Multimodal analysis of cell-free DNA whole-genome sequencing for pediatric cancers with low mutational burden

Peter Peneder, Adrian M Stütz, Didier Surdez, Manuela Krumbholz, Sabine Semper, Mathieu Chicard, Nathan C Sheffield, Gaelle Pierron, Eve Lapouble, Marcus Tötzl, Bekir Ergüner, Daniele Barreca, André F Rendeiro, Abbas Agaimy, Heidrun Boztug, Gernot Engstler, Michael Dworzak, Marie Bernkopf, Sabine Taschner-Mandl, Inge M Ambros, Ola Myklebost, Perrine Marec-Bérard, Susan Ann Burchill, Bernadette Brennan, Sandra J Strauss, Jeremy Whelan, Gudrun Schleiermacher, Christiane Schaefer, Uta Dirksen, Caroline Hutter, Kjetil Boye, Peter F Ambros, Olivier Delattre, Markus Metzler, Christoph Bock, Eleni M Tomazou, Peter Peneder, Adrian M Stütz, Didier Surdez, Manuela Krumbholz, Sabine Semper, Mathieu Chicard, Nathan C Sheffield, Gaelle Pierron, Eve Lapouble, Marcus Tötzl, Bekir Ergüner, Daniele Barreca, André F Rendeiro, Abbas Agaimy, Heidrun Boztug, Gernot Engstler, Michael Dworzak, Marie Bernkopf, Sabine Taschner-Mandl, Inge M Ambros, Ola Myklebost, Perrine Marec-Bérard, Susan Ann Burchill, Bernadette Brennan, Sandra J Strauss, Jeremy Whelan, Gudrun Schleiermacher, Christiane Schaefer, Uta Dirksen, Caroline Hutter, Kjetil Boye, Peter F Ambros, Olivier Delattre, Markus Metzler, Christoph Bock, Eleni M Tomazou

Abstract

Sequencing of cell-free DNA in the blood of cancer patients (liquid biopsy) provides attractive opportunities for early diagnosis, assessment of treatment response, and minimally invasive disease monitoring. To unlock liquid biopsy analysis for pediatric tumors with few genetic aberrations, we introduce an integrated genetic/epigenetic analysis method and demonstrate its utility on 241 deep whole-genome sequencing profiles of 95 patients with Ewing sarcoma and 31 patients with other pediatric sarcomas. Our method achieves sensitive detection and classification of circulating tumor DNA in peripheral blood independent of any genetic alterations. Moreover, we benchmark different metrics for cell-free DNA fragmentation analysis, and we introduce the LIQUORICE algorithm for detecting circulating tumor DNA based on cancer-specific chromatin signatures. Finally, we combine several fragmentation-based metrics into an integrated machine learning classifier for liquid biopsy analysis that exploits widespread epigenetic deregulation and is tailored to cancers with low mutation rates. Clinical associations highlight the potential value of cfDNA fragmentation patterns as prognostic biomarkers in Ewing sarcoma. In summary, our study provides a comprehensive analysis of circulating tumor DNA beyond recurrent genetic aberrations, and it renders the benefits of liquid biopsy more readily accessible for childhood cancers.

Trial registration: ClinicalTrials.gov NCT02613962.

Conflict of interest statement

The authors declare no competing interests.

Figures

Fig. 1. Whole-genome sequencing of cfDNA enables…
Fig. 1. Whole-genome sequencing of cfDNA enables fragment-based liquid biopsy analysis in Ewing sarcoma.
The top row of the figure describes the fragmentation and fragment-based analysis of cfDNA in cancer patients. The center row introduces the study cohort (center left) and illustrates the quantification of tumor-derived DNA based on genetic evidence, which is used as a reference in this study (center right). The bottom part of the figure outlines three complementary approaches to fragment-based cfDNA analyses: global fragment-size distribution; regional fragment-size distribution along the genome; and fragment coverage at regions-of-interest (bottom left). CNA profiles were used for comparing cfDNA to matched tumors biopsies and for time-resolved monitoring of tumor evolution. Fragment-based cfDNA metrics were combined for machine learning-based tumor detection and classification (bottom right). The main figures describing each of the analyses are indicated in brackets.
Fig. 2. Global fragment-size analysis detects highly…
Fig. 2. Global fragment-size analysis detects highly fragmented EwS tumor DNA.
a Histogram (top) showing the cfDNA fragment-size distribution for three representative samples with high (95%), low (2%), and undetectable (0%) tumor-derived DNA (ctDNA) content. The range of cfDNA fragment sizes in 22 healthy controls is shown in gray. Heatmap (bottom) showing the relative fragment-size distribution of 235 cfDNA samples subjected to whole-genome sequencing, each normalized against the median of 22 healthy controls. EwS samples are grouped by genetically inferred tumor-derived DNA content. The three samples shown in the histogram are marked by arrows. b Proportion of short cfDNA fragments (20-150 bp) for pediatric sarcomas and healthy controls (data from this study) and for adult cancers (published data). Boxes correspond to interquartile ranges (IQR), black lines indicate the median, and the whiskers extend to the lowest or highest data points that are still within 1.5 IQR of the bottom or top quartile, respectively. Significance versus the 22 healthy controls was assessed using two-sided Mann–Whitney U tests.
Fig. 3. CNA profiles in liquid biopsies…
Fig. 3. CNA profiles in liquid biopsies reflect tumor aberrations and allow monitoring of tumor evolution.
a Comparison of CNAs detected in cfDNA versus matched tumor biopsies. Only sample pairs with tumor-derived DNA detected in cfDNA based on ichorCNA are shown (n = 29 sample pairs); four copy-number neutral sample pairs were omitted from the plot. Patients are grouped according to CNA state in cfDNA relative to the matched tumor biopsy. Gray represents CNAs detected in both cfDNA and matched tumor biopsy, orange indicates CNAs detected only in cfDNA, and turquoise represents CNAs detected only in the tumor biopsy. The CNAs detected in cfDNAs versus matched tumor biopsies are summarized in a bar plot (bottom). b CNA plot (ichorCNA) of an EwS cfDNA sample (EwS_90_1) before (middle) and after (bottom) in silico size selection to the range of 90–150 bp. A subclonal CNA on chromosome 16 (indicated by black arrows) that was clearly visible in the tumor biopsy (top) became detectable in the matched cfDNA sample only after in silico size selection. c CNA profiles (ichorCNA) of longitudinal cfDNA samples derived from the same patient (EwS_5) support the monitoring of somatic clonal evolution for individual patients. The filtered CNA profiles of samples collected at diagnosis and two subsequent relapses are shown. The day of sample collection relative to the day of diagnosis is indicated (left). Inferred chromosomal gains are shown in red, inferred deletions are shown in green, and CNA-neutral regions are shown in blue.
Fig. 4. Regional fragment-size analysis detects an…
Fig. 4. Regional fragment-size analysis detects an EwS tumor-specific epigenetic signature in cfDNA samples.
a Schematic illustration of the regional fragment-size analysis, measuring the ratio of short (S) versus long (L) cfDNA fragments in 100 kb bins along the genome. Genomic regions that overlap with CNAs are excluded in order to focus the analysis on epigenetic signatures reflected in the cfDNA fragmentation patterns. b Heatmap comparing the genome-wide fragmentation profiles of cfDNA samples from patients with pediatric sarcoma to those of healthy controls. In each 100 kb bin (n = 20,706 bins), the log2(S/L ratio) of each sarcoma sample was compared to the distribution of log2(S/L ratios) of healthy controls via z-scores. Both CNA-affected and CNA-neutral bins are shown. EwS samples are grouped by genetically inferred tumor-derived DNA content. c Regional cfDNA fragmentation in patients with pediatric sarcoma compared to healthy controls. Only chromosome arms that are recurrently affected by CNAs in EwS tumors are shown. Box plots illustrate z-scores for EwS samples with genetic tumor evidence and without detected CNAs on the chromosomal arm (red), non-EwS sarcomas with genetic tumor evidence and without detected CNAs on the chromosomal arm (black), EwS samples without genetic tumor evidence (yellow), and healthy controls (green). The significance of the first group versus each of the other three groups was assessed using the two-sided Mann–Whitney U test; Bonferroni-corrected p-values are shown. Boxes correspond to interquartile ranges (IQR), thick black lines indicate the median, and the whiskers extend to the lowest or highest data point that are still within 1.5 IQR of the bottom or top quartile, respectively. d Functional enrichment analysis for regions with significantly shorter/longer cfDNA fragment size compared to healthy controls based on the LOLA software. A selection of enriched terms is shown, while the full list is provided in Supplementary Data 6. e EwS tumor-specific epigenome profiles for selected regions with significantly shorter cfDNA fragment size compared to healthy controls. The genome browser profiles show open chromatin-associated histone H3K27 acetylation (for regions with shorter fragments) based on ChIP-seq data for primary EwS tumors. EwS-specific DHSs along the selected genomic region are also indicated.
Fig. 5. Fragment analysis for EwS-specific genomic…
Fig. 5. Fragment analysis for EwS-specific genomic regions quantifies tumor-derived cfDNA in EwS patients.
a Conceptual outline of the LIQUORICE method and software for fragment analysis of cfDNA based on tumor-specific epigenetic alterations. b Aggregated, bias-corrected, and normalized coverage signals at selected genomic region sets shown for healthy controls, for non-EwS sarcomas, and for EwS cfDNA samples. EwS samples are grouped by genetically inferred tumor-derived DNA content and clinical tumor evidence. cfDNA samples with coverage signals significantly different (three standard deviations) from healthy controls are displayed in red; the total number of those samples and the direction of the deviation (arrow) are indicated. Total dip depth was used as the metric of choice for the sharp dips at hematopoietic-specific and universal DHSs; area over the curve (AOC) was used for the other region sets. c cfDNA-based coverage signal at EwS-specific DHSs (bottom, nEwS = 38) reflects the aggregate DNA methylation profiles at these regions in matched tumor biopsies (top, nEwS = 38). d Scatterplots showing the correlation of the coverage signal at EwS-specific DHSs with the genetically inferred tumor-derived DNA content of the cfDNA samples. Pearson correlation coefficients (r) and linear trend lines are shown. The x-axes are shown in a log scale from 1% onwards. e Same as d but showing the coverage signal at hematopoietic-specific DHSs. Blue arrows indicate samples with significant liver signature. f Aggregated, bias-corrected, and normalized coverage signal (AOC) at alveolar rhabdomyosarcoma (ARMS)-specific DHSs for cfDNA samples from healthy controls and patients with EwS, RMS, and other pediatric sarcomas (left; p-values were calculated using two-sided Mann–Whitney U tests without correction for multiple testing). For ARMS patients with at least 9% ctDNA (genetic-based evidence), a striking reduction of fragment coverage was observed (right). A cfDNA sample from a patient with embryonal rhabdomyosarcoma (ERMS) did not show any reduction of fragment coverage at ARMS-specific DHSs (bottom right).
Fig. 6. Fragment-based analysis of cfDNA enables…
Fig. 6. Fragment-based analysis of cfDNA enables accurate tumor detection and classification.
Prediction performance of machine learning classifiers trained to distinguish patients with EwS from healthy controls (a) and from patients with other pediatric sarcomas (b), based on the following sets of input features: global fragment-size distribution (blue); fragment coverage at EwS-specific DHSs (orange); read depth in 5 Mb bins (green); and regional fragmentation patterns (red). Results are also shown for a meta-learner combining the predictions of all individual classifiers into a weighted consensus prediction (purple). The performance of each model was evaluated by and averaged over 100 iterations of bootstrapping, separately for the different sequencing coverage levels (median of 12×, 1×, and 0.1×). CI is the 95% confidence interval obtained by bootstrapping. a ROC curves show, for each feature set, the performance for distinguishing between cfDNA samples from patients with clinical evidence for EwS (nsamples = 103) and healthy controls from three independent sets (22 controls sequenced in this study; 22 controls from Cristiano et al.; and 24 controls from Ulz et al.). Machine learning models were trained separately for each of the 3 control sets; the mean results over the 3*100 bootstrap iterations are shown. b ROC curves show the performance of each feature set for distinguishing between cfDNA samples from patients with EwS (nsamples = 98) and from patients with other pediatric sarcomas (nsamples = 18). For both sets of samples, we ensured the presence of tumor-derived cfDNA in the blood based on genetic evidence.

References

    1. Corcoran RB, Chabner BA. Application of cell-free DNA analysis to cancer treatment. N. Engl. J. Med. 2018;379:1754–1765. doi: 10.1056/NEJMra1706174.
    1. Heitzer E, Haque IS, Roberts CES, Speicher MR. Current and future perspectives of liquid biopsies in genomics-driven oncology. Nat. Rev. Genet. 2019;20:71–88. doi: 10.1038/s41576-018-0071-5.
    1. Siravegna G, Marsoni S, Siena S, Bardelli A. Integrating liquid biopsies into the management of cancer. Nat. Rev. Clin. Oncol. 2017;14:531–548. doi: 10.1038/nrclinonc.2017.14.
    1. Wan JCM, et al. Liquid biopsies come of age: towards implementation of circulating tumour DNA. Nat. Rev. Cancer. 2017;17:223–238. doi: 10.1038/nrc.2017.7.
    1. van der Pol Y, Mouliere F. Toward the early detection of cancer by decoding the epigenetic and environmental fingerprints of cell-free DNA. Cancer Cell. 2019;36:350–368. doi: 10.1016/j.ccell.2019.09.003.
    1. Abbosh C, et al. Phylogenetic ctDNA analysis depicts early-stage lung cancer evolution. Nature. 2017;545:446–451. doi: 10.1038/nature22364.
    1. Adalsteinsson VA, et al. Scalable whole-exome sequencing of cell-free DNA reveals high concordance with metastatic tumors. Nat. Commun. 2017;8:1324. doi: 10.1038/s41467-017-00965-y.
    1. Dawson SJ, et al. Analysis of circulating tumor DNA to monitor metastatic breast cancer. N. Engl. J. Med. 2013;368:1199–1209. doi: 10.1056/NEJMoa1213261.
    1. Parikh AR, et al. Liquid versus tissue biopsy for detecting acquired resistance and tumor heterogeneity in gastrointestinal cancers. Nat. Med. 2019;25:1415–1421. doi: 10.1038/s41591-019-0561-9.
    1. Cohen JD, et al. Detection and localization of surgically resectable cancers with a multi-analyte blood test. Science. 2018;359:926–930. doi: 10.1126/science.aar3247.
    1. Bettegowda C, et al. Detection of circulating tumor DNA in early- and late-stage human malignancies. Sci. Transl. Med. 2014;6:224ra224. doi: 10.1126/scitranslmed.3007094.
    1. Visvanathan K, et al. Monitoring of serum DNA methylation as an early independent marker of response and survival in metastatic breast cancer: TBCRC 005 Prospective Biomarker Study. J. Clin. Oncol. 2017;35:751–758. doi: 10.1200/JCO.2015.66.2080.
    1. Chan KC, et al. Noninvasive detection of cancer-associated genome-wide hypomethylation and copy number aberrations by plasma DNA bisulfite sequencing. Proc. Natl Acad. Sci. USA. 2013;110:18761–18768. doi: 10.1073/pnas.1313995110.
    1. Sun K, et al. Plasma DNA tissue mapping by genome-wide methylation sequencing for noninvasive prenatal, cancer, and transplantation assessments. Proc. Natl Acad. Sci. USA. 2015;112:E5503–E5512. doi: 10.1073/pnas.1508736112.
    1. Mattox, A. K. et al. Applications of liquid biopsies for cancer. Sci. Transl. Med.11, eaay1984 (2019).
    1. Cescon DW, Bratman SV, Chan SM, Siu LL. Circulating tumor DNA and liquid biopsy in oncology. Nat. Cancer. 2020;1:276–290. doi: 10.1038/s43018-020-0043-5.
    1. Shulman DS, et al. Detection of circulating tumour DNA is associated with inferior outcomes in Ewing sarcoma and osteosarcoma: a report from the Children’s Oncology Group. Br. J. Cancer. 2018;119:615–621. doi: 10.1038/s41416-018-0212-9.
    1. Andersson, D., Fagman, H., Dalin, M. G. & Stahlberg, A. Circulating cell-free tumor DNA analysis in pediatric cancers. Mol. Aspects Med. 72, 100819 (2019).
    1. Abbou SD, Shulman DS, DuBois SG, Crompton BD. Assessment of circulating tumor DNA in pediatric solid tumors: the promise of liquid biopsies. Pediatr. Blood Cancer. 2019;66:e27595. doi: 10.1002/pbc.27595.
    1. Chicard, M. et al. Whole-exome sequencing of cell-free DNA reveals temporo-spatial heterogeneity and identifies treatment-resistant clones in neuroblastoma. Clin. Cancer Res.24, 939–949 (2017).
    1. Jimenez I, et al. Circulating tumor DNA analysis enables molecular characterization of pediatric renal tumors at diagnosis. International journal of cancer. J. Int. Cancer. 2019;144:68–79. doi: 10.1002/ijc.31620.
    1. Klega, K. et al. Detection of somatic structural variants enables quantification and characterization of circulating tumor DNA in children with solid tumors. JCO Precis. Oncol.2018, PO.17.00285 (2018).
    1. Krumbholz M, et al. Genomic EWSR1 fusion sequence as highly sensitive and dynamic plasma tumor marker in Ewing sarcoma. Clin. Cancer Res. 2016;22:4356–4365. doi: 10.1158/1078-0432.CCR-15-3028.
    1. Eguchi-Ishimae M, et al. Early detection of the PAX3-FOXO1 fusion gene in circulating tumor-derived DNA in a case of alveolar rhabdomyosarcoma. Genes Chromosomes Cancer. 2019;58:521–529. doi: 10.1002/gcc.22734.
    1. Shukla, N. N. et al. Plasma DNA-based molecular diagnosis, prognostication, and monitoring of patients with EWSR1 fusion-positive sarcomas. JCO Precis. Oncol.2017, PO.16.00028 (2017).
    1. Barris DM, et al. Detection of circulating tumor DNA in patients with osteosarcoma. Oncotarget. 2018;9:12695–12704. doi: 10.18632/oncotarget.24268.
    1. Van Roy N, et al. Shallow whole genome sequencing on circulating cell-free DNA allows reliable noninvasive copy-number profiling in neuroblastoma patients. Clin. Cancer Res. 2017;23:6305–6314. doi: 10.1158/1078-0432.CCR-17-0675.
    1. Phallen J, et al. Direct detection of early-stage cancers using circulating tumor DNA. Sci. Transl. Med. 2017;9:eaan2415. doi: 10.1126/scitranslmed.aan2415.
    1. Zviran A, et al. Genome-wide cell-free DNA mutational integration enables ultra-sensitive cancer monitoring. Nat. Med. 2020;26:1114–1124. doi: 10.1038/s41591-020-0915-3.
    1. Grobner SN, et al. The landscape of genomic alterations across childhood cancers. Nature. 2018;555:321–327. doi: 10.1038/nature25480.
    1. Ma X, et al. Pan-cancer genome and transcriptome analyses of 1,699 paediatric leukaemias and solid tumours. Nature. 2018;555:371–376. doi: 10.1038/nature25795.
    1. Cristiano S, et al. Genome-wide cell-free DNA fragmentation in patients with cancer. Nature. 2019;570:385–389. doi: 10.1038/s41586-019-1272-6.
    1. Snyder MW, Kircher M, Hill AJ, Daza RM, Shendure J. Cell-free DNA comprises an in vivo nucleosome footprint that informs its tissues-of-origin. Cell. 2016;164:57–68. doi: 10.1016/j.cell.2015.11.050.
    1. Ulz P, et al. Inference of transcription factor binding from cell-free DNA enables tumor subtype prediction and early detection. Nat. Commun. 2019;10:4666. doi: 10.1038/s41467-019-12714-4.
    1. Sun K, et al. Orientation-aware plasma cell-free DNA fragmentation analysis in open chromatin regions informs tissue of origin. Genome Res. 2019;29:418–427. doi: 10.1101/gr.242719.118.
    1. Erkek S, et al. Comprehensive analysis of chromatin states in atypical teratoid/rhabdoid tumor identifies diverging roles for SWI/SNF and polycomb in gene regulation. Cancer Cell. 2019;35:95–110. doi: 10.1016/j.ccell.2018.11.014.
    1. van Groningen T, et al. Neuroblastoma is composed of two super-enhancer-associated differentiation states. Nat. Genet. 2017;49:1261–1266. doi: 10.1038/ng.3899.
    1. Gryder BE, et al. PAX3-FOXO1 establishes myogenic super enhancers and confers BET bromodomain vulnerability. Cancer Discov. 2017;7:884–899. doi: 10.1158/-16-1297.
    1. Banito A, et al. The SS18-SSX oncoprotein hijacks KDM2B-PRC1.1 to drive synovial sarcoma. Cancer Cell. 2018;33:527–541. doi: 10.1016/j.ccell.2018.01.018.
    1. Northcott PA, et al. The whole-genome landscape of medulloblastoma subtypes. Nature. 2017;547:311–317. doi: 10.1038/nature22973.
    1. Sheffield NC, et al. DNA methylation heterogeneity defines a disease spectrum in Ewing sarcoma. Nat. Med. 2017;23:386–395. doi: 10.1038/nm.4273.
    1. Boeva V, et al. Heterogeneity of neuroblastoma cell identity defined by transcriptional circuitries. Nat. Genet. 2017;49:1408–1413. doi: 10.1038/ng.3921.
    1. Larson JD, et al. Histone H3.3 K27M accelerates spontaneous brainstem glioma and drives restricted changes in bivalent gene expression. Cancer Cell. 2019;35:140–155. doi: 10.1016/j.ccell.2018.11.015.
    1. Thirant C, et al. ETO2-GLIS2 hijacks transcriptional complexes to drive cellular identity and self-renewal in pediatric acute megakaryoblastic leukemia. Cancer Cell. 2017;31:452–465. doi: 10.1016/j.ccell.2017.02.006.
    1. Grünewald TGP, et al. Ewing sarcoma. Nat. Rev. Dis. Prim. 2018;4:5. doi: 10.1038/s41572-018-0003-x.
    1. Brohl AS, et al. The genomic landscape of the ewing sarcoma family of tumors reveals recurrent STAG2 mutation. PLoS Genet. 2014;10:e1004475. doi: 10.1371/journal.pgen.1004475.
    1. Crompton BD, et al. The genomic landscape of pediatric ewing sarcoma. Cancer Discov. 2014;4:1326–1341. doi: 10.1158/-13-1037.
    1. Tirode F, et al. Genomic landscape of ewing sarcoma defines an aggressive subtype with co-association of STAG2 and TP53 mutations. Cancer Discov. 2014;4:1342–1353. doi: 10.1158/-14-0622.
    1. Mouliere, F. et al. Enhanced detection of circulating tumor DNA by fragment size analysis. Sci. Transl. Med.10, eaat4921 (2018).
    1. Underhill HR, et al. Fragment length of circulating tumor DNA. PLoS Genet. 2016;12:e1006162. doi: 10.1371/journal.pgen.1006162.
    1. Cheng L, et al. Integration of genomic copy number variations and chemotherapy-response biomarkers in pediatric sarcoma. BMC Med. Genomics. 2019;12:23. doi: 10.1186/s12920-018-0456-5.
    1. Gerstung M, et al. The evolutionary history of 2,658 cancers. Nature. 2020;578:122–128. doi: 10.1038/s41586-019-1907-7.
    1. Ivanov M, Baranova A, Butler T, Spellman P, Mileyko V. Non-random fragmentation patterns in circulating cell-free DNA reflect epigenetic regulation. BMC Genomics. 2015;16:S1. doi: 10.1186/1471-2164-16-S13-S1.
    1. Sheffield NC, Bock C. LOLA: enrichment analysis for genomic region sets and regulatory elements in R and Bioconductor. Bioinformatics. 2016;32:587–589. doi: 10.1093/bioinformatics/btv612.
    1. Moss J, et al. Comprehensive human cell-type methylation atlas reveals origins of circulating cell-free DNA in health and disease. Nat. Commun. 2018;9:5068. doi: 10.1038/s41467-018-07466-6.
    1. van der Lelij P, et al. Synthetic lethality between the cohesin subunits STAG1 and STAG2 in diverse cancer contexts. Elife. 2017;6:e26980. doi: 10.7554/eLife.26980.
    1. Boulay G, et al. Cancer-specific retargeting of BAF complexes by a prion-like domain. Cell. 2017;171:163–178. doi: 10.1016/j.cell.2017.07.036.
    1. Riggi N, et al. EWS-FLI1 utilizes divergent chromatin remodeling mechanisms to directly activate or repress enhancer elements in Ewing sarcoma. Cancer Cell. 2014;26:668–681. doi: 10.1016/j.ccell.2014.10.004.
    1. Tomazou EM, et al. Epigenome mapping reveals distinct modes of gene regulation and widespread enhancer reprogramming by the oncogenic fusion protein EWS-FLI1. Cell Rep. 2015;10:1082–1095. doi: 10.1016/j.celrep.2015.01.042.
    1. Shern JF, et al. Comprehensive genomic analysis of rhabdomyosarcoma reveals a landscape of alterations affecting a common genetic axis in fusion-positive and fusion-negative tumors. Cancer Discov. 2014;4:216–231. doi: 10.1158/-13-0639.
    1. Meuleman W, et al. Index and biological spectrum of human DNase I hypersensitive sites. Nature. 2020;584:244–251. doi: 10.1038/s41586-020-2559-3.
    1. Renzi S, Anderson ND, Light N, Gupta A. Ewing-like sarcoma: an emerging family of round cell sarcomas. J. Cell Physiol. 2019;234:7999–8007. doi: 10.1002/jcp.27558.
    1. Machado I, Navarro S, Llombart-Bosch A. Ewing sarcoma and the new emerging Ewing-like sarcomas: (CIC and BCOR-rearranged-sarcomas). A systematic review. Histol. Histopathol. 2016;31:1169–1181.
    1. Ulz P, et al. Inferring expressed genes by whole-genome sequencing of plasma DNA. Nat. Genet. 2016;48:1273–1278. doi: 10.1038/ng.3648.
    1. Jiang P, Lo YMD. The long and short of circulating cell-free DNA and the Ins and outs of molecular diagnostics. Trends Genet. 2016;32:360–371. doi: 10.1016/j.tig.2016.03.009.
    1. Gaspar N, et al. Ewing sarcoma: current management and future approaches through collaboration. J. Clin. Oncol. 2015;33:3036–3046. doi: 10.1200/JCO.2014.59.5256.
    1. Ferrari S, et al. Nonmetastatic Ewing family tumors: high-dose chemotherapy with stem cell rescue in poor responder patients. Results of the Italian Sarcoma Group/Scandinavian Sarcoma Group III protocol. Ann. Oncol. 2011;22:1221–1227. doi: 10.1093/annonc/mdq573.
    1. Heitzer E, et al. Establishment of tumor-specific copy number alterations from plasma DNA of patients with cancer. J. Int. Cancer. 2013;133:346–356. doi: 10.1002/ijc.28030.
    1. Gerber T, et al. Assessment of pre-analytical sample handling conditions for comprehensive liquid biopsy analysis. J. Mol. Diagn. 2020;22:1070–1086. doi: 10.1016/j.jmoldx.2020.05.006.
    1. Chen S, Zhou Y, Chen Y, Gu J. fastp: an ultra-fast all-in-one FASTQ preprocessor. Bioinformatics. 2018;34:i884–i890. doi: 10.1093/bioinformatics/bty560.
    1. Li, H. Aligning sequence reads, clone sequences and assembly contigs with BWA-MEM. Preprint at (2013).
    1. Faust GG, Hall IM. SAMBLASTER: fast duplicate marking and structural variant read extraction. Bioinformatics. 2014;30:2503–2505. doi: 10.1093/bioinformatics/btu314.
    1. Lawson, J. T., Tomazou, E. M., Bock, C. & Sheffield, N. C. MIRA: an R package for DNA methylation-based inference of regulatory activity. Bioinformatics34, 2649–2650 (2018).
    1. Robinson JT, et al. Integrative genomics viewer. Nat. Biotechnol. 2011;29:24–26. doi: 10.1038/nbt.1754.
    1. Kent WJ. BLAT–the BLAST-like alignment tool. Genome Res. 2002;12:656–664. doi: 10.1101/gr.229202.
    1. Lai, D., Ha, G. & Shah, S. HMMcopy: copy number prediction with correction for GC and mappability bias for HTS data. R package version 1.2.0. (2019).
    1. Ramirez F, et al. deepTools2: a next generation web server for deep-sequencing data analysis. Nucleic Acids Res. 2016;44:W160–W165. doi: 10.1093/nar/gkw257.
    1. Amemiya HM, Kundaje A, Boyle AP. The ENCODE blacklist: identification of problematic regions of the genome. Sci. Rep. 2019;9:9354. doi: 10.1038/s41598-019-45839-z.
    1. Benjamini Y, Speed TP. Summarizing and correcting the GC content bias in high-throughput sequencing. Nucleic Acids Res. 2012;40:e72. doi: 10.1093/nar/gks001.
    1. Romiguier J, Roux C. Analytical biases associated with GC-content in molecular evolution. Front. Genet. 2017;8:16. doi: 10.3389/fgene.2017.00016.
    1. Derrien T, et al. Fast computation and applications of genome mappability. PLoS ONE. 2012;7:e30377. doi: 10.1371/journal.pone.0030377.
    1. Levenberg K. A method for the solution of certain non-linear problems in least squares. Q. Appl. Math. 1944;2.2:164–168. doi: 10.1090/qam/10666.
    1. Pedregosa, F., Varoquaux, G., Gramfort, A., Michel, V. & Thirion, B. Scikit-learn: machine learning in Python. J. Machine Learn. Res.12, 2825–2830 (2011).

Source: PubMed

3
Suscribir