Open Access

Integrated analysis of whole genome and transcriptome sequencing in a young patient with gastric cancer provides insights for precision therapy

  • Authors:
    • Kongwang Hu
    • Weiqiang Yu
    • Olugbenga Emmanuel Ajayi
    • Longlong Li
    • Zhiguo Huang
    • Qiqi Rong
    • Shuaili Wang
    • Qing-Fa Wu
  • View Affiliations

  • Published online on: August 12, 2020     https://doi.org/10.3892/ol.2020.11976
  • Article Number: 115
  • Copyright: © Hu et al. This is an open access article distributed under the terms of Creative Commons Attribution License.

Metrics: Total Views: 0 (Spandidos Publications: | PMC Statistics: )
Total PDF Downloads: 0 (Spandidos Publications: | PMC Statistics: )


Abstract

Gastric cancer is a leading cause of cancer associated deaths worldwide and is considered to be an age‑related disease. In younger patients, gastric cancer is biologically more aggressive, and prognosis is worse compared with that in elderly patients. In the present case report, the whole genome and transcriptome was sequenced in a 26‑year‑old patient with gastric cancer who presented with gastric cancer‑related symptoms and was admitted to the First Affiliated Anhui Medical Hospital (Hefei, China) in December 2016. In total, 9 germline and 4 somatic mutations were identified in the patient, and there were more deleterious sites in the germline mutated genes. Genes with somatic mutations, such as MUC2, MUC4, SLC8A2, and with structural variations, including CCND3, FGFR2 and FGFR3, were found to be differentially expressed. Cancer‑associated pathways, such as the ‘calcium signaling pathway’, ‘cGMP‑PKG signaling pathway’ and ‘transcriptional mis‑regulation’ were also enriched at both the genomic and transcriptomic levels. The genes found to have germline (SFRP4), somatic (MUC2, MUC4, SLC8A2) mutations, or structural variations (CCND3, FGFR2 and FGFR3) were differentially expressed in the patient and could be promising precision therapy targets.

Introduction

Gastric cancer (GC) is a leading cause of cancer-related deaths worldwide and is known to be an age-related disease (1). The mean age of patients with GC at diagnosis is ~60 years and <3% of GC cases are reported in patients <30 years of age (2,3). Several reports have found that younger patients (~30 years) are often diagnosed with advanced stages of GC and have worse prognosis compared with that in elderly patients (~60 years) (4,5). In younger patients, the cancer was found to spread rapidly and was biologically more aggressive (6). Integrative genomic approaches, which associate genomic and transcriptomic analysis have been found to increase the identification of numerous events and processes associated with tumor aggressiveness and may assist in selecting candidate genes from normal and pathological samples for targeted therapy (7). A previous study aimed to investigate the biological and genomic mechanisms in the progression of GC has led to the development of target-oriented therapy for advanced GC to be used in a clinical setting, with base substitution mutations in genes such as tumor protein p53 (TP53), AT-rich interaction domain 1A (ARID1A), KRAS proto-oncogene, GTPase (KRAS), phosphatidylinositol-4,5-bisphosphate 3-kinase catalytic subunit alpha (PIK3CA), ring finger protein 43 (RNF43), APC regulator of WNT signaling pathway (APC), RAS p21 protein activator 1 (RASA1) and erb-b2 receptor tyrosine kinase 2 (ERBB2) identified as important therapeutic targets in anti-cancer treatments (8). Developing personalized therapy for optimal individual cancer patient outcome has become more feasible due to the rapid advancement in next-generation sequencing techniques and technologies that enable fast and comprehensive characterizations of tumors at the molecular level (912). As every patient harbors a unique combination of variants that influence the risk, onset, and progression of the disease, an effective personalized therapy is dependent on a well-profiled genome from the individual patient with cancer and understanding the oncogenic mechanisms that regulate the progression of the tumor (13). With a view to identifying potential clinically actionable therapeutic targets that may inform individualized treatment strategies, the whole genome was sequenced in a 26-year-old patient with advanced stage GC, by integrating the whole genome and whole transcriptome sequencing data.

Materials and methods

Sample collection and preservation

A 26-year-old female patient was admitted to the First Affiliated Anhui Medical Hospital in December 2016 having presented with GC-related symptoms. The patient felt uncomfortable in the upper left quadrant of her abdomen and had repeated black melena for one month. This symptom recurred and became more serious for 2 weeks. The patient felt full and uncomfortable in the upper left quadrant of her abdomen repeatedly following a meal, one month prior to the development of black melena. She also had nausea, although there was no hematemesis or hematochezia. The abdominal discomfort worsened and the patient received treatment in Luan Jinkai Hospital. Thereafter, the patient visited the First Affiliated Hospital of Anhui Medical University where the present study was conducted. The gastroscopy of the patient indicated GC with obstruction, while pathology indicated antrum considered as carcinoma mucocellulare. Pathological examination of the samples was performed using hematoxylin and eosin staining. Briefly, tissues were sliced with a dimension of 1.5×1.5×0.2-0.3 cm and soaked in 40°C warm water. The sections were deparaffinized in xylene and were gradually hydrated through graded ethanol (100, 95, 80 and 70%) each for 3 min at 25°C. The sections were stained in hematoxylin solution for 10 min at 25°C and differentiated in 1% hydrochloric alcohol. Then the slides were rinsed with tap water and distilled water until the nuclei became blue, and dehydrated in 95% ethanol. The sections were counterstained in 1% eosin solution for 5 min at 25°C, washed with 70% ethanol twice and absolute ethanol, and were cleared in two washes of xylene. The sections were mounted with neutral balsam and observed under a light microscope magnification, ×400 (×40 objective and ×10 ocular magnification; BX-42; Olympus). An abdominal computer tomography scan + enhancement showed that both armpits had multiple small lymph nodes and the gastric antrum had thickened walls along with multiple swollen lymph nodes in the surrounding area. The patient was diagnosed with a T3N3bM0 (advanced stage IIIC) GC using AJCC TNM staging system 7th edition (14).

The patient underwent a gastrostomy for tumor removal. Tumor tissue and adjacent normal tissue ~5-cm away from the tumor were resected and preserved in liquid nitrogen until use.

DNA and RNA extraction and library preparation

Genomic DNA was isolated from the resected tissues using a genomic DNA isolation kit (Bioo Scientific; http://www.biooscientific.com) according to the manufacturer's recommendations. Extracted DNA was quantified using a NanoDrop ND-1000 spectrophotometer (NanoDrop Technologies; Thermo Fisher Scientific, Inc.) and the integrity was assessed using 1% agarose gel electrophoresis and visualized in gel imager (Tanon; http://www.biotanon.com). RNA was extracted from the tumor and adjacent normal tissue using TRIzol® (Invitrogen; Thermo Fisher Scientific, Inc.), and eluted in RNAse-free water. RNA quantity and quality were assessed using an Agilent 2100 Bioanalyzer (Agilent Technologies Inc.). DNA and RNA libraries were prepared according to standard protocols according to Bioo Scientific.

Whole transcriptome sequencing and pre-analysis

For whole transcriptome sequencing, raw sequencing reads were produced using an Illumina X10 sequencer (Illumina Inc.) with 150-bp paired-end reads according to the manufacturer's instructions. Raw RNA-Sequencing (RNA-Seq) data was trimmed by removing the adaptors and low quality reads using trim_galore v0.4.4 software (15). The data was subsequently aligned to the human reference genome GRCh37.p13 using STAR v2.6.1a software (16) in a 2-pass mapping. Read count in the tumor and adjacent normal tissues was calculated using htseq-count v0.10.0 software (17), and differentially expressed genes were detected using edgeR v3.24.3 software (18). Genes with P≤0.001 and log2 fold change ≥1 were considered as differentially expressed.

Whole genome sequencing and pre-analysis

Whole genome sequencing of the prepared libraries was performed using the Illumina X10 sequencer (Illumina Inc.) with 150-bp paired-end reads according to the manufacturer's instructions. The raw fastq data was mapped to the human reference genome (b37) provided by the Broad Institute with the Burrows-Wheeler Aligner v.7.0.12 (19). Read duplications were marked using the Picard tool v.2.10.10; local realignment and base quality score recalibration was performed using Genome Analysis ToolKit (GATK) v. 3.8-0 (20) and the final bam files were used for variant calling.

Somatic and germline variants detection

Germline variants were called using the GATK HaplotyperCaller joint v3.8 software (21) and then filtered using GATK variant quality score recalibration (VQSR). Variants that passed using the VQSR module with coverage ≥6X and supporting reads ≥0 were retained, while the variants showing genotype ‘0/0’ or ‘./.’ in the tumor sample were filtered out. For somatic variant calling, three different types of software were used: GATK HaplotypeCaller joint v3.8, Mutect v1.1.4 and MuTect2 v3.8 (22). In Mutect and MuTect2, all variants that were flagged as ‘PASS’ were kept. Other variants were retained with a coverage ≥5X and supporting ≥3 reads, if they were flagged among a subset of ‘alternative (23) allele in normal’, ‘triallelic site’, ‘possible contamination’, ‘clustered read position’ in Mutect, or ‘germline risk’, ‘alt allele in normal’, ‘t lod fstar’ (tumor does not meet likelihood threshold), ‘str contraction’ (site filtered due to contraction of short tandem repeat region), ‘triallelic site’ (site filtered because > two alternate alleles pass tumor likelihood threshold) in MuTect2. Germline variants were removed from both the Mutect and MuTect2 output data. In all three softwares, variant allele frequency in the tumor sample had to be 3 times higher compared with that in the normal sample, and >5% of the variant reads in the tumor sample was required. Functional annotations for both germline and somatic variants were added to each mutation using the ANNOVAR software v.2018Apr16 (17) and several publicly available databases, such as 1,000 Genome Project (24), the Exome Aggregation Consortium (25), the Genome Aggregation Database (26) and the compiled scores prediction system dbnsfp33a (27), dbSNP (28) and COSMIC (29).

Transition to transversion ratio calculation

The transition to transversion ratio for germline variants was calculated by dividing total transitions by total transversions. Transitions refer to variations that involve a change from purine to purine, while transversions refer to variations that involve a change from purine to pyrimidine or vice versa (30). The transition to transversion ratio between homologous strands of DNA is generally ~2, and for human, the ratio is ~2.1 (31).

Copy number alternation and structural variation analysis

Copy number variants (CNV), tumor purity and ploidy was analyzed using cnv_facets software (https://github.com/dariober/cnv_facets) based on FACETS v0.5.14 (32), and run with the ‘-cval 25 400’ command. Duplicated and deleted segments were retained and annotated using an ENCODE gene symbol in R v3.5.1 (33). iCallSV (https://github.com/rhshah/iCallSV), which applied Delly v0.7.5 prediction method (34) was used to detect structural variations and fusions using default parameters. Inter- and intra-chromosomal fusions >50 kb were retained and only in-frame deletions, duplications and fusions were selected for subsequent analysis.

Mutation significance analysis

The 1,000 Genome Project (24), the Exome Aggregation Consortium (25) and the Genome Aggregation Database (26) provided alternative allele frequency data in different populations that ANNOVAR software applied to the present variations. Annotations using a set of algorithm scoring system from dbNSFP v.33a, including FATHMM, LRT, Mutation Assessor, Mutation Taster, M-CAP, Polyphen2, PROVEAN and SIFT (35), were also applied to the variations. Then the common variants were filtered out, which are sites with a minor allele frequency ≥1%, to get the most informative ones. Mutations predicted to be deleterious by at least two algorithms of dbNSFP were considered to be potential pathogenic sites of these informative variants.

Identification of mutations in cancer associated genes

To identify genes with a high possibility of involvement in cancer development, the Cancer Gene Census (CGC) (https://cancer.sanger.ac.uk/census) (36) and Cancer Predisposition Genes (CPG) (37) databases were used. In total, 850 candidate genes from cancer-associated databases (723 genes from CGC and 114 genes from CPG with 87 overlapping genes), were investigated. The gene lists were downloaded directly from the two databases with no specific selection criteria.

Pathway analysis

Gene enrichment was performed using the R package clusterProfiler v3.10.1 software (38), which implements a hypergeometric model to test for gene set overrepresentation relative to a background gene set. Enrichment was analyzed using the functions: ‘Enriched Kyoto Encyclopedia of Genes and Genomes (KEGG)’, ‘gseKEGG’ (Gene Set Enrichment Analysis (GSEA) (39) of KEGG, and a function of clusterProfiler), P<0.05, OrgDb=org.Hs.eg.db (a database for Genome Wide Annotation for Human, referenced by clusterProfiler).

Integrated analysis of whole genome and transcriptome sequencing data

The differentially expressed genes were compared with mutated genes identified by whole genome sequencing. To achieve this, genes that were mutated and with differentially expressed transcripts were merged in R v3.5.1 (33).

Analysis of public GC RNA-Seq data

The RNA-Seq data from The Cancer Genome Atlas stomach adenocarcinoma (TCGA STAD) was downloaded from TCGAbiolinks online tool (https://github.com/BioinformaticsFMRP/TCGAbiolinks). To identify the most significantly differentially expressed genes in GC, the results from three RNA-Seq differential expression analysis tools, including limma v3.42.2 (40), DEseq2 v1.26.0 (41) and edgeR v3.28.1 (18), were combined and the overlapping genes were separated for further analysis. For all three tools, P<0.05 and the average log2 fold change ± standard deviation were used as the cut-off values. Gene Ontology (42) enrichment analysis was performed using the Database for Annotation, Visualization and Integrated Discovery (DAVID) v6.7 functional annotation clustering tool to determine the potential functions of the differentially expressed genes. The KEGG pathway analysis was performed to determine the involvement of differentially expressed genes in different biological pathways. The-log10 (P) denotes the enrichment score, indicating the significance of the pathway associations.

Kaplan-Meier plotter (KM-Plotter) analysis

KM-plotter (https://kmplot.com/analysis/) is an online database and tool which can be used for the discovery and validation of survival biomarkers (43). It contains cancer gene expression data and survival information from Gene Expression Omnibus (44), European Genome-phenome Archive (EGA) (https://www.ebi.ac.uk/ega/home), and TCGA (https://www.cancer.gov/tcga), and uses each percentile of mRNA expression between the lower (25%) and upper quartiles (75%) of expression as a cut-off point to divide patients into high and low expression groups. The KM-plotter software, which included available transcriptome and survival data (both overall and disease-free survival times), was used to analyze a total of 1,440 patients with GC, based on default parameters.

Results

Tumor purity and coverage statistics

The proportion and the ploidy of tumor cells in the sample were ~22% and 1.74, respectively. Using whole genome sequencing, the total number of paired reads were 8.13 and 8.34 billion, the mean read coverage was 37.16 and 35.48, while the mean library insert size was 319 and 323 bp for normal and tumor tissue samples, respectively. For the consensus coding sequence region, ~90% of the bases had over 30X coverage. Using whole transcriptome sequencing, 1.04 and 1.11 billion clean reads were obtained with an average input read length of 253 and 261 bp for normal and tumor tissue, respectively.

Gene expression analysis

The RNA-Seq pipeline was used (Fig. S1A) for the whole transcriptome sequencing. Genes with P≤0.001 and absolute log2 fold change value ≥1 were considered as significantly differentially expressed genes. In total, 766 down- and 765 upregulated genes were obtained (Fig. S1B). The RNA-Seq result was subsequently matched to the whole genome sequencing data for further analysis.

Germline mutation analysis

Germline mutation analysis revealed a total of 4,228,339 single nucleotide polymorphisms (SNPs) and Indels with 95.03% of the sites previously reported in the dbSNP database. The transition to transversion ratio was 2.05, indicating that the germline analysis had high confidence, as the empirical value for whole genome analysis is ~2.1 (31). Among the germline variants, 21,405 SNPs/Indels were located in the exonic regions, of which 10,136 were predicted to alter the protein, including 9,504 non-synonymous single nucleotide variations (SNVs), 112 frameshift deletions, 90 frameshift insertions, 173 non-frameshift deletions, 169 non-frameshift insertions, 78 stop-gain, and 10 stop-loss. More than 1% minor allele frequency variants from publicly available databases were removed, as those sites are common in the population. Of the candidate genes from the CGC and CPG databases, 13 SNPs and two Indels were identified in 15 genes (Table I). A deleterious score evaluation of the 15 variants revealed that mutations in DOCK8 (p.D63N), HLF (p.V59A), NKX2-1 (p.G322S), NRG1 (p.S17C), PRDM2 (p.S823T), WNK2 (p.C652R) were deleterious, while mutations in KMT2D (p.V401M), RAD51D (p.A181V) and ROS1 (p.E1605Q) were possibly damaging.

Table I.

Functional assessment of the germline mutations and prediction of their effect.

Table I.

Functional assessment of the germline mutations and prediction of their effect.

GeneDatabseChromosomePositionReference alleleAlterationTypeAA. ChangeCosmic82Final prediction
AKAP9CGC791739445CTNon-SNVp.T3899INU
DOCK8CPG9286491GANon-SNVp.D63NYD
HIF1ACGC1462207747CANon-SNVp.T669NNU
HLFCGC1753345172TCNon-SNVp.V59AND
KMT2DCGC1249446404CTNon-SNVp.V401MNP
MLLT3CGC920414344CTGNon-FS DELp.163_164delYU
MN1CGC2228194933TGCTGCNon-FSp.Q533delinsNU
TGCTGCINSQQQQQ
NKX2-1CGC1436986635CTNon-SNVp.G322SND
NRG1CGC832505286CGNon-SNVp.S17CND
PRDM2CGC114107360TANon-SNVp.S823TYD
PTPRCCGC1198675962AGNon-SNVp.N101SNU
RAD51DCPG1733428245GANon-SNVp.A181VNP
RNF213CGC1778346531CANon-SNVp.P4250TNU
ROS1CGC6117662652CGNon-SNVp.E1605QNP
WNK2CGC996015284TCNon-SNVp.C652RND

[i] Non-SNV, non-synonymous single nucleotide variation; non-FS DEL, non-frameshift deletion; non-FS INS, non-frameshift insertion; CGC, Cancer Gene Census; CPG, Cancer Predisposition Genes; P, possibly damaging; D, deleterious; U, uncertain significance; Y, yes; N, no; AA, amino acid.

Somatic mutation analysis

A total of 3 softwares from the Broad Institute were used to identify the somatic variants and 114 variants in the exonic regions were found. This included 45 non-synonymous SNVs, 6 frameshift deletions, 11 frameshift insertions, 17 non-frameshift deletions, 3 non-frameshift insertions, 2 stop-gain and 30 synonymous SNVs. Similar to the germline mutation analysis, common variants were filtered out with publicly available datasets to select the most informative somatic variants. In total, 29 variants were obtained, including 13 mutations which were reported in the Catalogue of Somatic Mutations in Cancer database (Table SI). Of these, 3 protein-truncating variants (BPNT1, p.S144I; FRG1, p.N153D; and TAS2R31, p.L59F) were predicted to be deleterious and one variant (KRTAP5-3, p.C185S) was predicted to be possibly damaging. However, for transcriptional regulation, 3 genes (MUC2, MUC4 and SLC8A2) were found to be differentially expressed.

Large-scale variation analysis

To identify large-scale variations in the patient, the copy number alterations and complex structural events in the tumor sample were investigated. The structural variant discovery tool, DELLY, reported 701 structural events; however, only in-frame events or fusions over 50 kb between two genes were retained. Ultimately, 20 significant structural variations were reported (Table SII). Another sensitive tool, FACETS, detected 58 large segments, of which 62 were deletions and 112 were duplications found in the coding regions (Table SIII). Combining the results from the two tools, nine genes (CDKN1B, FBXW7, MUC4, CCND3, ETV6, FGFR2, FGFR3, TFEB and KMT2C) were reported in either the CGC or CPG databases. CDKN1B, FBXW7 and MUC4 showed duplications, and there were deletions in CCND3, ETV6, FGFR2, FGFR3 and TFEB. Although KMT2C rearranged with BAGE2, the latter was not reported in the 2 databases. The mRNA expression levels of the genes with structural variations was also investigated and 21 were found to be differentially expressed in the tumor sample (Fig. 1).

Integrated pathway analysis results

RNA-Seq and mutation data were analyzed to characterize the genomic alterations in known signaling pathways. The GO enrichment for somatic mutated genes revealed terms, which included 3 mucin family genes, MUC2, MUC4 and MUC6, and were enriched in ‘maintenance of gastrointestinal epithelium’, ‘O-glycan processing’, and digestive system process, while KRTAP5-3, TCHH, RPTN, POU3F1 were enriched in epidermal cell differentiation pathway (Table SIV). The KEGG enrichment analysis revealed SLC8A2 and SLC25A5 were enriched in calcium signaling and cGMP-PKG signaling pathways (Table SIV). From the RNA-Seq data, GSEA showed that the upregulated genes were enriched in ‘cGMP-PKG signaling’, ‘calcium signaling’, and ‘oxytocin signaling’ pathways, while the downregulated genes were enriched in ‘cell cycle’ pathway, ‘oxidative phosphorylation’ pathway, ‘transcriptional mis-regulation in cancer’ pathway and ‘GC pathway’ (Fig. 2).

Comparison with TCGA STAD GC features

To ensure the present study was representative and valuable, the genomic features of the patient was compared with data from TCGA STAD dataset, which was downloaded from TCGAbiolinks, and 329 (136 up- and 193 downregulated genes) differentially expressed genes were identified, which were also found in the RNA-Seq analysis. A total of three tools were used to identify the most significantly differentially expressed genes, and 136 up- and 193 downregulated genes were identified (Fig. S2A), and were subsequently investigated in the tumor tissue sample obtained from the patient with GC and compared with that in the adjacent normal tissue. In total, 27 genes were identified, including five that were up- and 22 that were downregulated (Fig. 3A). GO and KEGG pathway enrichment analysis revealed one (GO:MF) and three (KEGG) significant terms, respectively, including digestion and gastric acid secretion (Fig. 3B).

To determine the associated functional genomic alterations of these 27 genes, the previous studies that reported them were analyzed. In the present study, there were no somatic mutations in the 27 genes, however, 12 of the genes, including gastrokine 2 (GKN2), prosaposin-like 1 (PSAPL1), ethanolamine-phosphate phospho-lyase (ETNPPL), cytochrome P450 family 2 subfamily W member 1 (CYP2W1), SFRP4, collagen type XXII α 1 chain, prostate stem cell antigen, lipase F, gastric type, solute carrier family 15 member 1 (SLC15A1), ATPase H+/K+ transporting subunit β (ATP4B), aldehyde dehydrogenase 3 family member A1 (ALDH3A1) and cystatin SN (CST1) were mutated in germline. Notably, two (SFRP4 and SLC15A1) of the 12 germline mutated genes have been reported in previous studies. Germline mutations in SFRP4 (P320T and R340K) were reported to be strongly associated with human cancer types (45) while mutations in SLC15A1 were linked with bowel diseases (46).

The KM plotter database was subsequently used with the log-rank test to investigate the association between the expression level of the 27 genes and clinical outcomes (both disease-free and overall survival times). In the database, patients with GC are divided into high and low expression groups, with each percentile of mRNA expression between the lower (25%) and upper quartiles (75%) used as a cut-off point. A total of 10 genes (ALDH3A1, ETNPPL, activity regulated cytoskeleton associated protein (ARC), ATP4B, CST1, CYP2W1, GKN2, PSAPL1, SFRP4 and SLC15A1) with differential expression were significantly associated (P<0.05) with survival times. In all the 10 genes, those with high expression (red curve) were correlated with a poorer survival while genes with low expression (black curve) correlated with higher survival in patients with GC (Fig. S2B).

Discussion

In an attempt to improve the understanding of the pathobiology of GC in young patients and provide insights for personalized treatment strategies, the genome of a 26-year old patient with GC was performed using both whole genome and whole transcriptome sequencing. Germline and somatic mutations analyses of the whole genome sequencing data revealed 15 cancer related genes with germline mutations, 27 genes with 29 informative somatic mutations and nine genes with structural events that could underlie the progression of GC in the patient. Of the 27 genes containing somatic mutations, 26 have not been previously reported as driver genes in TCGA (42). Only MUC6, a immunohistochemical marker used to support the diagnosis of gastric-type EA (47), has been reported as a driver gene. In the mutation significance analysis, a set of scoring system was used to predict deleterious amino acid substitutions. This analysis revealed nine germline variants and four somatic mutations with potentially pathogenic sites. Compared with somatic mutations, more deleterious germline sites were detected and genes containing these sites were all found to be cancer-associated, suggesting that inheritance disorders might be a key risk factor for the development of GC in the patient.

To improve the understanding of the inter-relationships between genomic alterations and the transcriptome, the mutated genes at the transcriptional level were also investigated. It was found that genes with somatic mutations, such as MUC2, MUC4, SLC8A2, and with somatic structural mutations, such as CCND3, FGFR2 and FGFR3, were differentially expressed, suggesting that they could be promising precision therapy targets.

Mucin genes are known to maintain the gastrointestinal epithelium, and persistent mucosal inflammation increase the risk of developing GC (48). In addition, the fibroblast growth factor receptor (FGFR) pathway plays a key role in GC pathogenesis, and detection of FGFR2 copy number in the plasma circulating tumor DNA are potential predictive biomarkers to FGFR inhibition (49).

With respect to heterogeneity of GC, it was hypothesized that mutations in different genes that are involved in the same pathway could result in similar clinical phenotypes. Therefore, different pathways were investigated that could be dysregulated in the patient with GC. The calcium signaling pathway, cGMP-PKG signaling pathway and transcriptional mis-regulation were enriched at both the genomic and transcriptomic levels, suggesting that these pathways might play crucial roles in young patients with GC. Notably, these pathways were not found from the analysis of TCGA STAD dataset (42), suggesting that the pathogenesis of GC in young individuals, particularly in the patient in the present study may be different. Nevertheless, the genomic features of the young patient with GC was compared with TCGA STAD data and 27 differential expression genes were identified, particularly germline mutations in SFRP4 (P320T and R340K), which are strongly associated with human cancers (45). SFRP4 is a member of the SFRP family (50). SFRP4s contain a cysteine-rich domain homologous to the putative Wnt-binding site of Frizzled proteins (50) and act as soluble modulators of Wnt signaling, which is a well-known pathway that is involved in tumorigenesis of GC (51). This indicates that such germline mutations could be informative in designing patient specific therapies. Several germline mutations were found in the 27 genes; however, no somatic alterations or structural events were identified, suggesting that genetic factors may be important in the development of early GC.

Notably, the youngest patient in TCGA STAD dataset was 30-years-old and >99% (439/443) of the patients were >40 years old, thus, TCGA STAD data could have originated from elderly patients with GC. This supports the requirement to investigate the genome in younger patients with GC to identify possible therapeutic targets. The patient in the present study was diagnosed at age 26, which was much younger compared with patients in TCGA STAD dataset. TCGA-STAD RNA-seq data was reanalyzed and 331 dysregulated genes were found (138 were up- and 193 were downregulated), however, only 27 of these were found in the patient in the present study, suggesting that there was a considerable difference between GC in young and elderly patients. There may be indications that the pathogenesis of the patient was substantially different from typical patients with GC, as cancer-related pathways, which were significantly enriched at both genomic and transcriptomic levels in the patient were also not found in TCGA STAD dataset.

In conclusion, genes with germline (SFRP4), and somatic mutations (MUC2, MUC4, SLC8A2), and those with structural variations (CCND3, FGFR2 and FGFR3), which were differentially expressed in the patient could be promising precision therapy targets.

Supplementary Material

Supporting Data

Acknowledgements

Not applicable.

Funding

This study was supported in part by the Hefei National Laboratory for Physical Sciences at Microscale (grant no. HFNLM20180012).

Availability of data and materials

We have deposited the data to National Genomics Data Center with the project number PRJCA002949. The datasets are not publicly accessible due to local data protection laws. However, access to the data can be granted upon a reasonable request to the corresponding author.

Authors' contributions

QW and KH conceived the study and designed the experiments. LL and ZH provided the tissue samples and supplied the clinical and pathological information of the patient. QR, OEA and SW performed sample preparation, DNA isolation for sequencing and library construction. KH and WY performed the bioinformatics analysis. OEA, WY, KH and QW drafted and revised the manuscript and supplementary information. All authors approved the final version of the manuscript.

Ethics approval and consent to participate

This study was approved by the Ethics Committee of the First Affiliated Hospital of Anhui Medical University. The patient provided written informed consent for the use of the resected tissue, and the sample was obtained under Institutional Review Board approval.

Patient consent for publication

Not applicable.

Competing interests

The authors declare that they have no competing interests.

Glossary

Abbreviations

Abbreviations:

GC

gastric cancer

CGC

Cancer Gene Census

CPG

Cancer Predisposition Genes

Indels

insertions and deletions

SNP

single nucleotide polymorphism

SNV

single nucleotide variation

GO

Gene Ontology

KEGG

Kyoto Encyclopedia of Genes and Genomes

TCGA STAD

The Cancer Genome Atlas stomach adenocarcinoma

References

1 

De Vita F, Di Martino N, Fabozzi A, Laterza MM, Ventriglia J, Savastano B, Petrillo A, Gambardella V, Sforza V, Marano L, et al: Clinical management of advanced gastric cancer: The role of new molecular drugs. World J Gastroentero. 20:14537–14558. 2014. View Article : Google Scholar

2 

Mori M, Sugimachi K, Ohiwa T, Okamura T, Tamura S and Inokuchi K: Early gastric carcinoma in Japanese patients under 30 years of age. Br J Surg. 72:289–291. 1985. View Article : Google Scholar : PubMed/NCBI

3 

Lim S, Lee HS, Kim HS, Kim YI and Kim WH: Alteration of E-cadherin-mediated adhesion protein is common, but microsatellite instability is uncommon in young age gastric cancers. Histopathology. 42:128–136. 2003. View Article : Google Scholar : PubMed/NCBI

4 

Smith BR and Stabile BE: Extreme aggressiveness and lethality of gastric adenocarcinoma in the very young. Arch Surg. 144:506–510. 2009. View Article : Google Scholar : PubMed/NCBI

5 

Quijano Orvananos F, Moreno Paquentin E, Alvarez JJ, Martinez Munive A and Butron Perez L: Gastric carcinoma in patients under 35 years. Rev Gastroenterol Mex. 64:75–77. 1999.(In Spanish). PubMed/NCBI

6 

Theuer CP, Kurosaki T, Taylor TH and Anton-Culver H: Unique features of gastric carcinoma in the young: A population-based analysis. Cancer. 83:25–33. 1998. View Article : Google Scholar : PubMed/NCBI

7 

Wierinckx A, Roche M, Raverot G, Legras-Lachuer C, Croze S, Nazaret N, Rey C, Auger C, Jouanneau E, Chanson P, et al: Integrated genomic profiling identifies loss of chromosome 11p impacting transcriptomic activity in aggressive pituitary PRL Tumors. Brain Pathol. 21:533–543. 2011.PubMed/NCBI

8 

Cancer Genome Atlas Research Network, . Bass AJ, Thorsson V, Shmulevich I, Reynolds SM, Miller M, Bernard B, Hinoue T, Laird PW, Curtis C, et al: Comprehensive molecular characterization of gastric adenocarcinoma. Nature. 513:202–209. 2014. View Article : Google Scholar : PubMed/NCBI

9 

Chmielecki J and Meyerson M: DNA sequencing of cancer: What Have We Learned? Annu Rev Med. 65:63–79. 2014. View Article : Google Scholar : PubMed/NCBI

10 

Garraway LA and Lander ES: Lessons from the Cancer Genome. Cell. 153:17–37. 2013. View Article : Google Scholar : PubMed/NCBI

11 

Vogelstein B, Papadopoulos N, Velculescu VE, Zhou SB, Diaz LA and Kinzler KW: Cancer genome landscapes. Science. 339:1546–1558. 2013. View Article : Google Scholar : PubMed/NCBI

12 

Garraway LA: Genomics-driven oncology: Framework for an emerging paradigm. J Clin Oncol. 31:1806–1814. 2013. View Article : Google Scholar : PubMed/NCBI

13 

Chen R: Abstract 4495: Development and clinical application of an integrative genomic approach to personalized cancer therapy. Cancer Res. 762016.doi: 10.1158/1538-7445.AM2016-4495.

14 

Edge SB and Compton CC: The American Joint Committee on Cancer: The 7th edition of the AJCC cancer staging manual and the future of TNM. Ann Surg Oncol. 17:1471–1474. 2010. View Article : Google Scholar : PubMed/NCBI

15 

Krueger F: Trim Galore: A wrapper tool around Cutadapt and FastQC to consistently apply quality and adapter trimming to FastQ files, with some extra functionality for MspI-digested RRBS-type (Reduced Representation Bisufite-Seq) libraries. 2012.

16 

Dobin A, Davis CA, Schlesinger F, Drenkow J, Zaleski C, Jha S, Batut P, Chaisson M and Gingeras TR: STAR: Ultrafast universal RNA-seq aligner. Bioinformatics. 29:15–21. 2013. View Article : Google Scholar : PubMed/NCBI

17 

Anders S, Pyl PT and Huber W: HTSeq-a Python framework to work with high-throughput sequencing data. Bioinformatics. 31:166–169. 2015. View Article : Google Scholar : PubMed/NCBI

18 

Robinson MD, McCarthy DJ and Smyth GK: edgeR: A Bioconductor package for differential expression analysis of digital gene expression data. Bioinformatics. 26:139–140. 2010. View Article : Google Scholar : PubMed/NCBI

19 

Li H: Aligning sequence reads, clone sequences and assembly contigs with BWA-MEM. (Submitted on 16 Mar 2013 (v1), last revised 26 May, 2013 (this version, v2)).

20 

DePristo MA, Banks E, Poplin R, Garimella KV, Maguire JR, Hartl C, Philippakis AA, del Angel G, Rivas MA, et al: A framework for variation discovery and genotyping using next-generation DNA sequencing data. Nat Genet. 43:491–498. 2011. View Article : Google Scholar : PubMed/NCBI

21 

Poplin R, Ruano-Rubio V, DePristo MA, Fennell TJ, Carneiro MO, Van der Auwera GA, Kling DE, Gauthier LD, Levy-Moonshine A, Roazen D, et al: Scaling accurate genetic variant discovery to tens of thousands of samples. BioRxiv. July 24–2017.doi: https://doi.org/10.1101/201178.

22 

Cibulskis K, Lawrence MS, Carter SL, Sivachenko A, Jaffe D, Sougnez C, Gabriel S, Meyerson M, Lander ES and Getz G: Sensitive detection of somatic point mutations in impure and heterogeneous cancer samples. Nat Biotechnol. 31:213–219. 2013. View Article : Google Scholar : PubMed/NCBI

23 

Colaprico A, Silva TC, Olsen C, Garofano L, Cava C, Garolini D, Sabedot TS, Malta TM, Pagnotta SM, Castiglioni I, et al: TCGAbiolinks: An R/Bioconductor package for integrative analysis of TCGA data. Nucleic Acids Res. 44:e712016. View Article : Google Scholar : PubMed/NCBI

24 

1000 Genomes Project Consortium, . Auton A, Brooks LD, Durbin RM, Garrison EP, Kang HM, Korbel JO, Marchini JL, McCarthy S, McVean GA and Abecasis GR: A global reference for human genetic variation. Nature. 526:68–74. 2015. View Article : Google Scholar : PubMed/NCBI

25 

Karczewski KJ, Weisburd B, Thomas B, Solomonson M, Ruderfer DM, Kavanagh D, Hamamsy T, Lek M, Samocha KE, Cummings BB, et al: The ExAC browser: Displaying reference data information from over 60 000 exomes. Nucleic Acids Res. 45:D840–D845. 2016. View Article : Google Scholar : PubMed/NCBI

26 

Karczewski KJ and Karczewski LF: The genome aggregation database (gnomAD). 2017.

27 

Liu X, Wu C, Li C and Boerwinkle E: dbNSFP v3.0: A One-stop database of functional predictions and annotations for human nonsynonymous and splice-site SNVs. Hum Mutat. 37:235–241. 2016. View Article : Google Scholar : PubMed/NCBI

28 

Sherry ST, Ward MH, Kholodov M, Baker J, Phan L, Smigielski EM and Sirotkin K: dbSNP: The NCBI database of genetic variation. Nucleic Acids Res. 29:308–311. 2001. View Article : Google Scholar : PubMed/NCBI

29 

Tate JG, Bamford S, Jubb HC, Sondka Z, Beare DM, Bindal N, Boutselakis H, Cole CG, Creatore C, Dawson E, et al: COSMIC: The catalogue of somatic mutations in cancer. Nucleic Acids Res. 47:D941–D947. 2019. View Article : Google Scholar : PubMed/NCBI

30 

Kimura M: A simple method for estimating evolutionary rates of base substitutions through comparative studies of nucleotide-sequences. J Mol Evol. 16:111–120. 1980. View Article : Google Scholar : PubMed/NCBI

31 

Mwenifumbo JC and Marra MA: Cancer genome-sequencing study design. Nat Rev Genet. 14:321–332. 2013. View Article : Google Scholar : PubMed/NCBI

32 

Shen R and Seshan VE: FACETS: Allele-specific copy number and clonal heterogeneity analysis tool for high-throughput DNA sequencing. Nucleic Acids Res. 44:e1312016. View Article : Google Scholar : PubMed/NCBI

33 

Team RC: R: A language and environment for statistical computing. 2013.

34 

Rausch T, Zichner T, Schlattl A, Stutz AM, Benes V and Korbel JO: DELLY: Structural variant discovery by integrated paired-end and split-read analysis. Bioinformatics. 28:i333–i339. 2012. View Article : Google Scholar : PubMed/NCBI

35 

Dong C, Wei P, Jian X, Gibbs R, Boerwinkle E, Wang K and Liu X: Comparison and integration of deleteriousness prediction methods for nonsynonymous SNVs in whole exome sequencing studies. Hum Mol Genet. 24:2125–2137. 2015. View Article : Google Scholar : PubMed/NCBI

36 

Futreal PA, Coin L, Marshall M, Down T, Hubbard T, Wooster R, Rahman N and Stratton MR: A census of human cancer genes. Nat Rev Cancer. 4:177–183. 2004. View Article : Google Scholar : PubMed/NCBI

37 

Rahman N: Realizing the promise of cancer predisposition genes. Nature. 505:302–308. 2014. View Article : Google Scholar : PubMed/NCBI

38 

Yu GC, Wang LG, Han YY and He QY: clusterProfiler: An R package for comparing biological themes among gene clusters. OMICS. 16:284–287. 2012. View Article : Google Scholar : PubMed/NCBI

39 

Subramanian A, Tamayo P, Mootha VK, Mukherjee S, Ebert BL, Gillette MA, Paulovich A, Pomeroy SL, Golub TR, Lander ES and Mesirov JP: Gene set enrichment analysis: A knowledge-based approach for interpreting genome-wide expression profiles. Proc Natl Acad Sci USA. 102:15545–15550. 2005. View Article : Google Scholar : PubMed/NCBI

40 

Ritchie ME, Phipson B, Wu D, Hu Y, Law CW, Shi W and Smyth GK: limma powers differential expression analyses for RNA-sequencing and microarray studies. Nucleic Acids Res. 43:e472015. View Article : Google Scholar : PubMed/NCBI

41 

Love MI, Huber W and Anders S: Moderated estimation of fold change and dispersion for RNA-seq data with DESeq2. Genome Biol. 15:5502014. View Article : Google Scholar : PubMed/NCBI

42 

Bailey MH, Tokheim C, Porta-Pardo E, Sengupta S, Bertrand D, Weerasinghe A, Colaprico A, Wendl MC, Kim J, Reardon B, et al: Comprehensive characterization of cancer driver genes and mutations. Cell. 174:1034–1035. 2018. View Article : Google Scholar : PubMed/NCBI

43 

Szász AM, Lánczky A, Nagy Á, Förster S, Hark K, Green JE, Boussioutas A, Busuttil R, Szabó A and Győrffy B: Cross-validation of survival associated biomarkers in gastric cancer using transcriptomic data of 1,065 patients. Oncotarget. 7:49322–49333. 2016. View Article : Google Scholar : PubMed/NCBI

44 

Taiaroa G, Rawlinson D, Featherstone L, et al: Direct RNA sequencing and early evolution of SARS-CoV-2. bioRxiv. 2020.(Epub ahead of print).

45 

He ML, Chen Y, Chen Q, He Y, Zhao J, Wang J, Yang H and Kung HF: Multiple gene dysfunctions lead to high cancer-susceptibility: Evidences from a whole-exome sequencing study. Am J Cancer Res. 1:562–573. 2011.PubMed/NCBI

46 

Zucchelli M, Torkvist L, Bresso F, Halfvarson J, Hellquist A, Anedda F, Assadi G, Lindgren GB, Svanfeldt M, Janson M, et al: PepT1 oligopeptide transporter (SLC15A1) gene polymorphism in inflammatory bowel disease. Inflamm Bowel Dis. 15:1562–1569. 2009. View Article : Google Scholar : PubMed/NCBI

47 

Hodgson A, Parra-Herran C and Mirkovic J: Immunohistochemical expression of HIK1083 and MUC6 in endometrial carcinomas. Histopathology. 75:552–558. 2019. View Article : Google Scholar : PubMed/NCBI

48 

Li M, Huang L, Qiu H, Fu Q, Li W, Yu Q, Sun L, Zhang L, Hu G, Hu J and Yuan X: Helicobacter pylori infection synergizes with three inflammation-related genetic variants in the GWASs to increase risk of gastric cancer in a Chinese population. PLoS One. 8:e749762013. View Article : Google Scholar : PubMed/NCBI

49 

Hierro C, Alsina M, Sanchez M, Serra V, Rodon J and Tabernero J: Targeting the fibroblast growth factor receptor 2 in gastric cancer: Promise or pitfall? Ann Oncol. 28:1207–1216. 2017. View Article : Google Scholar : PubMed/NCBI

50 

Cruciat CM and Niehrs C: Secreted and transmembrane wnt inhibitors and activators. Cold Spring Harb Perspect Biol. 5:a0150812013. View Article : Google Scholar : PubMed/NCBI

51 

Herr P, Hausmann G and Basler K: WNT secretion and signalling in human disease. Trends Mol Med. 18:483–493. 2012. View Article : Google Scholar : PubMed/NCBI

Related Articles

Journal Cover

October-2020
Volume 20 Issue 4

Print ISSN: 1792-1074
Online ISSN:1792-1082

Sign up for eToc alerts

Recommend to Library

Copy and paste a formatted citation
x
Spandidos Publications style
Hu K, Yu W, Ajayi OE, Li L, Huang Z, Rong Q, Wang S and Wu Q: Integrated analysis of whole genome and transcriptome sequencing in a young patient with gastric cancer provides insights for precision therapy. Oncol Lett 20: 115, 2020
APA
Hu, K., Yu, W., Ajayi, O.E., Li, L., Huang, Z., Rong, Q. ... Wu, Q. (2020). Integrated analysis of whole genome and transcriptome sequencing in a young patient with gastric cancer provides insights for precision therapy. Oncology Letters, 20, 115. https://doi.org/10.3892/ol.2020.11976
MLA
Hu, K., Yu, W., Ajayi, O. E., Li, L., Huang, Z., Rong, Q., Wang, S., Wu, Q."Integrated analysis of whole genome and transcriptome sequencing in a young patient with gastric cancer provides insights for precision therapy". Oncology Letters 20.4 (2020): 115.
Chicago
Hu, K., Yu, W., Ajayi, O. E., Li, L., Huang, Z., Rong, Q., Wang, S., Wu, Q."Integrated analysis of whole genome and transcriptome sequencing in a young patient with gastric cancer provides insights for precision therapy". Oncology Letters 20, no. 4 (2020): 115. https://doi.org/10.3892/ol.2020.11976