Skip to main content

Evolutionary dependency of cancer mutations in gene pairs inferred by nonsynonymous-synonymous mutation ratios

Abstract

Background

Determining the impact of somatic mutations requires understanding the functional relationship of genes acquiring mutations; however, it is largely unknown how mutations in functionally related genes influence each other.

Methods

We employed non-synonymous-to-synonymous or dNdS ratios to evaluate the evolutionary dependency (ED) of gene pairs, assuming a mutation in one gene of a gene pair can affect the evolutionary fitness of mutations in its partner genes as mutation context. We employed PanCancer- and tumor type-specific mutational profiles to infer the ED of gene pairs and evaluated their biological relevance with respect to gene dependency and drug sensitivity.

Results

We propose that dNdS ratios of gene pairs and their derived cdNS (context-dependent dNdS) scores as measure of ED distinguishing gene pairs either as synergistic (SYN) or antagonistic (ANT). Mutation contexts can induce substantial changes in the evolutionary fitness of mutations in the paired genes, e.g., IDH1 and IDH2 mutation contexts lead to substantial increase and decrease of dNdS ratios of ATRX indels and IDH1 missense mutations corresponding to SYN and ANT relationship with positive and negative cdNS scores, respectively. The impact of gene silencing or knock-outs on cell viability (genetic dependencies) often depends on ED, suggesting that ED can guide the selection of candidates for synthetic lethality such as TCF7L2-KRAS mutations. Using cell line-based drug sensitivity data, the effects of targeted agents on cell lines are often associated with mutations of genes exhibiting ED with the target genes, informing drug sensitizing or resistant mutations for targeted inhibitors, e.g., PRSS1 and CTCF mutations as resistant mutations to EGFR and BRAF inhibitors for lung adenocarcinomas and melanomas, respectively.

Conclusions

We propose that the ED of gene pairs evaluated by dNdS ratios can advance our understanding of the functional relationship of genes with potential biological and clinical implications.

Background

While cancer genomes often harbor tens of thousands of somatic mutations, only a small fraction of these mutations, known as cancer driver mutations, are implicated in tumor initiation and progression [1, 2]. Selective advantages conferred by driver mutations leave genomic footprints as evolutionary consequences of positive selection and these footprints can help in identifying cancer driver mutations [1, 3]. Such genomic hallmarks of driver mutations include higher-than-expected mutation frequencies compared to neutral mutations that are relatively free of selection [4, 5]. In addition to mutation frequencies, assessing evolutionary selection in cancer genomes can be adopted for cancer driver discovery. One metric is the normalized nonsynonymous-to-synonymous (dN/dS or dNdS) ratio. The dNdS ratios exploit that the frequency of silent mutations can serve as proxy to model expected mutation frequencies. This ratio evaluates the type of selection (i.e., dNdS ratios > 1 and < 1 indicate positive and negative selection, respectively) and the magnitude of selective pressure on individual cancer genes or genomes that shape the mutational landscape of cancer genomes [6].

Multiple cancer driver mutations often co-exist in cancer genomes, and their coordinated actions may play important roles in cancer development and progression [7, 8]. Mutations in synergistic relationships will be co-selected, whereas those in antagonistic relationships or functional redundancies will be excluded, leading to co-occurrence (CO) and mutual exclusiveness (ME), respectively. Such gene pairs have been evaluated in terms of functionality, such as the synergistic relationship between APC-KRAS [9] and SMAD4-KRAS mutations [10]. ME gene pairs may explain the functional redundancies of cancer driver mutations with complementary functions [11, 12] but also indicate the potentially antagonistic relationships of mutations leading to synthetic lethality [13]. The functional relationship across multiple mutations adds additional complexity to the selection of biomarkers for targeted therapies and poses challenges in combined therapy, for example when combined mutations of functionally associated genes predict the resistance to targeted therapy [14]. Thus, the accurate identification of functional gene pairs and assessments of their relationship are important in precision oncology with potentially significant clinical relevance [6].

Previous methodologies to identify functional pairs of mutations have primarily relied on the genomic distribution of gene pairs such as CO and ME gene pairs as reviewed elsewhere [15]. Alternative approaches such as network-based identification of recurrently mutated subnetworks and gene pairs are also available [16,17,18]. Given the utility of evolutionary measures in identifying singleton cancer drivers [3], evolutionary relationships might offer quantitative means to assess the presence and potential functionality of gene pairs in cancer genomes. However, it is still challenging to determine which evolutionary metrics can be used to infer the functional relationship of mutations in gene pairs.

In this study, we postulated that the relationships of mutations in gene pairs might represent evolutionary dependency (ED). This implies that a mutation occurring in a gene (as a mutation context) confers positive or negative selective pressure on the other mutations in the genes of the functional relationship. To quantify the level of selective pressure, we employed dNdS ratios assuming that dNdS ratios of a gene measured in the presence or absence of the mutations of other genes in the paired relationship (as mutation contexts), reflect the ED of the corresponding gene pairs. Using PanCancer and tumor type-specific mutational profiles of the Cancer Genome Atlas (TCGA) consortium [19], we evaluated the pairwise relationships of gene pairs in terms of ED. We first tested whether dNdS ratios and their derived cdNS (context-dependent dNdS ratios) scores can distinguish types of functional ED of known functional gene pairs. Gene pairs exhibiting synergistic (SYN) and antagonistic (ANT) relationships (with cdNS scores > 0 and < 0, respectively) demonstrated a good agreement with CO-ME gene pairs identified based on genomic distribution. Furthermore, we integrated data of cell essentiality and pharmacological perturbations in cancer cell lines to assess the impact of ED on genetic dependency and sensitivity to targeted agents.

Methods

Mutations

We obtained somatic mutation calls for over 10,000 tumor specimens from the TCGA consortium (“mc3.v0.2.8.PUBLIC.maf.gz”; https://gdc.cancer.gov/about-data/publications/pancanatlas) available at the GDC (Genomic Data Commons portal) website [19, 20]. The coordinates of mutations are based on human reference genome GRCh37/hg19. The consequences of mutations on amino acid residues encoding and tumor-normal (alt. and ref. counts) allele frequencies of individual mutations were also obtained from the same resources. As an independent resource, mutational profiles of 10,000 patients identified by panel-sequencing (MSKCC-IMPACT data) were obtained along with patient outcome data (overall survival) via cBioPortal (https://www.cbioportal.org/study/summary?id=msk_impact_2017) [21]. We used a reference set of gene pairs with functional relationship, obtaining 517 gene pairs (referred to as “Mut-Mut” interactions in the literature) via the SELECT algorithm [22]. We also gathered information on two specific categories of gene pairs, CO and ME (co-occurring and mutually exclusive, respectively), from the same datasets. For tumor type-specific ED analysis, we restricted the mutational profiles to those occurring in 20 tumor types with over 500 cases each. The abbreviated tumor types include bladder urothelial carcinoma (BLCA, n = 412), breast invasive carcinoma (BCRA, n = 1097), cervical squamous cell carcinoma (CESC, n = 307), colorectal adenocarcinoma (COAD/READ, n = 629), esophageal carcinoma (ESCA, n = 185), head and neck squamous carcinoma (HNSC, n = 528), kidney renal clear cell carcinoma (KIRC, n = 537), kidney renal papillary cell carcinoma (KIRP, n = 291), brain lower-grade glioma (LGG, n = 515), glioblastoma multiforme (GBM, n = 596), liver hepatocellular carcinoma (LIHC, n = 377), lung adenocarcinoma (LUAD, n = 522), lung squamous cell carcinoma (LUSC, n = 504), ovarian serous cystadenocarcinoma (OV, n = 587), pancreatic adenocarcinoma (PAAD, n = 185), prostate adenocarcinoma (PRAD, n = 500), sarcoma (SARC, n = 261), skin cutaneous melanoma (SKCM, n = 470), stomach adenocarcinoma (STAD, n = 443), and uterine corpus endometrial carcinoma (UCEC, n = 548). COAD-READ and LGG-GBM were merged for the analysis unless indicated. Microsatellite instability (MSI) status of MSI-H (high), MSI-L (low), and MSS (microsatellite stable) as well as the LGG subtypes of astrocytoma, oligoastrocytoma, and oligodendroglioma were obtained from the publications [23, 24].

ED assessment with dNdS ratios and cdNS scores

The dNdS ratios were calculated using dNdScv R packages (https://github.com/im3sanger/dNdScv) using the genome version of GRCh37/hg19, which is compatible with those of mutational profiles [6]. To assess the level of ED for gene pairs (e.g., gene A and gene B), we determined the dNdS ratios of gene B in two different mutation contexts: with and without mutations of gene A. This led to two distinct dNdS ratios: dNdSB|Amut+ (with gene A mutations) and dNdSB|Amut- (without gene A mutations), or simply, dNdS,context + and dNdS,context − , respectively. As mutation contexts (gene A mutations), we considered non-silent mutations occurring in the gene A, dividing the genomes into those with (mutation contexts) and without these mutations. We either directly compared these two dNdS ratios through scatter plots or combined them as log-odds to form a cdNS (context-dependent dNdS ratios) score (as shown in main Fig. 1b, c, respectively). The cdNS score, calculated as the log2 ratio of the dNdS ratio, context + to dNdS ratios, context − , serves to depict both the direction (e.g., mode the selection, SYN or ANT) and the magnitude of ED. We estimated the dNdS ratios and cdNS scores across three types of mutations: missense, truncating (nonsense and splice-site mutations), and indels, based on the dNdScv output.

Fig. 1
figure 1

Schematics of dNdS ratio-based evolutionary dependency (ED) for mutation pairs. a Examples of synergistic (SYN) and antagonistic (ANT) gene pairs are shown for their presentations of mutations (depicted as the relationship between gene A and B). Two dNdS ratios (dNdS, context + and dNdS, context −) are calculated for gene B mutations in the presence and absence of mutation contexts (gene A mutations considered as mutation contexts). b Two dN/dS ratios are shown for known gene pairs with segregating gene pairs reflecting the types of functional relationship (red and blue for gene pairs representing co-occurrences (CO) and mutual exclusivity (ME), respectively). Three types of dN/dS ratios from missense, truncating, and indel mutations are shown separately, for those with significant ED (61 missense, 47 truncating mutations, and 41 indels, P < 0.01, permutation tests). c Summarized log2 odds of dN/dS ratios are calculated as cdNS (context-dependent dNdS ratios) scores and shown for missense, truncating mutations and indels. d cdNS scores of known gene pairs are compared with those obtained from independent MSK-IMPACT mutational profiles (TCGA and MSK-IMPACT, x- and y-axis, respectively)

To discover novel gene pairs exhibiting mutation-based ED, we selected a set of 312 genes. This comprised 220 genes identified in PanCancer mutational profiles (as determined by dNdScv with q_global < 0.1) and additional 92 genes that showed significant ED across tumor type-specific mutational profiles (significant in at least one tumor type). Details on these 312 genes and the tumor types analyzed are provided in Additional file 1:Table S1. The 312 genes served as mutation contexts, wherein dNdScv was employed to analyze genomes with and without their individual mutational contexts, utilizing the default set of 20,092 genes provided by the packages. The cdNS scores of individual gene pairs were calculated as the log2 ratio of two dNdS ratios (context + /context −), and their statistical significance was assessed via permutation tests unless otherwise specified.

Statistical evaluation of gene pairs

The significance of cdNS scores was determined through permutation tests, where we shuffled the mutation context labels of genomes to generate a set of permuted dNdS ratios, context + and dNdS ratio, context − along with their cdNS scores for each gene pair. The significance was then assessed based on the frequency with which the permuted cdNS scores (1000 permutation tests) exceeded the actual score, providing nominal P values. In this exploratory analysis, we set the significance cutoff of gene pairs to be P < 0.01. To alleviate the tumor type-specific influence on the identification of gene pairs exhibiting ANT relationship in PanCancer analyses, we employed previously proposed sample size correction methods [25]. As previously described, a correction factor w was employed to reduce the number of mutation-negative genomes by rescaling the original denominator N (the total number of genomes). The correction factor w was calculated using sigmoid functions with recommended parameters [25]. We also used Fisher’s exact test to calculate the significance level for gene pairs in ANT relationship in PanCancer analysis by using the corrected sample sizes instead of permutation tests.

For power analysis, we selected two known gene pairs of KRAS-STK11 synergistic (SYN) pair and KRAS-NRAS antagonistic (ANT) pair. To test the subsampling effect, we first selected 138 STK11-mutant and 272 NRAS-mutant genomes in PanCancer dataset as genomes with mutation contexts. Then, we subsampled varying number of these mutant genomes (ranging from 1 to 100 genomes), performing 100 subsampling for each group. We then calculated dNdS ratios (mutation context +) of KRAS missense mutations in subsample dataset to compare with dNdS ratios (mutation context −) to derive cdNS scores of KRAS-STK11 and KRAS-NRAS pairs (Additional file 2: Figure S1a and S1b for details). The number of KRAS missense mutations in each permutation is also counted in subsampled datasets to use as proxy to determine the minimum number of mutations. Consistent cdNS scores were observed in at least 10 STK11 and 50 NRAS mutant genomes where 3 and 1 KRAS missense mutations were observed. Based on this, we set a global, minimum threshold of mutations for SYN pairs (≥ 3 mutations) and ANT pairs (≥ 1 mutation). We also confirmed the cdNS values were relatively stable for these two gene pairs, even when the number of context − genomes was reduced to match that of the context + genomes (Additional file 2: Figure S1c). Additionally, we investigated the robustness of cdNS values with respect to the expression levels of mutation contexts. For this, the genomes with mutation contexts were divided into two categories: those whose genes of mutation contexts were expressed above the median level and those not (context-expressed and context-nonexpressed, respectively). For 85 gene pairs identified in PanCancer analyses (main Fig. 1), cdNS scores were calculated for genomes with context-expressed and context-nonexpressed. A significant correlation was observed between these two types of cdNS scores, indicating that the cdNS evaluation is relatively robust to the expression levels of mutation contexts (Additional file 2: Figure S1d).

Finally, SYN pairs in PanCancer data as well as SYN/ANT tumor type-specific pairs were evaluated for statistical significance by permutation tests. In PanCancer ED analyses, we selected gene pairs that showed significant ED (P < 0.01, permutation tests for PanCancer SYN pairs and P < 0.01, Fisher’s exact test for PanCancer ANT pairs), with a minimum number of mutations (SYN pairs ≥ 3 and ANT pairs ≥ 1), and ED ratios (dNdS context + /dNdS context −  > 3 or < 0.3) resulting in 85 gene pairs in PanCancer dataset. For tumor type-specific gene pairs, we also used the same criteria of PanCancer gene pairs (P < 0.01 in permutation tests both for SYN and ANT pairs, minimum number of mutations and ED ratios) identifying 3870 gene pairs across 20 tumor types.

Mutation features

Variant allele frequencies (VAF) were determined by dividing the count of altered (alt. counts) alleles by the total number of reads (alt. counts + ref. counts) at the corresponding genetic loci. VAFs were computed separately for each gene of tumor type-specific gene pairs and gene-wise summarized as median values in corresponding tumor types. We then calculated the difference in VAF between two genes in each pair, by subtracting the median VAF of the context genes (gene A) from the median VAF of genes associated with the contexts (gene B). Thus, we expect the negative difference in VAF values (ΔVAF) when the mutations in context genes and their partner genes are clonal and subclonal, respectively, and positive values for vice versa. The ΔVAF were compared between SYN and ANT gene pairs across tumor types examined. For association with tumor mutation burdens (TMB), we considered the number of non-silent mutations in each genome. For each gene pair, we calculate the TMB of genomes harboring both mutations in the gene pairs and those only harboring mutations on one gene of the pairs. The differences in TMB (ΔTMB) were also compared between SYN and ANT gene pairs in a tumor type-specific manner. For the baseline distribution of ΔVAF and ΔTMB, we permuted the genes across the 85 PanCancer gene pairs. We conducted a hundred permutation tests to obtain the baseline distributions of ΔVAF and ΔTMB, then compared them with those of the SYN and ANT gene pairs.

Relationship of ED and gene dependency

Gene dependencies of cancer cell lines were obtained from three databases: ANANA [26], DEMETER2 [27] and DRIVE [28] as indicated previously [22]. For each database, we estimated the genetic dependencies of gene pairs by calculating effect size with respect to mutation contexts. For instance, in the case of KRAS-NRAS gene pairs, where NRAS mutations represent mutation contexts, we collected cell lines containing KRAS missense mutations along with KRAS gene knockouts. These cell lines were then categorized into two groups: those with NRAS context mutations and those without. The effect size was determined as the difference in the median of cell viability between these two groups, aiming to evaluate whether NRAS context mutations impact KRAS knockouts in KRAS-mutated cell lines, either enhancing or compromising the cell viability. We then compared the effect size with their cdNS values as calculated in PanCancer or tumor type-specific mutational profiles. We also obtained gene pairs that appeared more than once across three databases of genetic dependency and calculated their average of effect size. Then, we classified gene pairs as either “compromising” or “rescuing,” determined by negative and positive effect sizes, respectively. A negative effect size (compromising) suggests that the presence of mutation contexts diminishes cell viability, while a positive effect size (rescuing) indicates an increase in cell viability. When combined with SYN-ANT ED types of gene pairs, this results in four categories: compromising-SYN/-ANT and rescuing-SYN/-ANT gene pairs. We also assessed pathway-level concordance of gene pairs by examining the frequency of occurrence of both genes across functional gene sets, utilizing Gene Ontology terms or gene sets from the MSigDB database (c5 gene sets as available in https://www.gsea-msigdb.org/gsea/msigdb) [29]. The occurrences of gene pairs in GO terms were z-normalized using the mean and standard deviation derived from permuted GO terms, where genes were randomly assigned across GO terms. Furthermore, we incorporated classifications of oncogenes and tumor suppressor genes as classified in the database of Cancer Census Genes for available genes [30].

Pharmacology data

Drug response data for human cancer cell lines were obtained from GDSC (Genomics of Drug Sensitivity in Cancer) database [31]. From a combined set of two dataset versions (GDSC1 and GDSC2), the IC50 (half maximal inhibitory concentration) values were obtained across cell lines with their mutations and other clinical features such as tumor types (https://www.cancerrxgene.org/). We focused on well-established target agents (EGFR and BRAF inhibitors) in two tumor types of lung adenocarcinomas (LUAD) and skin melanomas (SKCM). The cdNS scores of gene pairs involving EGFR and BRAF mutations were collected from LUAD and SKCM tumor type-specific data. For drug sensitivity data, we obtained IC50 values of cell lines with respect to the mutations of target genes (e.g., EGFR and BRAF mutations) and their ED-associated genes (e.g., KRAS and NRAS mutations, respectively). To extend analyses of the relationship between ED and drug response, we collected IC50 values of drugs whose targets correspond to target mutations (e.g., EGFR and BRAF inhibitors) obtaining combinations of tumor type-drug-gene pairs (e.g., lung adenocarcinomas/LUAD – gefitinib – EGFR/KRAS pairs). This facilitated the comparison of drug sensitivity to target gene mutations (e.g., EGFR mutations) relative to mutation contexts (e.g., KRAS mutations) using effect size measurements. Thus, effect size was computed as the difference between IC50 values of cell lines with and without mutation contexts, both harboring the mutations corresponding to targeting agents and tumor types [31], e.g., IC50 of EGFRmt/KRASmt minus IC50 of EGFRmt/KRASwt LUAD cell lines. The effect sizes were also obtained from CCLE (Cancer Cell Line Encyclopedia) databases (https://sites.broadinstitute.org/ccle/) [32] for gene pairs supported by both datasets.

Results

dNdS ratios and cdNS scores as measures of ED of gene pairs

Given that multiple gene mutations are unlikely to occur simultaneously in a single cell, the selective pressure of paired gene mutations can be dissected into those imposed on the first and second mutation hits of the gene pair. The evolutionary pressure imposed on the second gene mutations in the presence of the first gene mutations is the major determinant of selective fitness of paired gene mutations leading to genomic footprints such as CO and ME gene pairs in cancer genomes. The first mutation establishes a mutation context that either supports or opposes the second mutation, based on their functional relationship. In this study, we adopted dNdS ratios to evaluate and quantify the levels of selective pressures on mutations in gene pairs. Specifically, we compare dNdS ratios for a second gene mutation in the context of an existing or absent first mutation, with schematics illustrated in Fig. 1a. Originally, dNdS ratios identified positive and negative evolutionary pressures (dNdS ratios > 1 or < 1, respectively) indicative of oncogenes and tumor suppressor genes (6), but can also distinguish between synergistic and antagonistic gene mutations pairs by considering mutation contexts. For instance, the dNdS ratios for gene B are evaluated in the presence of (dNdSB|Amut+) and absence (dNdSB|Amut−) of mutations in its partner gene A. A higher ratio (dNdSB|Amut+  > dNdSB|Amut−) suggests a synergistic relationship, favoring gene B mutations when gene A is mutated. Conversely, a lower ratio (dNdSB|Amut+  < dNdSB|Amut−) indicates either antagonism or functional redundancy. The pairs are termed “synergistic” (SYN, dNdSB|Amut+  > dNdSB|Amut−) and “antagonistic” (ANT, dNdSB|Amut+  < dNdSB|Amut−), respectively. These ED-based annotations of SYN/ANT are distinguished from CO and ME gene pairs inferred from genomic distribution. We define mutation contexts (mutations of gene A) as those that either advantageous or disadvantageous in the acquisition of functionally related gene mutations (gene B), thereby determining the selective fitness of gene A-B pairs. Also, we use dNdS ratios (context +) and dNdS ratios (context −) as a concise representation of dNdSB|Amut+ and dNdSB|Amut−, respectively.

As a proof-of-concept, we first examined known gene mutation pairs previously identified as CO or ME pairs (333 CO and 184 ME pairs) as available in a previous report (22). We calculated dNdS ratios for each gene in a pair considering the mutations in its partner gene’s mutations as mutation context. The analysis was conducted for three types of mutation categories: missense, truncating (nonsense and splice site), and indels, using PanCancer mutational profiles of TCGA consortium. Scatter plots demonstrate dNdS ratios for three types of mutation categories, with the x- and y-axis displaying dNdS ratio (context −) and dNdS ratio (context +) for 61 missense, 47 truncating, and 41 indel mutation gene pairs with significant ED (P < 0.01, permutation tests, Fig. 1b). The majority of genes showed dNdS ratios greater than 1 (log2 of dNdS ratio > 0), either with or without mutation context, suggesting that most genes in these pairs are likely cancer drivers under positive selection. Notably, CO and ME pairs (red and blues dots, respectively) tend to segregate along the diagonal line implying the types of selective pressure conferred on gene pairs (i.e., SYN and ANT pairs, respectively) are concordant with CO and ME genomic presentation of gene pairs. Compared with dNdS ratios of unfiltered 517 gene pairs (Additional file 2: Fig. S2a), gene pairs that located near the diagonal lines were mostly filtered based on their significance suggesting a substantial variability of ED for gene pairs previously categorized based on genomic distribution. Discordant gene pairs between two types of annotations (e.g., SYN-ME and ANT-CO pairs) were further examined regarding their level of enrichment for cancer-related genes (i.e., Cancer Census Genes) and compared with those of concordant gene pairs (e.g., SYN-CO and ANT-ME) (Additional file 2: Fig. S2b). The enrichment analysis suggests that concordant pairs are more enriched to cancer-related genes than discordant pairs. This indicates that the selective pressure conferred on cancer-related genes are likely to result in their anticipated genomic consequences.

We then introduced the “context-dependent dNdS ratios (cdNS),” determined by the log2 transformed ratio between the dNdS ratio (context +) and the dNdS (context −). This approach consolidates the two dN/dS ratios into a single value, effectively capturing the ED of gene mutation pairs. The cdNS values for missense, truncating, and indel mutation pairs (149 significant pairs in Fig. 1b) are depicted in Fig. 1c. The cdNS values were also able to distinguish between the CO and ME gene pairs across the three mutation types again highlighting the overall concordance of the SYN-ANT relationship with CO-ME genomic representation of gene pairs.

To further validate our findings, we compared cdNS values of known gene pairs estimated using PanCancer mutational profiles with those from independent cohorts of MSK-IMPACT cohorts [21]. Among 517 known gene pairs, 133 pairs were available for mutation counts in the MSK-IMPACT dataset and we observed similar trends in segregating CO and ME gene pairs based on cdNS score, as shown in Additional file 2: Fig. S3. Furthermore, the cdNS values for 48 missense, 31 truncating, and 32 indel pairs whose ED are both significant in two datasets (TCGA and MSK-IMPACT) were significantly correlated suggesting that the functional relationships of gene pairs are consistent across databases (Fig. 1d). For example, cdNS scores of missense mutations of TCGA and MSK-IMPACT were significantly correlated (r = 0.75, P = 1.1e − 09) along with truncating mutations (r = 0.38, P = 0.035) and indels (r = 0.88, P = 5.4e − 11). These findings indicate that the ED-based SYN-ANT relationship of gene pairs remains consistent across various mutation databases, thus serving as a reliable indicator of their functionality and evolutionary relationship.

Identification of gene mutation pairs with ED

To identify gene pairs exhibiting ED, we selected a set of 312 genes composed of 220 genes and 92 genes identified from PanCancer and tumor type-specific mutational profiles, respectively (Additional File 1: Table S1). These genes exhibited significant overlap with known cancer-related genes [30], with 148 of them overlapping with a set of 719 known cancer-census genes. Additionally, 155 of these genes overlapped with 248 genes previously identified in known 517 gene pairs (where 248 gene members formed 517 gene pairs). We performed cdNS analyses for possible gene pairs of 312 genes using PanCancer mutational profiles. To assess the significance of cdNS scores, we employed permutation tests and implemented supplementary filters to account for tumor types and sample sizes (for detailed methods, refer to “Methods”). This analysis revealed 85 gene pairs with statistically significant ED, consisting of 41 missense gene pairs and 44 indel gene pairs. Among the pairs, 24 out of the 85 identified pairs (28.2%) overlapped with the 517 known gene pairs. For identified pairs, the dNdS ratios with or without mutation contexts are shown in a scatter plot (Fig. 2a) along with a barplot representing cdNS scores of the pairs (Fig. 2b). A comprehensive list of 85 gene pairs, along with related details can be found in Additional File 1:Supplementary Table S2. We observed that the smallest cdNS value was observed with IDH1 missense mutations in the presence of IDH2 mutation context (cdNS =  − 2.56) and the highest cdNS value was noted for ATRX indels with IDH1 mutation contexts (cdNS = 3.69), indicative of ANT and SYN relationship, respectively (arrows indicated, Fig. 2a, b). This finding suggests that IDH1 missense mutations occur approximately four times less frequently in genomes with IDH2 mutations compared to genomes without IDH2 mutations, indicating a potential antagonistic relationship or functional redundancy observed in brain tumors [33]. On the other hand, ATRX indels are found approximately ten times more frequently in genomes with IDH1 mutations compared to those without IDH1 mutations. This implies that ATRX losses, which lead to an alternative lengthening of telomeres, may not be sufficient on their own to drive tumor formation but may complement mutant IDH1 expression, suggesting their potential functional synergism [34].

Fig. 2
figure 2

PanCancer gene pairs under ED. a A total of 85 gene pairs identified in PanCancer mutational profiles are displayed for their dNdS ratios. The x-axis represents the dNdS ratios with mutation contexts (context +), while the y-axis represents the dNdS ratios without mutation contexts (context −). Blue and red dots denote missense and indel gene pairs, respectively. b cdNS scores of 85 gene pairs are shown in order of the cdNS scores. Blue and red lines represent the missense and indel gene pairs, respectively. c–f The dNdS ratios in the presence and absence of mutation contexts are separately illustrated in scatter plots for KRAS, TP53, PTEN, and IDH1 mutations. The y-axis represents the dNdS ratios with mutation contexts, while the x-axis represents the dNdS ratios without mutation contexts. Blue and red dots indicate missense and indel pairs, respectively

To further evaluate the impact of mutation contexts, we investigated four specific mutation contexts—KRAS, TP53, PTEN, and IDH1—that exhibited significant effects on gene pairs, with each occurring in 8, 7, 9, and 12 gene pairs, respectively. In the case of KRAS mutation contexts, we found that the dNdS ratios of eight gene pairs were substantially altered with KRAS mutations, indicating a functional association with KRAS mutations (Fig. 2c). For instance, the dNdS ratios of BRAF and EGFR mutations decreased (from 3.6 and 1.1 to 1.0 and − 0.9, respectively, on a log2 scale) in the presence of KRAS mutation contexts, suggesting an ANT relationship with KRAS mutations. Furthermore, SMAD4 and APC indels exhibited a SYN relationship with KRAS mutations, consistent with previous reports [35, 36]. Other mutations, such as TGIF1 loss, exhibited a SYN relationship with KRAS mutation contexts, consistent with their role in accelerating Kras-driven malignant transformation in the pancreas [37]. We also examined the impact of NRAS and HRAS mutation contexts on genes affected by KRAS mutation contexts using available cdNS scores for NRAS and HRAS mutation contexts (Additional File 2: Supplementary Fig. S4). Consistent observations of an ANT relationship with BRAF mutations (observed for both NRAS and HRAS mutation contexts) and a SYN relationship with APC mutations (for the NRAS mutation context) suggested that the ED relationship is largely preserved across members of the Ras gene family.

For TP53 mutation contexts (Fig. 2d), genes like ATRX encoding SWI/SNF family chromatin remodeling proteins known to cooperate with p53 deficiency, showed elevated dNdS ratios, indicating synergistic roles [38]. Additionally, previously reported synergistic gene pairs, such as CDKN2A-TP53, were observed as SYN pairs [25]. Among genes exhibiting SYN relationship with TP53 mutations, AJUBA and PRKCI have been reported to interact directly with p53 [39] and cooperate with TP53 losses across various tumor types [40], respectively. Conversely, as genes showing an ANT relationship with TP53 mutation contexts, GATA3 mutations have been previously reported for the synthetic lethality of GATA3 and MDM2 in breast cancers [41]. In addition, the CDH1 losses are also known to promote tumorigenesis with TP53 losses in endometrial cancers [42]. Our analyses also revealed a number of PTEN-synergistic mutations such as PIK3R1 and CCND1, e.g., PIK3R1 missense mutations dNdS ratios from 0.53 to 2.39 and CCND1 indels dNdS ratios from − 0.66 to 3.67, respectively with PTEN mutation contexts (Fig. 2e). Among the members in a phosphatidylinositol 3-kinase (PI3K) pathway, mutations of PIK3R1 disrupt the genome stabilizing roles of PTEN, potentially synergizing PTEN losses in endometrial cancers [43]. CCND1 overexpression has been often observed with PTEN alterations in lung cancers suggesting SYN relationship [44]. Although CCND1 amplification has been proposed as major alterations in cancer genomes, we observed both missense mutations and indels of CCND1 were associated with PTEN mutation contexts. In addition, the ANT presentation of PTEN and IDH1 mutations have been previously recognized for some tumor types such as gliomas [45]. Genes with synergistic relationship with IDH1 mutations in glioma such as ATRX have been previously reported [34], representing a glioma subgroup distinct from those with co-mutations of IDH1 with CIC and FUBP1 mutations [46] (Fig. 2f). Synergistic relationship of missense and indel TP53 mutations were observed within IDH1 mutation contexts, consistent with the known relationship in brain tumors where up to 70% of IDH1-mutant astrocytomas harbor TP53 mutations [47]. IDH2 ANT relationship suggests functional redundancy with IDH1 mutations in glioma pathogenesis [34].

Identification of genes pairs with tumor type-specific ED

The 312 genes examined using PanCancer profiles, were further evaluated across 20 tumor type-specific mutational profiles leading to the identification of 3870 gene pairs with tumor type-specific significant ED (see “Methods” for the selection criteria). The 3870 gene pairs with tumor type-specific ED, comprising 2523 missense, 587 truncating, and 760 indels, are presented for their dNdS ratios with or without mutation contexts (Fig. 3a) along with cdNS values (Fig. 3b). Detailed information of 3870 tumor type-specific gene pairs are available in Additional file 1: Table S3. The distribution of gene pairs across tumor types is illustrated (Fig. 3c). Tumor types with high tumor mutation burdens (TMB) including lung cancers and melanomas as well as those frequently showing microsatellite instability-high (MSI-H) such as colorectal, stomach, and endometrial cancers demonstrated a higher frequency of gene pairs with tumor type-specific ED. This suggests that the identification of gene pairs in tumor types with low mutation frequencies may not be exhaustive in current mutational profiles.

Fig. 3
figure 3

Tumor type-specific gene pairs under ED. a,b A total of 3870 mutation pairs gene pairs exhibiting tumor type-specific ED are depicted, showcasing their dNdS ratios and cdNS scores. c The prevalence of tumor type-specific gene pairs are shown across the examined tumor types. Colors represent the ED types (SYN and ANT) and mutation types (missense, truncating mutations, and indels). d For recurrent gene pairs of EPHA2-ATM and EP300-ATM, dNdS ratios are shown in tumor types where the gene pairs were observed. Shaded and unshaded boxes represent the dNdS ratios with or without mutation contexts, respectively. e Differences in variant allele frequency (VAF) are illustrated for SYN and ANT gene pairs as measured in PanCancer datasets. f VAF differences of gene pairs as estimated in the corresponding tumor types are shown separately across the tumor types. *, **, *** represent P < 0.05, P < 0.01, and P < 0.001, respectively. g Differences in tumor mutation burden (TMB) are presented in a boxplot (log10 scale). h TMB differences are shown across tumor types examined. Asterisks indicates the level of statistical significance

Among the gene pairs displaying the lowest cdNS values, indicating an ANT relationship, we noted that RB1 and RNF43 mutation contexts are antagonistic to IDH1 missense and APC truncating mutations in gliomas and colorectal cancers, respectively (ranked 5th and 6th with cdNS scores of − 3.83 and − 3.76). The ANT relationship of IDH1 and RB1 mutations in brain tumors may have arisen due to the mutation preference of low- and high-grade gliomas that are respectively enriched with IDH1 mutations and mutations in Rb pathways including RB1 mutations [48]. In colorectal cancers, RNF43 mutation contexts exhibit an ANT relationship with APC truncating mutations, particularly associated with MSI status, i.e., RNF43 mutations are prevalent in MSI-H tumors [49] but APC mutations are relatively enriched in MSS cases [23]. Therefore, some of tumor type-specific ANT relationships may reflect mutations that are enriched in specific tumor subtypes, such as the grade of gliomas and the microsatellite instability (MSI) status of colorectal cancers. The highest cdNS value (cdNS = 9.72) was observed for kidney cancers for TTN missense mutations in the context of TP53 mutations. Given that TP53 mutations are associated with elevated genomic instability and a higher TMB and TTN is the largest gene with a higher propensity for mutations, this correlation may be indicative of elevated genomic instability. The second highest cdNS values (cdNS = 7.312) was noted for NHLRC1 mutation contexts for FAT1 truncating mutations in head and neck cancers. NHLRC1 overexpression is known to stimulate cellular proliferation and invasion, possibly acting as AKT downstream effectors [50]. These phenotypic effects are similar to those associated with FAT1 losses [51] indicative of their potential synergistic roles.

We also identified 92 gene pairs that occurred in more than one tumor type. The two most frequently occurring gene pairs were observed in three tumor types, EPHA2-ATM and EP300-ATM gene pairs. EPHA2 and EP300 mutations consistently show decreased dNdS ratios in the presence of ATM mutation contexts compared to those without ATM mutation contexts across three tumor types, respectively (Fig. 3d). While ATM-encoded peptides are known for their roles in activating the homologous recombination pathway for DNA repair [52, 53], recent findings suggest that ATM might also counteract the non-homologous end-joining (NHEJ) process by removing DNA-PKcs [54]. Furthermore, EphA2, encoded by EPHA2 have a role in DNA repair through direct binding to DNA-PKcs, and thus, mutation of EPHA2 can consequently affect the NHEJ pathway [55]. This functional connection with ATM losses could elucidate the observed ANT relationship EPHA2 and ATM mutations across tumor types. Similarly, regarding EP300 mutations, there is evidence suggesting a functional association where ATM losses lead to the failure of p300 protein phosphorylation [56], indicating functional redundancy between ATM losses and p300 deficiency.

We further investigated the association between ED of gene pairs and variant allele frequency (VAF), a metric indicating the clonality for mutations. For each gene pair, we calculated the median VAF of mutations for each gene. We then determined the difference between the median VAF values (ΔVAF), i.e., VAF of the partner gene minus VAF of the context gene, and compared them with respect to SYN and ANT relationship of gene pairs (Fig. 3e). Our analysis revealed a significant decrease and increase of ΔVAF values for SYN and ANT gene pairs (P = 1.5e − 27 and P = 3.9e − 23, t-test, respectively). A positive ΔVAF indicates that mutation contexts in ANT relationships are likely to be subclonal, while mutations of their partner gene are likely to be clonal. The difference in VAF between SYN and ANT gene pairs is consistently observed across tumor types (Fig. 3f). We also calculated the difference in TMB (ΔTMB) between genomes harboring both mutations and those with singleton mutations in paired genes. We observed that genomes harboring mutations in both genes of ANT gene pairs had a higher TMB as indicated by positive ΔTMB, while the opposite was observed for SYN gene pairs (P = 6.9e − 50 and P = 7.9e − 176, respectively) (Fig. 3g). This is also consistently observed across tumor types (Fig. 3h). Considering the specificity of tumor subtypes, we examined the relationship of VAF and TMB with SYN-ANT gene pairs in glioma subtypes and MSI status in colorectal cancers (Additional file 2: Supplementary Fig. 5), revealing a consistent association. These findings offer insights into the mutation acquisition in gene pairs with an ANT relationship, which are less likely to be fixed in cancer genomes due to negative selection. For example, the impact of subclonal, ANT mutation contexts on the acquisition of mutations in their partner genes may be relatively small compared to clonal SYN mutation contexts. The impact of negative selection is primarily attributed to clonal mutations rather than subclonal mutations [57]. Moreover, mutations in ANT relationship are tolerant in genomes with high TMB, which are relatively tolerant to genomic alterations.

The association of genetic dependency and ED

We next investigated the relation between the impact of genetic perturbation of cancer cell lines with respect to their mutation configuration of genes in ED relationship. To explore the relationship between ED and genetic dependencies, we employed three databases of genetic screenings. One, AVANA [26] was based on CRISPR (clustered regularly interspaced short palindromic repeats). The other two, DEMETER2 [27] and DRIVE [28], were based on short hairpin or shRNA. Across each database, we collected data from cell lines harboring mutations targeted by gene silencing and further distinguished them into those with or without the mutation contexts. The effect size was determined by comparing the median of cell viability between these identified two groups of cell lines as previously described [22]. For example, the impact of KRAS silencing was assessed in cell lines with KRAS mutations and the effect size corresponding to KRAS-NRAS gene pairs was calculated by comparing the cell viability of those with and without NRAS mutations (≥ 2 cell lines for both groups were considered). We then determined the correlation between these effect sizes with cdNS scores of the corresponding gene pairs, as estimated in PanCancer mutational profiles, to evaluate the relationship between ED and genetic dependency.

It has been assumed that SYN and ANT mutation contexts will strengthen and lessen the impact of gene knockouts leading to decrease and increase of cell viability (here, compromising and rescuing effects on cell viability, respectively), since SYN gene pairs are implicated in oncogenic addiction and ANT gene pairs can rescue the gene silencing effects of partner mutations [22]. Thus, mutation contexts with high cdNS scores are expected to exhibit decreased effect sizes with knock-out mutations, while those with low cdNS scores are expected to exhibit elevated effect sizes. This expectation aligns with our observation that mutation contexts corresponding to SYN and ANT gene pairs showed substantial difference in effect size across databases for the selected gene pairs available for gene dependencies, i.e., AVANA (36 SYN–8 ANT pairs, P = 0.114; t-test), DEMETER2 (48 SYN–19 ANT pairs, P = 0.157; t-test), and DRIVE (35 SYN–14 ANT pairs, P = 0.403; t-test). The relationship was largely consistent for gene pairs examined in tumor type-specific datasets (Additional File 2: Supplementary Fig. 6).

We further compiled a list of 38 gene pairs that appeared more than once in three databases. These pairs showed a moderate yet substantial inverse relationship between ED and genetic dependencies, as indicated by cdNS scores and average effect size, respectively. (r =  − 0.29, P = 0.073; Fig. 4b, Additional file 1: Supplementary Table 4). These gene pairs were further categorized based on their impact on cell viability (either rescuing or compromising with effect sizes greater than 0 or less than 0, respectively) or ED (SYN and ANT with respect to cdNS). This classification resulted in four gene pair categories. We then estimated the pathway concordance of gene pairs (i.e., the number of molecular terms with gene members in the pair both included) and found that rescuing-antagonistic (Res.-ANT) gene pairs exhibited higher pathway concordance, followed by compromising-antagonistic (Com.-ANT) gene pairs (Fig. 4c). Therefore, the rescuing effects of gene pairs in ANT relationships can be largely ascribed to functional redundancy, as evidenced by a high level of pathway dependency, the degree of shared gene members within similar functional pathways. Further analysis based on whether the gene pairs represent the pairs of oncogenes or tumor suppressors showed that Res.-ANT pairs were more likely to be pairs of oncogenes (Fig. 4d). Thus, functionally redundant oncogene pairs appearing as ANT gene pairs can also provide rescuing effects in cell viability upon silencing of their paired mutations.

Fig. 4
figure 4

ED and genetic dependencies. a Effect sizes and dN/dS ratios for individual mutation pairs gene pairs are displayed across three databases of genetic dependencies (AVANA, DEMETER2, and DRIVE), with the y- and x-axis representing the mutation pairs and their respective dNdS ratios, respectively. The level of correlation and statistical significance by t-test is indicated. b Thirty eight mutation pairs gene pairs occurring more than once across databases are selected, and their effect sizes and dNdS ratios are presented. Four mutational categories are shown for rescuing and compromising effects (effect size > 0 and < 0, respectively) and synergistic and redundant/antagonistic effects (cdNS scores > 0 and < 0, respectively). c Boxplots depict pathway concordances for the four categories of mutation pairs. d The mutations in pairs are analyzed to determine whether they are oncogenes (OG) or tumor suppressor genes (TSG), and the abundance across the four categories of mutation pairs gene pairs is shown. e–h For selected examples of KRAS*-NRAS, BRAF*-KRAS, KRAS*-BRAF, and TCF7L2*-KRAS mutation pairs (where asterisk indicates the genes with knockouts), the cell viability of selected cancer cell lines with respect to their mutational states of gene members in the pairs are shown. Cell viability was used as available in AVANA database. Asterisks indicate knock-out genes with gene silencing. The dNdS ratios of mutations in genomes with or without the mutation contexts are also shown in barplots

The cell viability is shown for selected gene pairs for one example of KRAS*-NRAS gene pairs (asterisk indicating silenced genes in cell lines) serving as an example of rescuing-antagonistic (Res.-ANT) gene pair (AVANA database, Fig. 4e). This specific case demonstrates that silencing mutant KRAS substantially impacts cell viability in NRAS wild type cells (KRASmt/NRASwt), while NRAS mutations counteract the effects of KRAS inhibition in double mutant cells (NRASmt/KRASmt; P = 0.296; t-test). This pattern was similarly observed when inhibiting BRAF mutants in both KRAS-mutant and -wildtype cells (P = 3.1e − 06; t-test, Fig. 4f). Interestingly, unlike the BRAF*-KRAS relationship, KRAS*-BRAF exhibits compromising effects, where BRAF mutations fail to mitigate the impact of KRAS knockout on KRAS mutant cells (Fig. 4g). It is presumed that the hierarchy of genes within the signaling cascade leads to varied effects on cell viability, i.e., in the Ras-Raf signaling pathway, mutations of BRAF that are downstream of Kras in the pathway, do not rescue KRAS mutant cells from KRAS knockouts. The cellular effects with respect to the mutation contexts are consistently observed across database (DEMETER2 and DRIVE, Additional File 2: Supplementary Fig S7).

The association of the drug sensitivity and ED

We next examined whether the mutation contexts can influence the pharmacological effects of targeted agents according to their ED with the targeting genes. For the analysis, we focused on agents targeting EGFR and BRAF that are used in lung adenocarcinomas and melanomas (LUAD and SKCM), respectively. Among 279 gene pairs with tumor type-specific significant ED in LUAD, we selected 19 gene pairs including EGFR as either context genes or their partner genes. The cdNS values of 19 gene pairs are shown in Fig. 5a, highlighting ANT relationship of EGFR with KRAS mutations (i.e., cdNS scores of − 4.3 and − 3.9 for KRAS missense mutations with EGFR mutation context and for EGFR missense mutations with KRAS mutation contexts, respectively, arrows indicated in Fig. 5a). This ANT relationship suggests that KRAS mutations may compensate the EGFR mutations that are inhibited by EGFR inhibitors, thereby reducing the inhibitory effects. To demonstrate this, we examined IC50 (half maximal inhibitory concentration) levels of LUAD cell lines to EGFR inhibitors with respect to the mutational status of EGFR and KRAS as obtained in GDSC (Genomics of Drug Sensitivity in Cancer) [31] (Fig. 5b). Compared to EGFR wildtypes, EGFR mutant cell lines showed sensitivity to EGFR inhibitors only in the absence of KRAS mutations (EGFRmt/KRASwt) and the IC50 levels of cell lines harboring both EGFR and KRAS mutations (EGFRmt/KRASmt) were comparable to those of EGFR wild type cell lines across six EGFR inhibitors. A similar relationship was observed between drug sensitivity to BRAF inhibitors and the ED of BRAF-NRAS mutations in SKCM cell lines. For example, among 11 gene pairs including BRAF observed in SKCM, ANT relationship was noted with NRAS mutations (cdNS scores of − 3.6 and − 1.8, arrows in Fig. 5c). We also observed that BRAFmt/NRASmt SKCM cell lines showed comparable IC50 levels to those of BRAF wild type cell lines across three BRAF inhibitors (Fig. 5d). These findings suggest that for mutations of genes in ED can potentially alter drug sensitivity where the inhibitory impact of targeting agents is compromised by mutations that exhibit an ANT-relationship with targeted genes.

Fig. 5
figure 5

ED with pharmacological impact. a cdNS scores of gene pairs containing EGFR as gene members, as estimated in lung adenocarcinoma (LUAD) mutational profiles. Blue, orange, and red denote partner genes harboring significant ED for missense, truncating mutations, and indels, respectively, with corresponding mutation contexts. Two arrows indicate pairs of EGFR and KRAS mutations as significant ANT gene partners. b IC50 levels across six EGFR inhibitors are displayed for four types of LUAD cell lines based on the mutational states of EGFR and KRAS (mt and wt representing mutant and wild types, respectively). Drug names and significance levels estimated between EGFRmt/KRASwt (blue) and EGFRmt/KRASmt (orange) by t-test are illustrated. The number of cell lines is also indicated. IC50 values of celllines with EGFRwt/KRASwt (grey) and EGFRwt/KRASmt (black) are separately shown. c The cdNS scores are shown for BRAF-harboring gene pairs identified in melanoma (SKCM) mutational profiles. Arrows indicate ANT relationship between the BRAF mutations and NRAS mutations. d IC50 values with BRAF inhibitors are shown for SKCM cell lines with respect to the mutational status of BRAF and NRAS mutations. e The effect size (the differential of IC50 values of cell lines between those with or without the mutations paired with mutation contexts) is shown against cdNS scores of mutation pairs (y- and x-axis, respectively). Red and blue dots represent the effect size as estimated from GDSC and CCLE database. Pairs of two databases were linked by arrows for selected mutation pairs. f,g IC50 values are similarly shown for mutation pairs of PRSS1-EGFR and CTCF-BRAF for LUAD and SKCM celllines, respectively

To demonstrate the relationship between ED and drug sensitivity, pharmacological effects were further evaluated across tumor type-specific gene pairs. For this analysis, effect sizes were estimated as the IC50 differential between tumor type-matched cell lines (i.e., those harboring mutations corresponding to the partner genes of mutation contexts) with or without mutations in genes of the mutation context. For example, IC50 differential between EGFR inhibitor-treated EGFR-mutant LUAD cell lines with or without KRAS mutation context (no less than two cell lines) was calculated as effect size across individual EGFR inhibitors. Then, they were compared with cdNS scores of EGFR-KRAS gene pair as evaluated in LUAD mutational profiles. We considered drugs whose targets correspond to the partner mutations, obtaining 18 gene pairs whose cdNS scores and effect sizes are available both in two database of GDSC and CCLE (Cancer Cell Line Encyclopedia) database [32]. The selected pairs are shown in scatter plots (Fig. 5e, those of GDSC and CCLE datasets shown in red and blue, respectively). Substantial level of correlation was observed although not in significant levels (r =  − 0.25, P = 0.146). Among the gene pairs, we observed other candidates of rescuing pairs of PRSS1 and CTCF mutations occurring with EGFR and BRAF mutation contexts in LUAD and SKCM, respectively. When PRSS1 and CTCF mutations were examined for their impact on IC50 of cell lines treated with EGFR and BRAF inhibitors, we observed similar compromising effects, e.g., their co-mutations with targeted mutations showed an elevation in the IC50 of the cell lines (Fig. 5f, g). This suggests that PRSS1 and CTCF mutations exhibit compensatory effects comparable to EGFR and BRAF mutations in the therapeutic response to EGFR and BRAF inhibitors in LUAD and SKCM cell lines.

Discussion

In this study, we adopted evolutionary measures based on dNdS (nonsynonymous-to-synonymous) ratios and their derived cdNS (context-dependent dNdS ratios) scores to assess the ED (evolutionary dependency) of gene mutation pairs in cancer genomes. Although positive and negative selective pressures have influenced the development of acquired mutations, ultimately shaping the mutation landscape of cancer genomes [21, 58, 59], the relationship between these mutations remains largely unknown. One assumption underlying the study is that the selection of mutations in cancer genomes is an evolutionarily constrained and context-dependent process leaving genomic footprints in snapshots of cancer genomes. Thus, we suggest that the comparison of dNdS ratios of mutations in relation to the mutation context, can serve as a measure of ED. Previous studies have proposed similar methodologies, e.g., the comparison of two dNdS ratios between clonal and subclonal mutations in lung cancers holds potential clinical significance [60]. However, our extensive analyses on a broader set of genes and tumor types uncovered a range of functional and cellular aspects of gene pairs influenced by ED across various tumor types. Our newly introduced ED measures referred to as cdNS scores, encapsulate the essence of dNdS ratios, i.e., they not only categorize the types of gene pairs (synergistic and antagonistic pairs, SYN and ANT) but also offer quantitative assessments of ED between two genes. Upon applying ED to known gene pairs identified by genomic distribution (517 CO-ME gene pairs as co-occurring and mutually exclusive ones) [25], we observed a strong correlation between the CO and ME categories of CO-ME gene pairs and cdNS-based SYN and ANT pairs. Furthermore, the variability of ED for the observed gene pairs also underscores the necessity for quantitative metrics like cdNS scores. It should also be noted that the ANT gene pairs in PanCancer analyses might appear between two tumor type-specific genes simply due to their differing tumor type specificities. Although we have used weighting factors to mitigate this impact [25], future research should focus on further minimizing the influence of tumor type-specific effects when identifying PanCancer-ANT gene pairs.

Utilizing cdNS scores as novel ED measures, we evaluated gene pairs in cancer genomes with respect to other molecular and genomic features including VAF (variant allele frequencies) and TMB (tumor mutation burdens) as well as genetic dependency and pharmacological perturbations. The quantitative nature of cdNS scores reveals significant variation in the degree of ED, often exceeding tenfold positive or negative selective pressure for mutational acquisition according to the mutation contexts. This relationship often leads to the changes in the cell survival of cell lines with knock-outs, which are consistent with the types of ED where ANT and SYN gene pairs tend to show rescuing or compromising effect on the survival of cells. For example, a KRAS-NRAS ANT gene pair with low cdNS score rescues cell survival, indicating that NRAS mutation contexts rescue the impact of KRAS knockouts on KRAS-mutant cell lines. In contrast, KRAS mutation contexts decreased the survival of TCF7L2-mutant cells with TCF7L2 knockout, suggestive of potential synthetic lethality [61]. Of note, the relationship of ED and genetic dependency depends on the hierarchy of gene members in the signaling cascade [62] as shown with the relationship of BRAF and KRAS. For example, the KRAS mutation contexts showed rescuing effects on survival of BRAF-mutant cells with BRAF knock-outs, whereas BRAF mutation contexts fail to rescue the effect of KRAS knock-outs on KRAS-mutant cells. Given the hierarchy within the receptor tyrosine kinase signaling pathway including Ras-Raf, mutation contexts of Raf exert relatively minor effects on the mutations and knock-outs of its upstream Ras, while Ras mutations exhibit greater influence as upstream mutation contexts [63].

The dNdS ratio-based ED can also highlight resistance-conferring mutations to targeted agents such as KRAS and NRAS mutations for EGFR and BRAF inhibitors, respectively [64, 65]. This suggests that ED may assist in selecting patients who might benefit from targeted agents or predicting disease recurrence [15]. For example, while the impact of KRAS and NRAS mutations on the EGFR and BRAF inhibitors have been well-recognized, the ED-based analyses further demonstrated genes with similar ED and pharmacological effects with KRAS and NRAS mutations. We observed that PRSS1 and CTCF mutations in lung adenocarcinomas and skin melanomas show similar resistance-conferring effects to KRAS and NRAS mutations when treated with EGFR and BRAF inhibitors. PRSS1 encodes pancreatic serine proteinase, trypsin-1 whose expression has been associated with the sensitivity to EGFR inhibitors of cetuximab in colorectal cancers, and its inhibition reduces the tumor growth [66]. CTCF encoding CCCTC-binding factor mediates transcriptional regulation of DUSP6 and when inactivated, can activate MAPK signaling [67], which explains the resistance mechanisms to BRAF inhibitors with CTCF losses.

One limitation of our study is that ED analyses are limited to somatic mutations, whereas cancer genomes include other types of genomic alterations, such as DNA copy number alterations and chromosomal rearrangements [68]. One important assumption of dNdS ratios is that alterations that can be distinguished into non-neutral vs. neutral changes. Compared to nonsynonymous-vs.-synonymous mutation calls, other types of genomic alterations cannot be readily distinguished as non-neutral or neutral because large-scale chromosomal changes often involve many genes and have non-binary nature (e.g., copy number states). The heterogeneity of cancer genomes sharing multiple types of alterations in single locus also complicates the determination of functionality for large-scale genomic alterations [69]. Nevertheless, the consideration of genomic alterations in addition to somatic mutations will be essential for gaining a comprehensive understanding of the mutational landscape and identification of potential interactions between various types of genomic alterations. Another assumption underlying our study is that mutations in cancer genomes do not simultaneously occur, so that the first mutational hit in the mutation pair provides the ED for the second hit. However, kataegis is an example that violates this assumption, because a number of mutations occur simultaneously in localized genomic regions [70]. Since this genomic event has unique properties such as local adjacency and common mutation signatures, it may be feasible that somatic mutations can be categorized further [71] and subject to cdNS-based ED analyses. In addition, it should be considered that our cell line-based results might reflect different ED compared to those observed in primary tumors, as in vitro system lacks immune cells and has abundant cultural resources.

In our study, mutation contexts were established using genomes containing non-silent mutations of genes, without discrimination based on types of mutations. Thus, there is still a potential for considering distinct mutation types within mutation contexts and incorporating functionality measures to refine the ED of mutation pairs. For example, hotspot missense mutations [72] and those categorized based on functional measures [3] can be further considered in evaluating dNdS ratios and cdNS scores. Our analysis of three cell lines harboring mutations on both KRAS and EGFR (Fig. 5b), all KRAS mutations are located on known hotspots (12th and 13th amino acid residues), while EGFR mutations were near or outside the known hotspots mutations [72] raising a concern related with the functionality of observed gene pairs. Although our current methods do not integrate site-level information, it prompts further investigation into whether integrating functionally filtered mutational profiles could refine ED relationship. However, it is important to note the limited availability of genomes with mutation contexts since the mutations occur in a small fraction of cancer genomes. The limited number of available datasets may introduce potential biases in this type of analysis particularly when focusing on the pairs of genomic events (e.g., double mutants) that are often extremely scarce in public resources such as the CCLE database. This scarcity inadvertently highlights molecular pathways whose genes frequently harbor somatic mutations, such as the Ras pathway, underscoring the need for additional resources to draw robust conclusions. In addition, one observation of our study is the prevalence of tumor type-specific gene pairs among those with high tumor mutational burden (TMB). While it is possible that the hypermutated genomes may harbor an increased number of functional gene pairs, it is still conceivable that the current list of gene pairs may not be fully comprehensive and could benefit from larger mutation databases in the future [3, 73].

Conclusions

In summary, our study provides evolutionary perspectives on functional gene pairs in cancer genomes by employing dNdS ratio-based ED analyses. The dNdS ratio as well as novel ED measure of cdNS score was able to distinguish SYN and ANT gene pairs along with ED reflecting their functionality. The ED of gene pairs were examined for associations with other evolutionary and phenotypic features, revealing that ED of gene pairs represents unique functional signatures of mutations in cancer genomes. By employing cell line-based studies of genetic dependencies and pharmacological perturbation, ED guided the selection of candidates underlying synthetic lethality and drug-sensitizing or resistant mutations.

Availability of data and materials

Somatic mutation calls of TCGA consortium and related information were downloaded from TCGA data portal [74]. Panel-based mutational profiles were downloaded from cBioPortal link [75]. Cancer-related genes were used as Cancer Census genes [76]. dvdscv R packages were downloaded from Github database [77]. Gene dependency scores were obtained from Dependency Map (DepMap) gene essentiality screenings [78]. Specifically, AVANA scores (18q3_v3, 17 k genes across 485 cell lines) and DEMETER2 scores (v2, 17 k genes across 712 cell lines) were obtained along with DRIVE database (7.8 k genes across 398 cell lines, version5, May 2019). Drug response data of cancer cell lines were obtained from GDSC (Genomics of Drug Sensitivity in Cancer) database as a combined set of two dataset versions (GDSC1 and GDSC2) [79]. Other pharmacological cancer cell line data were obtained from CCLE (Cancer Cell Line Encyclopedia) databases [80]. The R code used to derive the cdNS scores are available (https://github.com/TkimLab/cdNS) [81].

References

  1. 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. 2018;173(371–385): e318.

    Google Scholar 

  2. Stratton MR, Campbell PJ, Futreal PA. The cancer genome. Nature. 2009;458:719–24.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  3. Martinez-Jimenez F, Muinos F, Sentis I, Deu-Pons J, Reyes-Salazar I, Arnedo-Pac C, Mularoni L, Pich O, Bonet J, Kranas H, et al. A compendium of mutational cancer driver genes. Nat Rev Cancer. 2020;20:555–72.

    Article  CAS  PubMed  Google Scholar 

  4. Dees ND, Zhang Q, Kandoth C, Wendl MC, Schierding W, Koboldt DC, Mooney TB, Callaway MB, Dooling D, Mardis ER, et al. MuSiC: identifying mutational significance in cancer genomes. Genome Res. 2012;22:1589–98.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  5. Lawrence MS, Stojanov P, Polak P, Kryukov GV, Cibulskis K, Sivachenko A, Carter SL, Stewart C, Mermel CH, Roberts SA, et al. Mutational heterogeneity in cancer and the search for new cancer-associated genes. Nature. 2013;499:214–8.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  6. Martincorena I, Raine KM, Gerstung M, Dawson KJ, Haase K, Van Loo P, Davies H, Stratton MR, Campbell PJ. Universal patterns of selection in cancer and somatic tissues. Cell. 2017;171(1029–1041):e1021.

    Google Scholar 

  7. Loeb KR, Loeb LA. Significance of multiple mutations in cancer. Carcinogenesis. 2000;21:379–85.

    Article  CAS  PubMed  Google Scholar 

  8. Yates LR, Campbell PJ. Evolution of the cancer genome. Nat Rev Genet. 2012;13:795–806.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  9. Sansom OJ, Meniel V, Wilkins JA, Cole AM, Oien KA, Marsh V, Jamieson TJ, Guerra C, Ashton GH, Barbacid M, Clarke AR. Loss of Apc allows phenotypic manifestation of the transforming properties of an endogenous K-ras oncogene in vivo. Proc Natl Acad Sci U S A. 2006;103:14122–7.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  10. Bardeesy N, Cheng KH, Berger JH, Chu GC, Pahler J, Olson P, Hezel AF, Horner J, Lauwers GY, Hanahan D, DePinho RA. Smad4 is dispensable for normal pancreas development yet critical in progression and tumor biology of pancreas cancer. Genes Dev. 2006;20:3130–46.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  11. Ciriello G, Cerami E, Sander C, Schultz N. Mutual exclusivity analysis identifies oncogenic network modules. Genome Res. 2012;22:398–406.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  12. Kim JW, Botvinnik OB, Abudayyeh O, Birger C, Rosenbluh J, Shrestha Y, Abazeed ME, Hammerman PS, DiCara D, Konieczkowski DJ, et al. Characterizing genomic alterations in cancer by complementary functional associations. Nat Biotechnol. 2016;34:539–46.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  13. Unni AM, Lockwood WW, Zejnullahu K, Lee-Lin SQ, Varmus H. Evidence that synthetic lethality underlies the mutual exclusivity of oncogenic KRAS and EGFR mutations in lung adenocarcinoma. Elife. 2015;4:e06907.

    Article  PubMed  PubMed Central  Google Scholar 

  14. Whittaker SR, Theurillat JP, Van Allen E, Wagle N, Hsiao J, Cowley GS, Schadendorf D, Root DE, Garraway LA. A genome-scale RNA interference screen implicates NF1 loss in resistance to RAF inhibition. Cancer Discov. 2013;3:350–62.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  15. Deng Y, Luo S, Deng C, Luo T, Yin W, Zhang H, Zhang Y, Zhang X, Lan Y, Ping Y, et al. Identifying mutual exclusivity across cancer genomes: computational approaches to discover genetic interaction and reveal tumor vulnerability. Brief Bioinform. 2019;20:254–66.

    Article  PubMed  Google Scholar 

  16. Glaab E, Baudot A, Krasnogor N, Schneider R, Valencia A. EnrichNet: network-based gene set enrichment analysis. Bioinformatics. 2012;28:i451–7.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  17. Leiserson MD, Vandin F, Wu HT, Dobson JR, Eldridge JV, Thomas JL, Papoutsaki A, Kim Y, Niu B, McLellan M, et al. Pan-cancer network analysis identifies combinations of rare somatic mutations across pathways and protein complexes. Nat Genet. 2015;47:106–14.

    Article  CAS  PubMed  Google Scholar 

  18. Shrestha R, Hodzic E, Sauerwald T, Dao P, Wang K, Yeung J, Anderson S, Vandin F, Haffari G, Collins CC, Sahinalp SC. HIT’nDRIVE: patient-specific multidriver gene prioritization for precision oncology. Genome Res. 2017;27:1573–88.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  19. Hoadley KA, Yau C, Hinoue T, Wolf DM, Lazar AJ, Drill E, Shen R, Taylor AM, Cherniack AD, Thorsson V, et al. Cell-of-Origin Patterns Dominate the Molecular Classification of 10,000 Tumors from 33 Types of Cancer. Cell. 2018;173(291–304): e296.

    Google Scholar 

  20. Cancer Genome Atlas Research N, Weinstein JN, Collisson EA, Mills GB, Shaw KR, Ozenberger BA, Ellrott K, Shmulevich I, Sander C, Stuart JM. The Cancer Genome Atlas Pan-Cancer analysis project. Nat Genet. 2013;45:1113–20.

    Article  Google Scholar 

  21. Zehir A, Benayed R, Shah RH, Syed A, Middha S, Kim HR, Srinivasan P, Gao J, Chakravarty D, Devlin SM, et al. Mutational landscape of metastatic cancer revealed from prospective clinical sequencing of 10,000 patients. Nat Med. 2017;23:703–13.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  22. Mina M, Iyer A, Tavernari D, Raynaud F, Ciriello G. Discovering functional evolutionary dependencies in human cancers. Nat Genet. 2020;52:1198–207.

    Article  CAS  PubMed  Google Scholar 

  23. Cancer Genome Atlas N. Comprehensive molecular characterization of human colon and rectal cancer. Nature. 2012;487:330–7.

    Article  Google Scholar 

  24. Cancer Genome Atlas Research N, Brat DJ, Verhaak RG, Aldape KD, Yung WK, Salama SR, Cooper LA, Rheinbay E, Miller CR, Vitucci M, et al. Comprehensive, integrative genomic analysis of diffuse lower-grade gliomas. N Engl J Med. 2015;372:2481–98.

    Article  Google Scholar 

  25. Mina M, Raynaud F, Tavernari D, Battistello E, Sungalee S, Saghafinia S, Laessle T, Sanchez-Vega F, Schultz N, Oricchio E, Ciriello G. Conditional selection of genomic alterations dictates cancer evolution and oncogenic dependencies. Cancer Cell. 2017;32(155–168): e156.

    Google Scholar 

  26. Meyers RM, Bryan JG, McFarland JM, Weir BA, Sizemore AE, Xu H, Dharia NV, Montgomery PG, Cowley GS, Pantel S, et al. Computational correction of copy number effect improves specificity of CRISPR-Cas9 essentiality screens in cancer cells. Nat Genet. 2017;49:1779–84.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  27. McFarland JM, Ho ZV, Kugener G, Dempster JM, Montgomery PG, Bryan JG, Krill-Burger JM, Green TM, Vazquez F, Boehm JS, et al. Improved estimation of cancer dependencies from large-scale RNAi screens using model-based normalization and data integration. Nat Commun. 2018;9:4610.

    Article  PubMed  PubMed Central  Google Scholar 

  28. McDonald ER 3rd, de Weck A, Schlabach MR, Billy E, Mavrakis KJ, Hoffman GR, Belur D, Castelletti D, Frias E, Gampa K, et al. Project DRIVE: a compendium of cancer dependencies and synthetic lethal relationships uncovered by large-scale, deep RNAi screening. Cell. 2017;170(577–592):e510.

    Google Scholar 

  29. Subramanian A, Tamayo P, Mootha VK, Mukherjee S, Ebert BL, Gillette MA, Paulovich A, Pomeroy SL, Golub TR, Lander ES, Mesirov JP. Gene set enrichment analysis: a knowledge-based approach for interpreting genome-wide expression profiles. Proc Natl Acad Sci U S A. 2005;102:15545–50.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  30. Sondka Z, Bamford S, Cole CG, Ward SA, Dunham I, Forbes SA. The COSMIC Cancer Gene Census: describing genetic dysfunction across all human cancers. Nat Rev Cancer. 2018;18:696–705.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  31. Iorio F, Knijnenburg TA, Vis DJ, Bignell GR, Menden MP, Schubert M, Aben N, Goncalves E, Barthorpe S, Lightfoot H, et al. A landscape of pharmacogenomic interactions in cancer. Cell. 2016;166:740–54.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  32. Barretina J, Caponigro G, Stransky N, Venkatesan K, Margolin AA, Kim S, Wilson CJ, Lehar J, Kryukov GV, Sonkin D, et al. The cancer cell line encyclopedia enables predictive modelling of anticancer drug sensitivity. Nature. 2012;483:603–7.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  33. Yan H, Parsons DW, Jin G, McLendon R, Rasheed BA, Yuan W, Kos I, Batinic-Haberle I, Jones S, Riggins GJ, et al. IDH1 and IDH2 mutations in gliomas. N Engl J Med. 2009;360:765–73.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  34. Mukherjee J, Johannessen TC, Ohba S, Chow TT, Jones L, Pandita A, Pieper RO. Mutant IDH1 cooperates with ATRX loss to drive the alternative lengthening of telomere phenotype in glioma. Cancer Res. 2018;78:2966–77.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  35. Bera A, Zhao S, Cao L, Chiao PJ, Freeman JW. Oncogenic K-Ras and loss of Smad4 mediate invasion by activating an EGFR/NF-kappaB Axis that induces expression of MMP9 and uPA in human pancreas progenitor cells. PLoS ONE. 2013;8:e82282.

    Article  PubMed  PubMed Central  Google Scholar 

  36. Fearon ER, Vogelstein B. A genetic model for colorectal tumorigenesis. Cell. 1990;61:759–67.

    Article  CAS  PubMed  Google Scholar 

  37. Weng CC, Hsieh MJ, Wu CC, Lin YC, Shan YS, Hung WC, Chen LT, Cheng KH. Loss of the transcriptional repressor TGIF1 results in enhanced Kras-driven development of pancreatic cancer. Mol Cancer. 2019;18:96.

    Article  PubMed  PubMed Central  Google Scholar 

  38. Oppel F, Tao T, Shi H, Ross KN, Zimmerman MW, He S, Tong G, Aster JC, Look AT. Loss of atrx cooperates with p53-deficiency to promote the development of sarcomas and other malignancies. PLoS Genet. 2019;15:e1008039.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  39. Xu B, Li Q, Zhang J, Chen F. LIM domain-containing protein Ajuba inhibits chemotherapy-induced apoptosis by negatively regulating p53 stability in colorectal cancer cells. Mol Oncol. 2023;17:1678–91.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  40. Sarkar S, Bristow CA, Dey P, Rai K, Perets R, Ramirez-Cardenas A, Malasi S, Huang-Hobbs E, Haemmerle M, Wu SY, et al. PRKCI promotes immune suppression in ovarian cancer. Genes Dev. 2017;31:1109–21.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  41. Bianco G, Coto-Llerena M, Gallon J, Kancherla V, Taha-Mehlitz S, Marinucci M, Konantz M, Srivatsa S, Montazeri H, Panebianco F, et al. GATA3 and MDM2 are synthetic lethal in estrogen receptor-positive breast cancers. Commun Biol. 2022;5:373.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  42. Stodden GR, Lindberg ME, King ML, Paquet M, MacLean JA, Mann JL, DeMayo FJ, Lydon JP, Hayashi K. Loss of Cdh1 and Trp53 in the uterus induces chronic inflammation with modification of tumor microenvironment. Oncogene. 2015;34:2471–82.

    Article  CAS  PubMed  Google Scholar 

  43. Cheung LW, Hennessy BT, Li J, Yu S, Myers AP, Djordjevic B, Lu Y, Stemke-Hale K, Dyer MD, Zhang F, et al. High frequency of PIK3R1 and PIK3R2 mutations in endometrial cancer elucidates a novel mechanism for regulation of PTEN protein stability. Cancer Discov. 2011;1:170–85.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  44. Dragoj M, Milosevic Z, Bankovic J, Dinic J, Pesic M, Tanic N, Stankovic T. Association of CCND1 overexpression with KRAS and PTEN alterations in specific subtypes of non-small cell lung carcinoma and its influence on patients’ outcome. Tumour Biol. 2015;36:8773–80.

    Article  CAS  PubMed  Google Scholar 

  45. Choi SW, Lee Y, Shin K, Koo H, Kim D, Sa JK, Cho HJ, Shin HM, Lee SJ, Kim H, et al. Mutation-specific non-canonical pathway of PTEN as a distinct therapeutic target for glioblastoma. Cell Death Dis. 2021;12:374.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  46. Jiao Y, Killela PJ, Reitman ZJ, Rasheed AB, Heaphy CM, de Wilde RF, Rodriguez FJ, Rosemberg S, Oba-Shinjo SM, Nagahashi Marie SK, et al. Frequent ATRX, CIC, FUBP1 and IDH1 mutations refine the classification of malignant gliomas. Oncotarget. 2012;3:709–22.

    Article  PubMed  PubMed Central  Google Scholar 

  47. Suzuki H, Aoki K, Chiba K, Sato Y, Shiozawa Y, Shiraishi Y, Shimamura T, Niida A, Motomura K, Ohka F, et al. Mutational landscape and clonal architecture in grade II and III gliomas. Nat Genet. 2015;47:458–68.

    Article  CAS  PubMed  Google Scholar 

  48. Ichimura K, Pearson DM, Kocialkowski S, Backlund LM, Chan R, Jones DT, Collins VP. IDH1 mutations are present in the majority of common adult gliomas but rare in primary glioblastomas. Neuro Oncol. 2009;11:341–7.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  49. Elez E, Ros J, Fernandez J, Villacampa G, Moreno-Cardenas AB, Arenillas C, Bernatowicz K, Comas R, Li S, Kodack DP, et al. RNF43 mutations predict response to anti-BRAF/EGFR combinatory therapies in BRAF(V600E) metastatic colorectal cancer. Nat Med. 2022;28:2162–70.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  50. Faltus C, Lahnsteiner A, Barrdahl M, Assenov Y, Husing A, Bogatyrova O, Laplana M, Johnson T, Muley T, Meister M, et al: Identification of NHLRC1 as a Novel AKT Activator from a Lung Cancer Epigenome-Wide Association Study (EWAS). Int J Mol Sci 2022;23.

  51. Lan T, Ge Q, Zheng K, Huang L, Yan Y, Zheng L, Lu Y, Zheng D. FAT1 Upregulates in oral squamous cell carcinoma and promotes cell proliferation via cell cycle and DNA repair. Front Oncol. 2022;12:870055.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  52. Bakr A, Oing C, Kocher S, Borgmann K, Dornreiter I, Petersen C, Dikomey E, Mansour WY. Involvement of ATM in homologous recombination after end resection and RAD51 nucleofilament formation. Nucleic Acids Res. 2015;43:3154–66.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  53. Morrison C, Sonoda E, Takao N, Shinohara A, Yamamoto K, Takeda S. The controlling role of ATM in homologous recombinational repair of DNA damage. EMBO J. 2000;19:463–71.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  54. Britton S, Chanut P, Delteil C, Barboule N, Frit P, Calsou P. ATM antagonizes NHEJ proteins assembly and DNA-ends synapsis at single-ended DNA double strand breaks. Nucleic Acids Res. 2020;48:9710–23.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  55. Kaminskyy VO, Haag P, Novak M, Vegvari A, Arapi V, Lewensohn R, Viktorsson K: EPHA2 interacts with DNA-PK(cs) in cell nucleus and controls ionizing radiation responses in non-small cell lung cancer cells. Cancers (Basel) 2021;13.

  56. Jang ER, Choi JD, Jeong G, Lee JS. Phosphorylation of p300 by ATM controls the stability of NBS1. Biochem Biophys Res Commun. 2010;397:637–43.

    Article  CAS  PubMed  Google Scholar 

  57. Zapata L, Pich O, Serrano L, Kondrashov FA, Ossowski S, Schaefer MH. Negative selection in tumor genome evolution acts on essential cellular functions and the immunopeptidome. Genome Biol. 2018;19:67.

    Article  PubMed  PubMed Central  Google Scholar 

  58. Van den Eynden J, Larsson E. Mutational signatures are critical for proper estimation of purifying selection pressures in cancer somatic mutation data when using the dN/dS metric. Front Genet. 2017;8:74.

    Article  PubMed  PubMed Central  Google Scholar 

  59. Zhou Z, Zou Y, Liu G, Zhou J, Wu J, Zhao S, Su Z, Gu X. Mutation-profile-based methods for understanding selection forces in cancer somatic mutations: a comparative analysis. Oncotarget. 2017;8:58835–46.

    Article  PubMed  PubMed Central  Google Scholar 

  60. Frankell AM, Dietzen M, Al Bakir M, Lim EL, Karasaki T, Ward S, Veeriah S, Colliver E, Huebner A, Bunkum A, et al. The evolution of lung cancer and impact of subclonal selection in TRACERx. Nature. 2023;616:525–33.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  61. Cook JH, Melloni GEM, Gulhan DC, Park PJ, Haigis KM. The origins and genetic interactions of KRAS mutations are allele- and tissue-specific. Nat Commun. 1808;2021:12.

    Google Scholar 

  62. Roberts PJ, Der CJ. Targeting the Raf-MEK-ERK mitogen-activated protein kinase cascade for the treatment of cancer. Oncogene. 2007;26:3291–310.

    Article  CAS  PubMed  Google Scholar 

  63. Samatar AA, Poulikakos PI. Targeting RAS-ERK signalling in cancer: promises and challenges. Nat Rev Drug Discov. 2014;13:928–42.

    Article  CAS  PubMed  Google Scholar 

  64. Van Allen EM, Wagle N, Sucker A, Treacy DJ, Johannessen CM, Goetz EM, Place CS, Taylor-Weiner A, Whittaker S, Kryukov GV, et al. The genetic landscape of clinical resistance to RAF inhibition in metastatic melanoma. Cancer Discov. 2014;4:94–109.

    Article  PubMed  Google Scholar 

  65. Westover D, Zugazagoitia J, Cho BC, Lovly CM, Paz-Ares L. Mechanisms of acquired resistance to first- and second-generation EGFR tyrosine kinase inhibitors. Ann Oncol. 2018;29:i10–9.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  66. Tan Z, Gao L, Wang Y, Yin H, Xi Y, Wu X, Shao Y, Qiu W, Du P, Shen W, et al. PRSS contributes to cetuximab resistance in colorectal cancer. Sci Adv. 2020;6:eaax5576.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  67. Shen CH, Kim SH, Trousil S, Frederick DT, Piris A, Yuan P, Cai L, Gu L, Li M, Lee JH, et al. Loss of cohesin complex components STAG2 or STAG3 confers resistance to BRAF inhibition in melanoma. Nat Med. 2016;22:1056–61.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  68. Consortium ITP-CAoWG. Pan-cancer analysis of whole genomes. Nature. 2020;578:82–93.

    Article  Google Scholar 

  69. Turajlic S, Sottoriva A, Graham T, Swanton C. Resolving genetic heterogeneity in cancer. Nat Rev Genet. 2019;20:404–16.

    Article  CAS  PubMed  Google Scholar 

  70. Nik-Zainal S, Alexandrov LB, Wedge DC, Van Loo P, Greenman CD, Raine K, Jones D, Hinton J, Marshall J, Stebbings LA, et al. Mutational processes molding the genomes of 21 breast cancers. Cell. 2012;149:979–93.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  71. Lee SY, Wang H, Cho HJ, Xi R, Kim TM. The shaping of cancer genomes with the regional impact of mutation processes. Exp Mol Med. 2022;54:1049–60.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  72. Chang MT, Asthana S, Gao SP, Lee BH, Chapman JS, Kandoth C, Gao J, Socci ND, Solit DB, Olshen AB, et al. Identifying recurrent mutations in cancer reveals widespread lineage diversity and mutational specificity. Nat Biotechnol. 2016;34:155–63.

    Article  CAS  PubMed  Google Scholar 

  73. Muinos F, Martinez-Jimenez F, Pich O, Gonzalez-Perez A, Lopez-Bigas N. In silico saturation mutagenesis of cancer genes. Nature. 2021;596:428–32.

    Article  CAS  PubMed  Google Scholar 

  74. The Cancer Genome Atlas consortium. PanCancer mutation profiles (mc3.v0.2.8.PUBLIC.maf.gz). https://gdc.cancer.gov/about-data/publications/pancanatlas.

  75. MSKCC-IMPACT clinical sequence cohort. Mutational profiles. https://www.cbioportal.org/study/summary?id=msk_impact_2017.

  76. Cancer Gene Census, COSMIC database, v96 2022 May. https://cancer.sanger.ac.uk/census.

  77. dNdScv R packages homepage. https://github.com/im3sanger/dNdScv.

  78. DepMap database, cell essentiality dataset. https://depmap.org/portal/.

  79. GDSC database. https://www.cancerrxgene.org/.

  80. CCLE database. https://sites.broadinstitute.org/ccle/.

  81. Github repository. https://github.com/TkimLab/cdNS.

Download references

Acknowledgements

We appreciate the support of this research by the Basic Medical Science Facilitation Program through the Catholic Medical Center of Catholic University of Korea funded by the Catholic Education Foundation.

Funding

This work was supported by the National Research Foundation of Korea (NRF) funded by the Ministry of Science and ICT (2019R1A5A2027588 and 2019M3E5D3073104 to T.-M.K.).

Author information

Authors and Affiliations

Authors

Contributions

D.H. and S.K. carried out bioinformatics analyses and troubleshooting. S.L., Y.M., S.K., J.Y., H.J, H.C., and J.J. designed and performed statistical analyses. S.L., R.X., and T.K. built schematics of the project and analyzed data. B.S. and J.K. analyzed data for validation. D.H., S.K., and T.K. wrote the paper. T.K. conceived and supervised the project. All authors read and approved the final manuscript.

Corresponding author

Correspondence to Tae-Min Kim.

Ethics declarations

Ethics approval and consent to participate

Not applicable.

Consent for publication

Not applicable.

Competing interests

The authors declare that they have no competing interests.

Additional information

Publisher’s Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Dong-Jin Han, Sunmin Kim wish it to be known that, in their opinion, the first two authors should be regarded as joint First Authors.

Supplementary Information

13073_2024_1376_MOESM1_ESM.xlsx

Additional file 1: Table S1. Gene used for ED analyses. Table S2. PanCancer gene pairs with ED. Table S3. Tumor type-specific gene pairs with ED. Table S4. ED and genetic dependencies. Table S5. Effect size of targeted agents with ED of gene pairs

13073_2024_1376_MOESM2_ESM.pdf

Additional file 2: Supplementary Fig. S1. The subsampling analysis of mutation contexts. Supplementary Fig. S2. ED of known gene pairs. Supplementary Fig. S3. ED in panel sequencing data. Supplementary Fig. S4. ED of mutations conserved across Ras mutation contexts. Supplementary Fig. S5. VAF and TMB of SYN and ANT gene pairs in tumor subtypes. Supplementary Fig. S6. Genetic dependency and tumor type-specific gene pairs. Supplementary Fig. S7. Cell viability of synergistic and antagonistic mutation pairs

Rights and permissions

Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if changes were made. The images or other third party material in this article are included in the article's Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article's Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by/4.0/. The Creative Commons Public Domain Dedication waiver (http://creativecommons.org/publicdomain/zero/1.0/) applies to the data made available in this article, unless otherwise stated in a credit line to the data.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Han, DJ., Kim, S., Lee, SY. et al. Evolutionary dependency of cancer mutations in gene pairs inferred by nonsynonymous-synonymous mutation ratios. Genome Med 16, 103 (2024). https://doi.org/10.1186/s13073-024-01376-7

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/s13073-024-01376-7

Keywords