Alternative tumour-specific antigens

Christof C Smith, Sara R Selitsky, Shengjie Chai, Paul M Armistead, Benjamin G Vincent, Jonathan S Serody, Christof C Smith, Sara R Selitsky, Shengjie Chai, Paul M Armistead, Benjamin G Vincent, Jonathan S Serody

Abstract

The study of tumour-specific antigens (TSAs) as targets for antitumour therapies has accelerated within the past decade. The most commonly studied class of TSAs are those derived from non-synonymous single-nucleotide variants (SNVs), or SNV neoantigens. However, to increase the repertoire of available therapeutic TSA targets, 'alternative TSAs', defined here as high-specificity tumour antigens arising from non-SNV genomic sources, have recently been evaluated. Among these alternative TSAs are antigens derived from mutational frameshifts, splice variants, gene fusions, endogenous retroelements and other processes. Unlike the patient-specific nature of SNV neoantigens, some alternative TSAs may have the advantage of being widely shared by multiple tumours, allowing for universal, off-the-shelf therapies. In this Opinion article, we will outline the biology, available computational tools, preclinical and/or clinical studies and relevant cancers for each alternative TSA class, as well as discuss both current challenges preventing the therapeutic application of alternative TSAs and potential solutions to aid in their clinical translation.

Figures

Figure 1:. Summary of tumour-specific antigen production…
Figure 1:. Summary of tumour-specific antigen production in the tumour cell.
Mutations and other tumour-specific nucleotide sequences (shown in red) can be observed at the genomic DNA level, where they undergo transcription (1) and splicing to form mRNA (2). Alternative splicing can occur at this step to form splice variant mRNA. Next, translation occurs on variant mRNA, resulting in production of variant proteins (3). Post-translational frameshifts (e.g. ribosomal slippage, among other mechanisms) can occur at this step, resulting in frameshifted protein variants. These proteins then can undergo proteasomal degradation (4) and transport to the endoplasmic reticulum (ER) to subsequently be loaded on major histocompatibility complexes (MHCs) (5). Other forms of post-translational frameshift can occur during these steps (e.g. protein splicing). Lastly, peptides containing variant sequences can be presented at the cell surface in the context of MHC, resulting in T-cell targetable tumour-specific antigens (6).
Figure 2:. Average tumour-specific antigen counts by…
Figure 2:. Average tumour-specific antigen counts by cancer type.
Plots represent number of unique identified epitopes by The Cancer Genome Atlas (TCGA) cancer type. Insertion or deletion (INDEL)-neoantigen counts demonstrated significant correlation with single nucleotide variant (SNV)-neoantigens among all cancer types (coefficient: 0.81, p 10-fold tumour-vs-mean normal expression by DESeq2) in Smith et al. (JCI, 2018). All TSA classes represent the average number of predicted class I human leukocyte antigen (HLA) binders (8–11mers,

Figure 3:. Computational workflow for tumour-specific antigen…

Figure 3:. Computational workflow for tumour-specific antigen calling.

a) Identification of tumour-specific antigens begins with…

Figure 3:. Computational workflow for tumour-specific antigen calling.
a) Identification of tumour-specific antigens begins with variant calling. This can be done through comparison of tumour versus normal tissue DNA sequences (single nucleotide variants (SNVs) and insertions or deletions (INDELs)) or RNA sequences (splice variants, fusions, viral sequences and retroelements) to look for tumour-specific variants in the exome or tumour-specific transcripts in the transcriptome, respectively. b) Tumour human leukocyte antigen (HLA)-typing is performed to enable downstream major histocompatibility complex (MHC) binding prediction. c) Peptide enumeration occurs through translation of variant nucleotide sequences into their respective amino acid sequences, filtering for translation incompatible sequences such as those containing intervening stop codons or those with low evidence of RNA expression. These polypeptides are then used to derive 8–11 mer sequences (for MHC class I epitopes) or 15mer sequences (MHC class II epitopes) to allow for d) downstream MHC or HLA binding prediction of each sequence. Binders are typically defined in the literature as those with predicted binding affinity (Kd) of ≤ 500 nM or are selected from those with the highest rank percentile for predicted binding affinity. Other filtering criteria may be performed after this step, such as immunogenicity prediction or filtering away sequences with high homology to self-antigens. e) Lastly, therapies are generated using predicted tumour-specific antigens. These can be either DNA, RNA, or peptide vaccines or cellular therapies such as adoptive T-cell (ACT) therapy.
Figure 3:. Computational workflow for tumour-specific antigen…
Figure 3:. Computational workflow for tumour-specific antigen calling.
a) Identification of tumour-specific antigens begins with variant calling. This can be done through comparison of tumour versus normal tissue DNA sequences (single nucleotide variants (SNVs) and insertions or deletions (INDELs)) or RNA sequences (splice variants, fusions, viral sequences and retroelements) to look for tumour-specific variants in the exome or tumour-specific transcripts in the transcriptome, respectively. b) Tumour human leukocyte antigen (HLA)-typing is performed to enable downstream major histocompatibility complex (MHC) binding prediction. c) Peptide enumeration occurs through translation of variant nucleotide sequences into their respective amino acid sequences, filtering for translation incompatible sequences such as those containing intervening stop codons or those with low evidence of RNA expression. These polypeptides are then used to derive 8–11 mer sequences (for MHC class I epitopes) or 15mer sequences (MHC class II epitopes) to allow for d) downstream MHC or HLA binding prediction of each sequence. Binders are typically defined in the literature as those with predicted binding affinity (Kd) of ≤ 500 nM or are selected from those with the highest rank percentile for predicted binding affinity. Other filtering criteria may be performed after this step, such as immunogenicity prediction or filtering away sequences with high homology to self-antigens. e) Lastly, therapies are generated using predicted tumour-specific antigens. These can be either DNA, RNA, or peptide vaccines or cellular therapies such as adoptive T-cell (ACT) therapy.

References

    1. Yarchoan M, Johnson BA, Lutz ER, Laheru DA & Jaffee EM Targeting neoantigens to augment antitumour immunity. Nature Reviews Cancer 17, 209–222 (2017).
    1. Sahin U. et al. Personalized RNA mutanome vaccines mobilize poly-specific therapeutic immunity against cancer. Nature 547, 222–226 (2017).
    1. Ott PA et al. An immunogenic personal neoantigen vaccine for patients with melanoma. Nature 547, 217–221 (2017).
    1. Gubin MM, Artyomov MN, Mardis ER & Schreiber RD Tumor neoantigens: Building a framework for personalized cancer immunotherapy. Journal of Clinical Investigation 125, 3413–3421 (2015).
    1. Hacohen N, Fritsch EF, Carter TA, Lander ES & Wu CJ Getting Personal with Neoantigen-Based Therapeutic Cancer Vaccines. Cancer Immunol. Res 1, 11–15 (2013).
    1. Schumacher TN & Schreiber RD Neoantigens in cancer immunotherapy. Science. 348, 69–74 (2015).
    1. Cristescu R. et al. Pan-tumor genomic biomarkers for PD-1 checkpoint blockade-based immunotherapy. Science. 362, (2018).
    1. Keskin DB et al. Neoantigen vaccine generates intratumoral T cell responses in phase Ib glioblastoma trial. Nature 565, 234–239 (2019).
    1. Hilf N. et al. Actively personalized vaccination trial for newly diagnosed glioblastoma. Nature 565, 240–245 (2019).
    1. Turajlic S. et al. Insertion-and-deletion-derived tumour-specific neoantigens and the immunogenic phenotype: a pan-cancer analysis. Lancet Oncol 18, 1009–1021 (2017).
    1. Smith CC et al. Endogenous retroviral signatures predict immunotherapy response in clear cell renal cell carcinoma. J. Clin. Invest 128, 4804–4820 (2018).
    1. Thorsson V. et al. The Immune Landscape of Cancer. Immunity 48, 812–830.e14 (2018).
    1. Mertens F, Antonescu CR & Mitelman F. Gene fusions in soft tissue tumors: Recurrent and overlapping pathogenetic themes. Genes Chromosom. Cancer 55, 291–310 (2016).
    1. Wang Y, Wu N, Liu D. & Jin Y. Recurrent Fusion Genes in Leukemia: An Attractive Target for Diagnosis and Treatment. Curr. Genomics 18, (2017).
    1. Pellagatti A. et al. Impact of spliceosome mutations on RNA splicing in myelodysplasia: Dysregulated genes/pathways and clinical associations. Blood 132, 1225–1240 (2018).
    1. Bartel F, Taubert H. & Harris LC Alternative and aberrant splicing of MDM2 mRNA in human cancer. Cancer Cell 2, 9–15 (2002).
    1. Perz JF, Armstrong GL, Farrington LA, Hutin YJF & Bell BP The contributions of hepatitis B virus and hepatitis C virus infections to cirrhosis and primary liver cancer worldwide. J. Hepatol 45, 529–538 (2006).
    1. Ambrosio MR & Leoncini L. in Tropical Hemato-Oncology (eds. Droz J-P, Carme B, Couppié P, Nacher M. & Thiéblemont C.) 127–141 (Springer International Publishing, 2015). doi:10.1007/978-3-319-18257-5_15
    1. Mahieux R. & Gessain A. HTLV-1 and Associated Adult T-cell Leukemia/Lymphoma. Reviews in Clinical and Experimental Hematology 7, 336–361 (2003).
    1. Mesri EA, Cesarman E. & Boshoff C. Kaposi’s sarcoma herpesvirus/ Human herpesvirus-8 (KSHV/HHV8), and the oncogenesis of Kaposi’s sarcoma. Nat. Rev. Cancer 10, 707–719 (2010).
    1. Harrington WJ, Wood C. & Wood C. in DNA Tumor Viruses 683–702 (2009). doi:10.1007/978-0-387-68945-6_26
    1. Shukla SA et al. Comprehensive analysis of cancer-associated somatic mutations in class i HLA genes. Nat. Biotechnol 33, 1152–1158 (2015).
    1. Szolek A. et al. OptiType: Precision HLA typing from next-generation sequencing data. Bioinformatics 30, 3310–3316 (2014).
    1. Bai Y, Wang D. & Fury W. in Methods in Molecular Biology 1802, 193–201 (2018).
    1. Ka S. et al. HLAscan: Genotyping of the HLA region using next-generation sequencing data. BMC Bioinformatics 18, (2017).
    1. Buchkovich ML et al. HLAProfiler utilizes k-mer profiles to improve HLA calling accuracy for rare and common alleles in RNA-seq data. Genome Med 9, (2017).
    1. Jurtz V. et al. NetMHCpan-4.0: Improved Peptide–MHC Class I Interaction Predictions Integrating Eluted Ligand and Peptide Binding Affinity Data. J. Immunol 199, 3360–3368 (2017).
    1. Rajasagi M. et al. Systematic identification of personal tumor-specific neoantigens in chronic lymphocytic leukemia. Blood 124, 453–462 (2014).
    1. Soria-Guerra RE, Nieto-Gomez R, Govea-Alonso DO & Rosales-Mendoza S. An overview of bioinformatics tools for epitope prediction: Implications on vaccine development. Journal of Biomedical Informatics 53, 405–414 (2015).
    1. Zhang Q. et al. Immune epitope database analysis resource (IEDB-AR). Nucleic Acids Res 36, (2008).
    1. Linnebacher M. et al. Frameshift peptide-derived T-cell epitopes: A source of novel tumor-specific antigens. Int. J. Cancer 93, 6–11 (2001).
    1. Thibodeau SN, Bren G. & Schaid D. Microsatellite instability in cancer of the proximal colon. Science. 260, 816–819 (1993).
    1. Ionov Y, Peinado MA, Malkhosyan S, Shibata D. & Perucho M. Ubiquitous somatic mutations in simple repeated sequences reveal a new mechanism for colonic carcinogenesis. Nature 363, 558–561 (1993).
    1. Sakurada K. et al. RIZ, the retinoblastoma protein interacting zinc finger gene, is mutated in genetically unstable cancers of the pancreas, stomach, and colorectum. Genes Chromosom. Cancer 30, 207–211 (2001).
    1. De Smedt L. et al. Microsatellite instable vs stable colon carcinomas: Analysis of tumour heterogeneity, inflammation and angiogenesis. Br. J. Cancer 113, 500–509 (2015).
    1. Dolcetti R. et al. High prevalence of activated intraepithelial cytotoxic T lymphocytes and increased neoplastic cell apoptosis in colorectal carcinomas with microsatellite instability. Am. J. Pathol 154, 1805–1813 (1999).
    1. Maby P. et al. Correlation between density of CD8 + T-cell infiltrate in microsatellite unstable colorectal cancers and frameshift mutations: A rationale for personalized immunotherapy. Cancer Res 75, 3446–3455 (2015).
    1. Tougeron D. et al. Tumor-infiltrating lymphocytes in colorectal cancers with microsatellite instability are correlated with the number and spectrum of frameshift mutations. Mod. Pathol 22, 1186–1195 (2009).
    1. Saeterdal I, Gjertsen MK, Straten P, Eriksen JA & Gaudernack G. A TGF betaRII frameshift-mutation-derived CTL epitope recognised by HLA-A2-restricted CD8+ T cells. Cancer Immunol. Immunother. 50, 469–76 (2001).
    1. Le DT et al. PD-1 Blockade in Tumors with Mismatch-Repair Deficiency. N. Engl. J. Med 372, 2509–2520 (2015).
    1. Gong J, Chehrazi-Raffle A, Reddi S. & Salgia R. Development of PD-1 and PD-L1 inhibitors as a form of cancer immunotherapy: A comprehensive review of registration trials and future considerations. Journal for ImmunoTherapy of Cancer 6, (2018).
    1. Motzer RJ et al. Nivolumab versus Everolimus in Advanced Renal-Cell Carcinoma. N. Engl. J. Med 373, 1803–13 (2015).
    1. Hundal J. et al. pVAC-Seq: A genome-guided in silico approach to identifying tumor neoantigens. Genome Med 8, (2016).
    1. Kim S. et al. Neopepsee: Accurate genome-level prediction of neoantigens by harnessing sequence and amino acid immunogenicity information. Ann. Oncol 29, 1030–1036 (2018).
    1. Bjerregaard AM, Nielsen M, Hadrup SR, Szallasi Z. & Eklund AC MuPeXI: prediction of neo-epitopes from tumor sequencing data. Cancer Immunol. Immunother. 66, 1123–1130 (2017).
    1. Rubinsteyn A. et al. Computational pipeline for the PGV-001 neoantigen vaccine trial. Front. Immunol 8 (2018).
    1. Rech AJ et al. Tumor Immunity and Survival as a Function of Alternative Neopeptides in Human Cancer. Cancer Immunol. Res 6, 276–287 (2018).
    1. Zhou Z. et al. TSNAD: An integrated software for cancer somatic mutation and tumour-specific neoantigen detection. R. Soc. Open Sci 4, (2017).
    1. Saunders CT et al. Strelka: Accurate somatic small-variant calling from sequenced tumor-normal sample pairs. Bioinformatics 28, 1811–1817 (2012).
    1. Saeterdal I. et al. Frameshift-mutation-derived peptides as tumor-specific antigens in inherited and spontaneous colorectal cancer. Proc. Natl. Acad. Sci 98, 13255–13260 (2001).
    1. Inderberg EM et al. T cell therapy targeting a public neoantigen in microsatellite instable colon cancer reduces in vivo tumor growth. Oncoimmunology 6, (2017).
    1. Jayasinghe RG et al. Systematic Analysis of Splice-Site-Creating Mutations in Cancer. Cell Rep 23, 270–281.e3 (2018).
    1. Yang Y, Swaminathan S, Martin BK & Sharan SK Aberrant splicing induced by missense mutations in BRCA1: Clues from a humanized mouse model. Hum. Mol. Genet 12, 2121–2131 (2003).
    1. Nyström-Lahti M. et al. Missense and nonsense mutations in codon 659 of MLHI cause aberrant splicing of messenger RNA in HNPCC kindreds. Genes Chromosom. Cancer 26, 372–375 (1999).
    1. Zhang K, Nowak I, Rushlow D, Gallie BL & Lohmann DR Patterns of missplicing caused by RB1 gene mutations in patients with retinoblastoma and association with phenotypic expression. Hum. Mutat 29, 475–484 (2008).
    1. Wadt K. et al. A cryptic BAP1 splice mutation in a family with uveal and cutaneous melanoma, and paraganglioma. Pigment Cell Melanoma Res 25, 815–818 (2012).
    1. Chen LL et al. A mutation-created novel intra-exonic pre-mRNA splice site causes constitutive activation of KIT in human gastrointestinal stromal tumors. Oncogene 24, 4271–4280 (2005).
    1. Smart AC et al. Intron retention is a source of neoepitopes in cancer. Nat. Biotechnol 36, 1056–1058 (2018).
    1. Jung H. et al. Intron retention is a widespread mechanism of tumor-suppressor inactivation. Nat. Genet 47 1242–1248 (2015).
    1. Dvinge H. & Bradley RK Widespread intron retention diversifies most cancer transcriptomes. Genome Med 7, (2015).
    1. Kawakami SA et al. The intronic region of an incompletely spliced The lntronic Region of an Incompletely Spliced gp700 Gene Transcript Encodes an Epitope Recognized by Melanoma-Reactive Tumor-Infiltrating Lymphocytes. J Immunol J. Immunol. by guest April 159, 303–308 (1997).
    1. Coulie PG et al. A mutated intron sequence codes for an antigenic peptide recognized by cytolytic T lymphocytes on a human melanoma. Proc. Natl. Acad. Sci 92, 7976–7980 (1995).
    1. Uenaka A. et al. Cryptic CTL Epitope on a Murine Sarcoma Meth A Generated by Exon Extension as a Novel Mechanism. J. Immunol 170, 4862–4868 (2003).
    1. Boultwood J, Dolatshad H, Varanasi SS, Yip BH & Pellagatti A. The role of splicing factor mutations in the pathogenesis of the myelodysplastic syndromes. Advances in Biological Regulation 54, 153–161 (2014).
    1. Yip BH, Dolatshad H, Roy S, Pellagatti A. & Boultwood J. Impact of Splicing Factor Mutations on Pre-mRNA Splicing in the Myelodysplastic Syndromes. Curr. Pharm. Des 22, 2333–44 (2016).
    1. Weiss RB, Dunn DM, Atkins JF & Gesteland RF Slippery runs, shifty stops, backward steps, and forward hops: −2, −1, +1, +2, +5, and +6 ribosomal frameshifting. Cold Spring Harb. Symp. Quant. Biol 52, 687–93 (1987).
    1. Saulquin X. et al. +1 Frameshifting as a novel mechanism to generate a cryptic cytotoxic T lymphocyte epitope derived from human interleukin 10. J Exp Med 195, 353–358 (2002).
    1. Macejak DG & Sarnow P. Internal initiation of translation mediated by the 5′ leader of a cellular mRNA. Nature 353, 90–94 (1991).
    1. Bullock TNJ, Patterson AE, Franlin LL, Notidis E. & Eisenlohr LC Initiation Codon Scanthrough versus Termination Codon Readthrough Demonstrates Strong Potential for Major Histocompatibility Complex Class I–restricted Cryptic Epitope Expression. J. Exp. Med 186, 1051–1058 (1997).
    1. Bullock TN Ribosomal scanning past the primary initiation codon as a mechanism for expression of CTL epitopes encoded in alternative reading frames. J. Exp. Med 184, 1319–1329 (1996).
    1. Malarkannan S, Horng T, Shih PP, Schwab S. & Shastri N. Presentation of out-of-frame peptide/MHC class I complexes by a novel translation initiation mechanism. Immunity 10, 681–690 (1999).
    1. Eynde BBJ Van Den et al. A new antigen recognized by cytolytic T lymphocytes on a human kidney tumor results from reverse strand transcription. J. Exp. Med 190, 1793–1800 (1999).
    1. Bruce A, Atkins J. & Gesteland R. tRNA anticodon replacement experiments show that ribosomal frameshifting can be caused by doublet decoding. Proc Natl Acad Sci U S A 83, 5062–5066 (1986).
    1. Dalet A. et al. An antigenic peptide produced by reverse splicing and double asparagine deamidation. Proc. Natl. Acad. Sci 108, E323–E331 (2011).
    1. Hanada KI, Yewdell JW & Yang JC Immune recognition of a human renal cancer antigen through post-translational protein splicing. Nature 427, 252–256 (2004).
    1. Liepe J. et al. A large fraction of HLA class I ligands are proteasome-generated spliced peptides. Science. 354, 354–358 (2016).
    1. Kahles A. et al. Comprehensive Analysis of Alternative Splicing Across Tumors from 8,705 Patients. Cancer Cell 34, 211–224.e6 (2018).
    1. Shukla GC & Singh J. Mutations of RNA splicing factors in hematological malignancies. Cancer Lett 409, 1–8 (2017).
    1. Ley TJ et al. Genomic and Epigenomic Landscapes of Adult De Novo Acute Myeloid Leukemia. N. Engl. J. Med 368, 2059–2074 (2013).
    1. Adamia S. et al. AGenome-wide aberrantRNASplicing in patients with acute myeloid leukemia identifies novel potential disease markers and therapeutic targets. Clin. Cancer Res 20, 1135–1145 (2014).
    1. Wang L. et al. SF3B1 and Other Novel Cancer Genes in Chronic Lymphocytic Leukemia. N. Engl. J. Med 365, 2497–2506 (2011).
    1. Yoshida K. et al. Frequent pathway mutations of splicing machinery in myelodysplasia. Nature 478, 64–69 (2011).
    1. Kar SA et al. Spliceosomal gene mutations are frequent events in the diverse mutational spectrum of chronic myelomonocytic leukemia but largely absent in juvenile myelomonocytic leukemia. Haematologica 98, 107–113 (2013).
    1. Visconte V, Makishima H, Maclejewski JP & Tiu RV Emerging roles of the spliceosomal machinery in myelodysplastic syndromes and other hematological disorders. Leukemia 26, 2447–2454 (2012).
    1. Quesada V. et al. Exome sequencing identifies recurrent mutations of the splicing factor SF3B1 gene in chronic lymphocytic leukemia. Nat. Genet 44, 47–52 (2012).
    1. Lee SCW et al. Modulation of splicing catalysis for therapeutic targeting of leukemia with mutations in genes encoding spliceosomal proteins. Nat. Med 22, 672–678 (2016).
    1. Lim KH & Fairbrother WG Spliceman-A computational web server that predicts sequence variations in pre-mRNA splicing. Bioinformatics 28, 1031–1032 (2012).
    1. Mort M. et al. MutPred Splice: Machine learning-based prediction of exonic variants that disrupt splicing. Genome Biol 15, (2014).
    1. Brooks AN et al. Conservation of an RNA regulatory map between Drosophila and mammals. Genome Res 21, 193–202 (2011).
    1. Rogers MF, Thomas J, Reddy ASN & Ben-Hur A. SpliceGrapher: Detecting patterns of alternative splicing from RNA-Seq data in the context of gene models and EST data. Genome Biol 13, (2012).
    1. Shen S. et al. rMATS: Robust and flexible detection of differential alternative splicing from replicate RNA-Seq data. Proc. Natl. Acad. Sci 111, E5593–E5601 (2014).
    1. Kahles A, Ong CS, Zhong Y. & Rätsch G. SplAdder: Identification, quantification and testing of alternative splicing events from RNA-Seq data. Bioinformatics 32, 1840–1847 (2016).
    1. Denti L. et al. ASGAL: Aligning RNA-Seq data to a splicing graph to detect novel alternative splicing events. BMC Bioinformatics 19, (2018).
    1. US National Library of Medicine. , (2019).
    1. Shyu A. Bin, Wilkinson MF & Van Hoof A. Messenger RNA regulation: To translate or to degrade. EMBO Journal 27, 471–481 (2008).
    1. Crainie M, Belch AR, Mant MJ & Pilarski LM Overexpression of the receptor for hyaluronan-mediated motility (RHAMM) characterizes the malignant clone in multiple myeloma: identification of three distinct RHAMM variants. Blood 93, 1684–96 (1999).
    1. Busse A. et al. Wilms’ tumor gene 1 (WT1) expression in subtypes of acute lymphoblastic leukemia (ALL) of adults and impact on clinical outcome. Ann. Hematol 88, 1199–1205 (2009).
    1. Kramarzova K. et al. Real-time PCR quantification of major Wilms tumor gene 1 (WT1) isoforms in acute myeloid leukemia, their characteristic expression patterns and possible functional consequences. Leukemia 26, 2086–2095 (2012).
    1. Siehl JM et al. Expression of Wilms’ tumor gene 1 at different stages of acute myeloid leukemia and analysis of its major splice variants. Ann. Hematol 83, 745–750 (2004).
    1. Mailänder V. et al. Complete remission in a patient with recurrent acute myeloid leukemia induced by vaccination with WT1 peptide in the absence of hematological or renal toxicity [2]. Leukemia 18, 165–166 (2004).
    1. Kohrt HE et al. Donor immunization with WT1 peptide augments antileukemic activity after MHC-matched bone marrow transplantation. Blood 118, 5319–5329 (2011).
    1. Oka Y. et al. Wilms tumor gene peptide-based immunotherapy for patients with overt leukemia from myelodysplastic syndrome (MDS) or MDS with myelofibrosis. Int. J. Hematol 78, 56–61 (2003).
    1. Rosenfeld C, Cheever MA & Gaiger A. WT1 in acute leukemia, chronic myelogenous leukemia and myelodysplastic syndrome: Therapeutic potential of WT1 targeted therapies. Leukemia 17, 1301–1312 (2003).
    1. Chapuis AG et al. Transferred WT1-reactive CD8+ T cells can mediate antileukemic activity and persist in post-transplant patients. Sci. Transl. Med 5, (2013).
    1. Tsuboi A. et al. WT1 Peptide-Based Immunotherapy for Patients with Lung Cancer: Report of Two Cases. Microbiol. Immunol 48, 175–184 (2004).
    1. liyama T. et al. WT 1 (Wilms’ tumor 1) peptide immunotherapy for renal cell carcinoma. Microbiol. Immunol 51, 519–530 (2007).
    1. Kawase T. et al. Alternative splicing due to an intronic SNP in HMSD generates a novel minor histocompatibility antigen. Blood 110, 1055–1063 (2007).
    1. Vauchy C. et al. CD20 alternative splicing isoform generates immunogenic CD4 helper T epitopes. Int. J. Cancer 137, 116–126 (2015).
    1. Rowley JD A new consistent chromosomal abnormality in chronic myelogenous leukaemia identified by quinacrine fluorescence and Giemsa staining. Nature 243, 290–293 (1973).
    1. Williams SV, Hurst CD & Knowles MA Oncogenic FGFR3 gene fusions in bladder cancer. Hum. Mol. Genet 22, 795–803 (2013).
    1. Tognon C. et al. Expression of the ETV6-NTRK3 gene fusion as a primary event in human secretory breast carcinoma. Cancer Cell 2, 367–376 (2002).
    1. The Cancer Genome Atlas Network. Comprehensive molecular characterization of clear cell renal cell carcinoma. Nature 499, 43–9 (2013).
    1. Seshagiri S. et al. Recurrent R-spondin fusions in colon cancer. Nature 488, 660–664 (2012).
    1. Young LC et al. Identification of novel isoforms of the EML4-ALK transforming gene in non-small cell lung cancer. Cancer Res 68, 4971–4976 (2008).
    1. Lyu X. et al. Detection of 22 common leukemic fusion genes using a single-step multiplex qRT-PCR-based assay. Diagn. Pathol 12, 55 (2017).
    1. Xiao X, Garbutt CC, Hornicek F, Guo Z. & Duan Z. Advances in chromosomal translocations and fusion genes in sarcomas and potential therapeutic applications. Cancer Treatment Reviews 63, 61–70 (2018).
    1. Translocations C. et al. Antigenicity of Fusion Proteins from Sarcoma-associated. Cytokines 61, 6868–6875 (2001).
    1. Druker BJ et al. Efficacy and Safety of a Specific Inhibitor of the BCR-ABL Tyrosine Kinase in Chronic Myeloid Leukemia. N. Engl. J. Med 344, 1031–1037 (2001).
    1. Jamal-Hanjani M. et al. Tracking the Evolution of Non–Small-Cell Lung Cancer. N. Engl. J. Med 376, 2109–2121 (2017).
    1. McGranahan N. et al. Allele-Specific HLA Loss and Immune Escape in Lung Cancer Evolution. Cell 171 1259–1271.e11 (2017).
    1. Rosenthal R. et al. Neoantigen-directed immune escape in lung cancer evolution. Nature 567, 479–485 (2019).
    1. Yu YP et al. Identification of recurrent fusion genes across multiple cancer types. Sci. Rep 9, 1074 (2019).
    1. Wang Q, Xia J, Jia P, Pao W. & Zhao Z. Application of next generation sequencing to human gene fusion detection: Computational tools, features and perspectives. Brief. Bioinform 14, 506–519 (2013).
    1. Zhang J, Mardis ER & Maher CA INTEGRATE-neo: A pipeline for personalized gene fusion neoantigen discovery. Bioinformatics 33, 555–557 (2017).
    1. Chang TC et al. The neoepitope landscape in pediatric cancers. Genome Med 9, (2017).
    1. Pinilla-Ibarz J. et al. Vaccination of patients with chronic myelogenous leukemia with bcr-abl oncogene breakpoint fusion peptides generates specific immune responses. Blood 95, 1781–1787 (2000).
    1. Cathcart K. et al. A multivalent bcr-abl fusion peptide vaccination trial in patients with chronic myeloid leukemia. Blood 103, 1037–1042 (2004).
    1. Mackall CL et al. A pilot study of consolidative immunotherapy in patients with high-risk pediatric sarcomas. Clin. Cancer Res 14, 4850 (2008).
    1. Bocchia M. et al. Effect of a p210 multipeptide vaccine associated with imatinib or interferon in patients with chronic myeloid leukaemia and persistent residual disease: A multicentre observational trial. Lancet 365, 657–662 (2005).
    1. Rojas JM, Knight K, Wang L. & Clark RE Clinical evaluation of BCR-ABL peptide immunisation in chronic myeloid leukaemia: Results of the EPIC study. Leukemia 21, 2287–2295 (2007).
    1. Yang W. et al. Immunogenic neoantigens derived from gene fusions stimulate T cell responses. Nat. Med 25, 767–775 (2019).
    1. Goodier JL & Kazazian HH Retrotransposons Revisited: The Restraint and Rehabilitation of Parasites. Cell 135, 23–35 (2008).
    1. Shen H. et al. Integrated Molecular Characterization of Testicular Germ Cell Tumors. Cell Rep 23, 3392–3406 (2018).
    1. Florl AR, Löwer R, Schmitz-Dräger BJ & Schulz WA DNA methylation and expression of LINE-1 and HERV-K provirus sequences in urothelial and renal cell carcinomas. Br. J. Cancer 80, 1312–1321 (1999).
    1. Brocks D. et al. DNMT and HDAC inhibitors induce cryptic transcription start sites encoded in long terminal repeats. Nat. Genet 49, 1052–1060 (2017).
    1. Chiappinelli KB et al. Inhibiting DNA Methylation Causes an Interferon Response in Cancer via dsRNA Including Endogenous Retroviruses. Cell 162, 974–986 (2015).
    1. Sheng W. et al. LSD1 Ablation Stimulates Anti-tumor Immunity and Enables Checkpoint Blockade. Cell 174, 549–563.e19 (2018).
    1. Goel S. et al. CDK4/6 inhibition triggers anti-tumour immunity. Nature 548, 471–475 (2017).
    1. Jones PA, Ohtani H, Chakravarthy A. & De Carvalho DD Epigenetic therapy in immune-oncology. Nature Reviews Cancer 19, 151–161 (2019).
    1. Belgnaoui SM, Gosden RG, Semmes OJ & Haoudi A. Human LINE-1 retrotransposon induces DNA damage and apoptosis in cancer cells. Cancer Cell Int 6, (2006).
    1. Scott EC et al. A hot L1 retrotransposon evades somatic repression and initiates human colorectal cancer. Genome Res 26, 745–755 (2016).
    1. Chen L, Dahlstrom JE, Chandra A, Board P. & Rangasamy D. Prognostic value of LINE-1 retrotransposon expression and its subcellular localization in breast cancer. Breast Cancer Res. Treat 136, 129–142 (2012).
    1. Patnala R. et al. Inhibition of LINE-1 retrotransposon-encoded reverse transcriptase modulates the expression of cell differentiation genes in breast cancer cells. Breast Cancer Res. Treat 143, 239–253 (2014).
    1. Löwer R, Löwer J. & Kurth R. The viruses in all of us: characteristics and biological significance of human endogenous retrovirus sequences. Proc. Natl. Acad. Sci 93, 5177–5184 (1996).
    1. Boller K. et al. Human endogenous retrovirus HERV-K113 is capable of producing intact viral particles. J. Gen. Virol 89, 567–572 (2008).
    1. Faff O. et al. Retrovirus-like particles from the human T47D cell line are related to mouse mammary tumour virus and are of human endogenous origin. J. Gen. Virol 73, 1087–1097 (1992).
    1. Wang-Johanning F. et al. Expression of multiple human endogenous retrovirus surface envelope proteins in ovarian cancer. Int. J. Cancer 120, 81–90 (2007).
    1. Büscher K. et al. Expression of human endogenous retrovirus K in melanomas and melanoma cell lines. Cancer Res 65, 4172–4180 (2005).
    1. Wang-Johanning F. et al. Expression of human endogenous retrovirus K envelope transcripts in human breast cancer. Clin. Cancer Res 7, 1553–1560 (2001).
    1. Contreras-Galindo R. et al. Human Endogenous Retrovirus K (HML-2) Elements in the Plasma of People with Lymphoma and Breast Cancer. J. Virol 82, 9329–9336 (2008).
    1. Wang-Johanning F. et al. Detecting the expression of human endogenous retrovirus E envelope transcripts in human prostate adenocarcinoma. Cancer 98, 187–197 (2003).
    1. Yoshida M, Miyoshi I. & Hinuma Y. Isolation and characterization of retrovirus from cell lines of human adult T-cell leukemia and its implication in the disease. Proc. Natl. Acad. Sci 79, 2031–2035 (1982).
    1. Kalyanaraman VS et al. A new subtype of human T-cell leukemia virus (HTLV-II) associated with a T-cell variant of hairy cell leukemia. Science. 218, 571–573 (1982).
    1. Sauter M. et al. Human endogenous retrovirus K10: expression of Gag protein and detection of antibodies in patients with seminomas. J. Virol 69, 414–21 (1995).
    1. Cherkasova E. et al. Detection of an immunogenic HERV-E envelope with selective expression in clear cell kidney cancer. Cancer Res 76, 2177–2185 (2016).
    1. Takahashi Y. et al. Regression of human kidney cancer following allogeneic stem cell transplantation is associated with recognition of an HERV-E antigen by T cells. J. Clin. Invest 118, 1099–109 (2008).
    1. Panda A. et al. Endogenous retrovirus expression is associated with response to immune checkpoint pathway in clear cell renal cell carcinoma. JCI Insight 3, (2018).
    1. Rooney MS, Shukla SA, Wu CJ, Getz G. & Hacohen N. Molecular and genetic properties of tumors associated with local immune cytolytic activity. Cell 160, 48–61 (2015).
    1. Mayer J, Blomberg J. & Seal RL A revised nomenclature for transcribed human endogenous retroviral loci. Mob. DNA 2, (2011).
    1. Cherkasova E. et al. Inactivation of the von Hippel-Lindau tumor suppressor leads to selective expression of a human endogenous retrovirus in kidney cancer. Oncogene 30, 4697–4706 (2011).
    1. Vargiu L. et al. Classification and characterization of human endogenous retroviruses; mosaic forms are common. Retrovirology 13, 7 (2016).
    1. Tokuyama M. et al. ERVmap analysis reveals genome-wide transcription of human endogenous retroviruses. Proc. Natl. Acad. Sci 115, 12565–12572 (2018).
    1. Smit A, Hubley R. & Green P. RepeatMasker Open-4.0. 2013–2015. (2013).
    1. Paces J. HERVd: the Human Endogenous RetroViruses Database: update. Nucleic Acids Res 32, 50D–50 (2004).
    1. Kim TH, Jeon YJ, Kim WY & Kim HS HESAS: HERVs expression and structure analysis system. Bioinformatics 21, 1699–1700 (2005).
    1. Tongyoo P. et al. EnHERV: Enrichment analysis of specific human endogenous retrovirus patterns and their neighboring genes. PLoS One 12, (2017).
    1. US National Library of Medicine. , (2019).
    1. Brandle D. A mutated HLA-A2 molecule recognized by autologous cytotoxic T lymphocytes on a human renal cell carcinoma. J. Exp. Med 183, 2501–2508 (1996).
    1. Huang J. et al. T Cells Associated with Tumor Regression Recognize Frameshifted Products of the CDKN2A Tumor Suppressor Gene Locus and a Mutated HLA Class I Gene Product. J. Immunol 172, 6057–6064 (2014).
    1. Van Hall T. et al. Selective cytotoxic T-lymphocyte targeting of tumor immune escape variants. Nat. Med 12, 417–424 (2006).
    1. Doorduijn EM et al. TAP-independent self-peptides enhance T cell recognition of immune-escaped tumors. J. Clin. Invest 126, 784–794 (2016).
    1. Marijt KA, Doorduijn EM & van Hall T. TEIPP antigens for T-cell based immunotherapy of immuneedited HLA class Ilow cancers. Mol. Immunol (2018). doi:10.1016/j.molimm.2018.03.029
    1. Doorduijn EM et al. T cells specific for a TAP-independent self-peptide remain naïve in tumor-bearing mice and are fully exploitable for therapy. Oncoimmunology 7, (2018).
    1. Marijt KA et al. Identification of non-mutated neoantigens presented by TAP-deficient tumors. J. Exp. Med 215, 2325–2337 (2018).
    1. Lansford JL et al. Computational modeling and confirmation of leukemia-associated minor histocompatibility antigens. Blood Adv 2, 2052–2062 (2018).
    1. Kreiter S. et al. Mutant MHC class II epitopes drive therapeutic immune responses to cancer. Nature 520, 692–696 (2015).
    1. Tran E. et al. Cancer immunotherapy based on mutation-specific CD4+ T cells in a patient with epithelial cancer. Science. 344, 641–645 (2014).
    1. Andreatta M. et al. Machine learning reveals a non-canonical mode of peptide binding to MHC class II molecules. Immunology 152, 255–264 (2017).
    1. Nielsen M. & Lund O. NN-align. An artificial neural network-based alignment algorithm for MHC class II peptide binding prediction. BMC Bioinformatics 10, 296 (2009).
    1. Andreatta M. et al. Accurate pan-specific prediction of peptide-MHC class II binding affinity with improved binding core identification. Immunogenetics 67, 641–650 (2015).
    1. Saito R. et al. Molecular subtype-specific immunocompetent models of high-grade urothelial carcinoma reveal differential neoantigen expression and response to immunotherapy. Cancer Res 78, 3954–3968 (2018).
    1. The problem with neoantigen prediction. Nat. Biotechnol 35, 97–97 (2017).
    1. Pearson H. et al. MHC class I-associated peptides derive from selective regions of the human genome. J. Clin. Invest 126, 4690–4701 (2016).
    1. Creech AL et al. The Role of Mass Spectrometry and Proteogenomics in the Advancement of HLA Epitope Prediction. Proteomics 18, (2018).
    1. Hunt DF et al. Characterization of peptides bound to the class I MHC molecule HLA-A2.1 by mass spectrometry. Science. 255, 1261–1263 (1992).
    1. Falk K, Rötzschke O, Stevanovié S, Jung G. & Rammensee HG Allele-specific motifs revealed by sequencing of self-peptides eluted from MHC molecules. Nature 351, 290–296 (1991).
    1. Michalski A, Cox J. & Mann M. More than 100,000 detectable peptide species elute in single shotgun proteomics runs but the majority is inaccessible to data-dependent LC-MS/MS. J. Proteome Res 10, 1785–1793 (2011).
    1. Griss J. et al. Recognizing millions of consistently unidentified spectra across hundreds of shotgun proteomics datasets. Nat. Methods 13, 651–656 (2016).
    1. Yaqüe J. et al. Peptide rearrangement during quadrupole ion trap fragmentation: Added complexity to MS/MS spectra. Anal. Chem 75, 1524–1535 (2003).
    1. Chawner R, Holman SW, Gaskell SJ & Eyers CE Peptide scrambling during collision-induced dissociation is influenced by n-terminal residue basicity. J. Am. Soc. Mass Spectrom 25, 1927–1938 (2014).
    1. Yadav M. et al. Predicting immunogenic tumour mutations by combining mass spectrometry and exome sequencing. Nature 515, 572–576 (2014).
    1. Polyakova A, Kuznetsova K. & Moshkovskii S. Proteogenomics meets cancer immunology: Mass spectrometric discovery and analysis of neoantigens. Expert Review of Proteomics 12, 533–541 (2015).
    1. Laumont CM et al. Noncoding regions are the main source of targetable tumor-specific antigens. Sci. Transl. Med 10, eaau5516 (2018).
    1. van der Lee DI et al. Mutated nucleophosmin 1 as immunotherapy target in acute myeloid leukemia. J. Clin. Invest 129, 774–785 (2019).
    1. Matsushita H. et al. Cancer exome analysis reveals a T-cell-dependent mechanism of cancer immunoediting. Nature 482, 400–404 (2012).
    1. Castle JC et al. Exploiting the mutanome for tumor vaccination. Cancer Res 72, 1081–1091 (2012).
    1. Gubin MM et al. Checkpoint blockade cancer immunotherapy targets tumour-specific mutant antigens. Nature 515, 577–81 (2014).
    1. Carreno BM et al. A dendritic cell vaccine increases the breadth and diversity of melanoma neoantigen-specific T cells. Science. 348, 803–808 (2015).
    1. Gao Q. et al. Driver Fusions and Their Implications in the Development and Treatment of Human Cancers. Cell Rep 23, 227–238.e3 (2018).
    1. Selitsky SR, Marron D, Mose LE, Parker JS & Dittmer DP Epstein-Barr Virus-Positive Cancers Show Altered B-Cell Clonality. mSystems 3, (2018).

Source: PubMed

3
Subskrybuj