commitment
1101Male37.2Conventional2LYST1235922291 GAstop_gainedArg2288 *high Driver
1164Male29.7Conventional1TCF7L1285529694rs147750102GAmissenseGly205SermoderateDD22.4 9.67 × 10−56.49 × 10−5Mesoderm commitment
1040Female32.7Conventional1EPB41L52120776677rs200315720GAmissenseArg6HismoderateDD29.9 1.98 × 10−41.30 × 10−4Mesoderm commitment
1020Female39.0Conventional1EPB41L52120833086rs766560121CAmissenseLeu148IlemoderateDD28.9 0 Mesoderm commitment
3001NANANANALRP22170038097rs137983840CTmissenseAla3344ThrmoderateDD26.3 6.16 × 10−56.48 × 10−5Sonic Hedgehog
1035Female26.1Chondroid1PDK12173460594 Tframeshift & stop_gainedAsn424fshigh PI3K/AKT/mTOR
1111Male51.3Chondroid1NFE2L22178095743 CAmissenseAsp530TyrmoderateDD22.5 Mesoderm commitment
1014Female44.3Conventional1BMPR22203420750rs146310981GAmissenseVal788IlemoderateDD22.7 3.52 × 10−5 Mesoderm commitment
1139Female39.5Conventional1RARB325611341 AGmissenseThr188AlamoderateDD22.9 8.79 × 10−6 Mesoderm commitment
1053Female54.6Conventional1HHIP4145581068 ACACframeshiftPhe304fshigh 0 Sonic Hedgehog
1127Male50.7Conventional3SRF643146888rs765592889TCmissenseVal496AlamoderateDD22.4 1.76 × 10−5 Mesoderm commitment
1048/
5320
NANANANATBXT6166571981rs368179445CTmissenseArg377GlnmoderateDD25.5 0 T super-enhancer
1080Female47.4Conventional1ATP6V1B2820068082 CTmissense & splice regionArg130TrpmoderateDD35 0 T super-enhancer
1161Male78.3Conventional1EXT18118832021rs145720047GCmissensePro477ArgmoderateDD22.4 1.24 × 10−46.48 × 10−5Mesoderm commitment
1001Female47.0Conventional1DEPTOR8120977595 AframeshiftGlu185fshigh PI3K/AKT/mTOR
1034Female51.0Conventional2SMARCA292039568rs774084308CTmissensePro153LeumoderateDD22.2 0 SWI/SNF complex
1080Female47.4Conventional1JAK295078325 AGmissenseHis671ArgmoderateDD21.7 Mesoderm commitment
1154Female44.1Conventional2PTCH1998209505rs556901417GAmissenseArg1345CysmoderateDD25.2 1.09 × 10−43.89 × 10−4Sonic Hedgehog
1042Female39.5Conventional1SUFU10104309821rs34406289GAmissenseAla138ThrmoderateDD33 2.64 × 10−50Sonic Hedgehog
1141Female57.6Conventional3SUFU10104375030rs79299301GAmissenseArg343HismoderateDD20.8 1.58 × 10−46.49 × 10−5Sonic Hedgehog
1006Male46.2Conventional1PAX61131823215 ATmissenseVal98GlumoderateDD27.5 Mesoderm commitment
1162Female66.4Conventional3EXT21144254000rs138495222CTmissenseThr620MetmoderateDD34DM9.24 × 10−48.43 × 10−4Mesoderm commitment
1090Female28.3Conventional1GDF3127842985rs146973734CTmissenseArg195GlnmoderateTT0.004P/DM1.85 × 10−45.84 × 10−4Notochord development
1001Female47.0Conventional1COL2A11248380213rs201823490GAmissensePro478LeumoderateDD23.7 00Notochord development
1113Female27.8Conventional1SOX211395364265 GCmissenseAsn13LysmoderateDD21.6 Mesoderm commitment
1050Male60.3Conventional2WDHD11455462357rs556223202GAstop_gainedArg373 *high 8.88 × 10−6 Mesoderm commitment
1127Male50.7Conventional3AKT114105246527rs397514644GAmissenseArg25CysmoderateDT29.6P/DM PI3K/AKT/mTOR
1091Female69.9Conventional2TSC2162130319 CTmissenseAla1184ValmoderateDD24 3.56 × 10−5 PI3K/AKT/mTOR
1028Female39.7Conventional1TSC2162136297rs373635516CTmissensePro1589LeumoderateDD21.9 0 PI3K/AKT/mTOR
1006Male46.2Conventional1ACACA1735603791 ATmissenseVal804GlumoderateDD23.7 Mesoderm commitment
1114Male44.4Conventional1PRKACA1914203391rs41296324ATstop_lostTer125Lysext *high 02.60 × 10−4Mesoderm commitment
1045Female31.2Conventional1FOXA22022563180 CTmissenseGly234SermoderateDD22.3 Mesoderm commitment
1098Female13.8Conventional3SMARCB12224129394 AA frameshiftLys13fshigh SWI/SNF complex
Site: 1, bones of skull and face and associate joints; 2, pelvic bones, sacrum and coccyx and associated joints; 3, vertebral column. Chr, chromosome; REF, reference allele; VAR, variant allele; Freq, frequency; MAF, minor allele frequency; T, tolerant; D, deleterious; P, pathogenic; DM, disease-causing mutation. a Pathogenicity prediction for missense variants based on in silico algorithms, METALR and METASVM, which are ensemble prediction scores that incorporate results from nine algorithms and allele frequency. CADD was also applied. b,c Genome Aggregation Database (gnomAD) exomes and genomes, respectively, in non-Finnish European population (NFE).
Table 2. Potentially pathogenic variants in the Chinese dataset.
Table 2. Potentially pathogenic variants in the Chinese dataset.
IDGenderAgeMorphologyGeneChrLocationSNP IDSREFVARVariant
Type
Protein
Change
Variant
Impact
Pathogenecity Prediction aHGMD/
ClinVar
MAF in Control DatasetsPathway/Process
METASVMMETALRCADDgnomAD Exomes EAS bgnomAD Genomes EAS c
1804M65ConventionalATP8B21154317567 ACmissenseLys836GlnmoderateDD26.2 1.74 × 10−40Mesoderm commitment
1686M64ConventionalLRP22170033035 CTmissenseGly3486GlumoderateDD33 00Sonic Hedgehog
1372NANANALRP22170062924rs577943281TAmissenseThr2436SermoderateDD22.7 2.90 × 10−46.17 × 10−4Sonic Hedgehog
1964M24ChondroidLRP22170055366rs755215116ACmissenseIle2836MetmoderateDD24.2 3.66 × 10−36.17 × 10−4Sonic Hedgehog
1442NANANAPDK12173460575rs745678398AGstructural interaction High 00PI3K/AKT/mTOR
1886F62ConventionalTCF7L1285536476 CTmissenseThr553IlemoderateDD23 5.82 × 10−50Mesoderm commitment
1790M65ConventionalTCF7L1285531113rs373770977GAmissenseVal252IlemoderateDD23.4 1.25 × 10−40Mesoderm commitment
1779M47ConventionalTCF7L1285536536rs555810312CTmissensePro573LeumoderateDD23.2 4.19 × 10−40Mesoderm commitment
2011F7ConventionalTBXT6166581010rs563349798CGmissenseVal24LeumoderateDD29.7 00T super-enhancer
1686M64ConventionalEXT18118819501rs753261171GAmissenseThr613MetmoderateDD33 00Mesoderm commitment
776M54ConventionalSUFU10104309738 CTstructural interaction High 5.80 × 10−50Sonic Hedgehog
1921M57ConventionalEXT21144193231rs767085143AGmissenseAsn448SermoderateDD24.3 2.90 × 10−40Mesoderm commitment
1372NANANACOL2A11248369322rs995646562CTmissenseAla1222ThrmoderateDD24.1 00Notochord development
1909M26ConventionalCOL2A11248372528 GTmissensePro916HismoderateDD25 00Notochord development
1936F71ConventionalTSC2162134572rs45517338CGmissensePro1450ArgmoderateDD25.8 2.35 × 10−40PI3K/AKT/mTOR
1417M59ConventionalTSC2162134572rs45517338CGmissensePro1450ArgmoderateDD25.8 2.35 × 10−40PI3K/AKT/mTOR
1979M59ChondroidTSC2162121610rs45509392GAmissenseAsp647AsnmoderateDD32DM5.80 × 10−40PI3K/AKT/mTOR
1570M24ChondroidTSC2162121610rs45509392GAmissenseAsp647AsnmoderateDD32DM5.80 × 10−40PI3K/AKT/mTOR
462F62ConventionalTSC2162121610rs45509392GAmissenseAsp647AsnmoderateDD32DM5.80 × 10−40PI3K/AKT/mTOR
1084NANANAACACA1735549102rs772483773CTmissenseVal1449MetmoderateDD25.2 6.96 × 10−46.17 × 10−4Mesoderm commitment
Chr, chromosome; REF, reference allele; VAR, variant allele; Freq, frequency; MAF, minor allele frequency; T, tolerant; D, deleterious; P, pathogenic; DM, disease-causing mutation. a Pathogenicity prediction for missense variants based on in silico algorithms, METALR and METASVM, which are ensemble prediction scores that incorporate results from nine algorithms and allele frequency. CADD was also applied. b,c Genome Aggregation Database (gnomAD) exomes and genomes, respectively, in East Asian population. Structural interaction: variants that impact the internal interactions of the resulting polypeptide structure.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Yepes, S.; Shah, N.N.; Bai, J.; Koka, H.; Li, C.; Gui, S.; McMaster, M.L.; Xiao, Y.; Jones, K.; Wang, M.; et al. Rare Germline Variants in Chordoma-Related Genes and Chordoma Susceptibility. Cancers 2021, 13, 2704. https://doi.org/10.3390/cancers13112704

AMA Style

Yepes S, Shah NN, Bai J, Koka H, Li C, Gui S, McMaster ML, Xiao Y, Jones K, Wang M, et al. Rare Germline Variants in Chordoma-Related Genes and Chordoma Susceptibility. Cancers. 2021; 13(11):2704. https://doi.org/10.3390/cancers13112704

Chicago/Turabian Style

Yepes, Sally, Nirav N. Shah, Jiwei Bai, Hela Koka, Chuzhong Li, Songbai Gui, Mary Lou McMaster, Yanzi Xiao, Kristine Jones, Mingyi Wang, and et al. 2021. "Rare Germline Variants in Chordoma-Related Genes and Chordoma Susceptibility" Cancers 13, no. 11: 2704. https://doi.org/10.3390/cancers13112704

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

Article Metrics

Back to TopTop