Skip to main content

ORIGINAL RESEARCH article

Front. Genet., 28 April 2020
Sec. Computational Genomics

Prioritizing Cancer Genes Based on an Improved Random Walk Method

  • 1Key Lab of Intelligent Computing and Signal Processing of Ministry of Education, College of Computer Science and Technology, Anhui University, Hefei, China
  • 2Division of Biomedical Engineering, University of Saskatchewan, Saskatoon, SK, Canada
  • 3Department of Computer Sciences, University of Saskatchewan, Saskatoon, SK, Canada
  • 4Department of Mechanical Engineering, University of Saskatchewan, Saskatoon, SK, Canada
  • 5Institutes of Physical Science and Information Technology, Anhui University, Hefei, China
  • 6College of Computer and Information Engineering, Fuyang Normal University, Fuyang, China

Identifying driver genes that contribute to cancer progression from numerous passenger genes, although a central goal, is a major challenge. The protein–protein interaction network provides convenient and reasonable assistance for driver gene discovery. Random walk–based methods have been widely used to prioritize nodes in social or biological networks. However, most studies select the next arriving node uniformly from the random walker's neighbors. Few consider transiting preference according to the degree of random walker's neighbors. In this study, based on the random walk method, we propose a novel approach named Driver_IRW (Driver genes discovery with Improved Random Walk method), to prioritize cancer genes in cancer-related network. The key idea of Driver_IRW is to assign different transition probabilities for different edges of a constructed cancer-related network in accordance with the degree of the nodes' neighbors. Furthermore, the global centrality (here is betweenness centrality) and Katz feedback centrality are incorporated into the framework to evaluate the probability to walk to the seed nodes. Experimental results on four cancer types indicate that Driver_IRW performs more efficiently than some previously published methods for uncovering known cancer-related genes. In conclusion, our method can aid in prioritizing cancer-related genes and complement traditional frequency and network-based methods.

Introduction

As one of the most complex and threatening diseases, cancer has attracted the attention of many research groups and large-scale programs [such as The Cancer Genome Atlas (TCGA) (Network, 2008) and the International Cancer Genome Consortium (Bobrow and Zhao, 2010)] to explore the molecular mechanisms and pathogenesis. With the rapid advances of technology, huge volumes of cancer genomics data have been generated containing many different types of genetic aberrations, such as single-nucleotide variants (SNVs), copy number variations (CNVs), and small and large insertions and deletions (Indels) (Zhang and Zhang, 2017; Dimitrakopoulos et al., 2018). It has been demonstrated that cancer is related to gene mutations, but only a few genes exist that confer selective growth advantage to cancer progression, known as driver genes. The remaining mutations are called passenger genes (Greenman et al., 2007; Stratton et al., 2009; Vogelstein et al., 2013; Iranzo et al., 2018). However, identifying and distinguishing driver genes from myriad passengers are a fundamental question and an intractable challenge (Haber and Settleman, 2007; Stratton et al., 2009; Vogelstein et al., 2013) and are crucial to gain insights into biological processes (Zhang et al., 2018).

Efforts have been made to address this challenge using a variety of novel methods. The most traditional approaches are based on genetic aberration frequencies among population cohorts of patients with cancer (Gui et al., 2011; Dees et al., 2012; Lawrence et al., 2013), which can detect some major driver genes with significantly higher mutation rates than background mutation rates (BMRs). However, the estimations of BMR significantly affect the identification of driver mutations. Additionally, it has been found that the BMR is dependent on the genomic context of the nucleotide, the type of mutation transcription rates, and replication time (Raphael et al., 2014). Therefore, it is difficult to estimate BMR accurately. Moreover, genes altered in only a few individuals may be relatively important in contributing to cancer progression (Stratton et al., 2009; Raphael et al., 2014; Hristov and Singh, 2017). Evidently, these frequency-based methods cannot reveal rarely mutated driver genes. Consequently, some promising methods considering somatic mutations in the context of pathways are proposed, because cellular signaling and regulatory pathways are usually affected by driver mutations (Network, 2008; Vandin et al., 2012; Jones et al., 2016). Additionally, most pathway-based methods are primarily based on the mutual exclusivity of mutations (Zhang et al., 2014; Wu et al., 2015; Zhang and Zhang, 2018). They are focused on analyzing somatic mutation rather than integrating different omics data, such as transcriptome and interactome. Given that proteins tend to be proximal if they take part in the same pathways (Hristov and Singh, 2017), in recent years, many novel methods based on networks have been successfully applied to cancer driver gene identification by integrating different omics data (Bashashati et al., 2012; Hou and Ma, 2014; Amgalan and Lee, 2015; Bertrand et al., 2015; Dimitrakopoulos et al., 2018; Song et al., 2019). However, some only map genes of different omics data into networks without collecting more information regarding network topology into account. Furthermore, some network diffusion approaches, such as DawnRank (Hou and Ma, 2014), propagate expression information through a protein interaction network by selecting the next arriving node from its neighbors uniformly. Moreover, Gentili et al. (2019) have proposed a BRW (biological random walk) method to leverage biological information in network propagation for gene prioritization. In addition, there are also some advances of random walk in different research field. For example, Chen et al. (2016) have proposed an improved random walk with restart method for lncRNA-disease association prediction (IRWRLDA). These two methods improve the initial probabilities of restart term by setting uniform probability of disease-associated seed nodes and considering lncRNA expression similarity and disease semantic similarity separately. However, in real-world scenarios, the random walker is more likely to have tendentiousness and preference for selecting the neighbors with a greater degree rather than uniformly (Liu et al., 2017). The aforementioned random walk–based methods rarely consider it. Although some methods have realized the importance of seed genes, they did not consider the topological characteristics. Therefore, it is beneficial for a novel method to take more graph topological characteristics and propagating tendency into consideration to identify cancer driver genes.

In this study, to mitigate these methodological limitations and improve the accuracy of driver gene identification, we proposed a novel approach based on the random walk method, named Driver_IRW (Driver genes discovery with Improved Random Walk method), for driver genes discovery by integrating transcriptomic data and interaction network. The assumption of our method is that genes in the interaction network with a higher degree have a higher transition probability from their upstream neighbors. First, we constructed different networks for different types of cancer by selecting those edges that exist in both the known PPI network and differential coexpression network (Guo et al., 2019), in which the known information of the PPI network used is a directed network from DanwRank (Hou and Ma, 2014). The tumor and normal expression data were used to construct the differential coexpression network for each type of cancer. Then, the degree, betweenness, and Katz centralities were obtained based on the constructed network. Third, based on the assumption, we adopt the strategy that the information in network is diffused in accordance with its neighbors' out-degree rather than uniformly (Liu et al., 2017). Moreover, the betweenness and Katz centrality of different seed genes of corresponding cancers were merged as random jumping probabilities to these nodes, in which the different seed nodes were extracted from CGC (Sondka et al., 2018) and DisGeNet databases (Piñero et al., 2016) for different cancers. The random walk scores were calculated by the improved random walk method. Finally, only mutated genes were retained. To evaluate the performance of the proposed method, data of four cancer types from TCGA were used, and the results indicate that it performs well. Moreover, the benchmark analysis showed that the proposed method is useful.

Materials and Methods

Datasets

In this work, four different types of cancer, breast cancer (BRCA) with 1097 samples, head and neck squamous cell carcinoma (HNSC) with 522 samples, kidney renal cell cancer (KIRC) with 534 samples, and thyroid cancer (THCA) with 513 samples, from TCGA were studied. The datasets used consisted of mutation and expression data from tumor and normal samples for every cancer. The mutation data were integrated by SNVs and CNVs. It was regarded as a mutated gene if there was an SNV or CNV present, in which the CNV data are downloaded from UCSC data portal (http://xena.ucsc.edu/) (Rosenbloom et al., 2015), which have transformed the data from TCGA using Gistic2 method. And we retained only those genes with +2 and −2 values, which are the high-level amplification and homozygous deletion. The RNAseq expression data were real values denoting the normalized abundance of each gene in each sample. To obtain the differential coexpression network, the expression data of normal and tumor samples were required. In addition, the seed genes of different cancers were derived from CGC (release v85, downloaded on May 8, 2018) (Sondka et al., 2018) and DisGeNet (release v6, downloaded in February, 2019) databases (Piñero et al., 2016). The CGC database encompasses 719 expert-curated descriptions of the genes driving human cancer (Sondka et al., 2018), and DisGeNet is one of the largest available collections of human disease-involved genes and variants (Piñero et al., 2016). Additionally, the reference network used in this study was downloaded from DanwRank (Hou and Ma, 2014), which integrates various sources, including MEMo (Ciriello et al., 2012), NCI-Nature Curated PID (Schaefer et al., 2008), Rectome (Croft et al., 2010), and KEGG (Kanehisa et al., 2011). It can be viewed as a directed graph.

The Construction of the Cancer-Related Network

To retrieve more specific peculiarity of different types of cancer, we constructed different networks for different types of cancer by integrating the known PPI network and differential coexpression network (Guo et al., 2019).

First, Pearson correlation coefficients with p-values of tumor and normal expression data were calculated separately as coexpression networks for different cancers. Then, the differential coexpression network was constructed as per the following two steps: (1) only the edges with p < 0.05 were selected and assigned to 1 as significantly correlated gene pairs; (2) the differential edges that were significantly correlated only in the tumor or normal coexpression networks, i.e., the edges, were statistically significant (p < 0.05) in tumor (normal) data but not (p-value > 0.05) in normal (tumor) data, were screened out. The consistent edges, i.e., edges that were statistically significant or not in both tumor and normal coexpression networks, were removed. Finally, the reference network downloaded from DawnRank (Hou and Ma, 2014) was integrated with differential coexpression network by selecting the common nodes and edges (Guo et al., 2019). The framework of Driver_IRW is shown in Figure 1. This reconstructed network, i.e., adjacency matrix, is a 0–1 matrix represented as A with A(i, j) = 1 if node i links to j in the constructed network, otherwise, A(i, j) = 0.

FIGURE 1
www.frontiersin.org

Figure 1. The framework of Driver_IRW for identifying cancer driver genes. (A) The procedure of data processing, including the construction of the network and the selection of the seed genes for different cancers. (B) The improved random walk, including the calculation of the improved transition probabilities and the calculation of the global centralities for seed genes (blue nodes). The mutation data are used as posterior information to filter the mutated candidate driver genes.

The Selection of Cancer-Related Seed Genes

In previous studies, seed genes have been widely used as prior information for disease gene discovery (Köhler et al., 2008; Moreau and Tranchevent, 2012). To take this prior information into account in the Driver_IRW framework, and in light of the previous conception, known cancer-related genes for the corresponding types of cancer were used as seed genes in this study. Different cancer-related genes were extracted for corresponding types of cancers from CGC (release v85, May 8th, 2018) (Sondka et al., 2018) and DisGeNet (Piñero et al., 2016) databases.

The Calculation of Centralities

By virtue of the adjacency matrix, the topological centralities, here are degree centrality, betweenness centrality, and Katz feedback centrality (termed as DC, BC, and KC, respectively), were calculated. The degree centrality was used to obtain the weighted transition matrix. Additionally, the betweenness and Katz feedback centralities were used to evaluate the random jumping probability to seed genes.

The degree centrality of a vertex is the number of edges incident to the vertex in a graph. That is,

DCi=j=1nAij    (1)

where n is the total number of genes in the network; Aij is the adjacency matrix of the network.

The betweenness centrality can be interpreted as a vertex being more central if it is needed to transport more information of others in the network. This is calculated as follows:

BCi=2n2-3n+2sitnst(i)gst    (2)

where gst indicates the total number of shortest paths from node s to node t, and nst(i) indicates the number of these paths that pass through vertex i. The betweenness centrality was obtained using the “igraph” R package, which provided handy tools to create, manipulate, and visualize networks, and calculate various structural properties (Csardi, 2006).

The Katz centrality, one of the feedback centralities, is calculated based on the impact of a vertex on others. This is defined as follows:

KC(i)=k=0j=1nαk(Ak)ji    (3)

where A is the adjacency matrix; (Ak)ji is the number of paths from j to i with length k; and α is a damping factor, which restricts that the longer the path between i and j, the smaller the impact of i on j should be. It has been proved that to guarantee convergence the α must be restricted as follows:

λ1<1αk=1 αkAk   converge    (4)

where λ1 is the largest eigenvalue of A. The closed form expression is Equation (5) when it converges:

KC=k=1αk(AT)k1n=((I-αAT)-1)1n    (5)

Then, the prior information used in the diffusing procedure as random jumping probabilities to seeds is represented by the normalized mean value of normalized betweenness centrality and Katz centrality of seeds.

The Algorithm of Improved RW

In light of the above assumption that the information in the network is diffused in accordance with its neighbors' out-degree instead of uniformly, the transition matrix whose values represent transition probability from the vertex i to any vertex j of the directed graph was defined as follows (Liu et al., 2017):

pij={αDCj+AijvN+(i)DCv++(1-α)1n,if vN+(i)DCv+0;     1n,otherwise    (6)

where DCv+ and DCj+ are the out-degree of vertex v and j, respectively. N+(i) denotes all the neighbors interacting with vertex i in the network. In addition, n is the total number of genes in the network. This means that if there are neighbors of one node, the transition probabilities from the node to the neighbors are proportional to the neighbors' out-degree; otherwise, the transition probability is uniform according to the total genes. A parameter of α (empirically, here α = 0.85) is presented. This is used to avoid neglecting the nodes whose out-degree of the neighbors of node i is zero.

Next, we defined the score of each gene iteratively according to the improved random walk approach:

ri(t+1)=d·j=1npji·rj(t)+(1-d)prior_ps    (7)

It can be presented in the matrix form:

r(t+1)=dPT×r(t)+(1-d)×prior_p    (8)

where r(t) and prior_p are n × 1 vectors representing the gene score in the t-th iteration of each node and prior information of seed nodes separately with the sum of the values equal to 1 (Köhler et al., 2008). The initial gene scores r(0) are the normalized mean of all tumor expression. Moreover, PT is the transposition of the transition matrix P obtained by Equation (6). Here, d is set to 0.85 according to the initial PageRank algorithm and is the damping factor that corresponds to a random walker periodically jumping to a random node (Page et al., 1999), which is used as the seed gene of different cancers. To retrieve more global information on the network, betweenness and Katz feedback centralities were used to represent the random jumping probabilities of a random walker jumping to the seeds in each iteration.

The final random walk scores of all nodes converged to a stationary distribution when there was no longer a significant update in the scores (Hou and Ma, 2014). When the difference (Equation 9) in scores between the (t+1)-th and previous t-th iteration was smaller than ε, the iteration stopped. Here, the threshold was set as ε = 1e − 8. Additionally, the iteration stopped after the maximum number of iterations, which was set to 1,000, when no solution was obtained. In practice, the improved method always converges to the stationary status.

diff=i=1n(ri(t+1)-ri(t))2    (9)

To retrieve the impact of the mutated genes in the population, only the mutated genes were retained in the final results.

Results

Performance Evaluation for Known Cancer-Related Genes

For evaluating and comparing the performance of Driver_IRW in predicting known cancer genes, two publicly available databases-20/20 rule (Vogelstein et al., 2013) and IntOGen (downloaded on May, 2019) (Gonzalez-Perez et al., 2013), datasets were utilized as approximate benchmarks. The 20/20 rule dataset contained 138 well-studied oncogenes and tumor suppressor genes, which were used to assess the ability of our method in identifying known cancer drivers. The IntOGen database lists previously detected drivers of different cancers, which were used to evaluate the performance of driver discovery on specific types of cancer. Different driver genes were extracted for different types of cancer as benchmarks for performance comparison. Therefore, we compared Driver_IRW with previous state-of-the-art methods, such as DawnRank (Hou and Ma, 2014), DriverNet (Bashashati et al., 2012), MUFFINN (Cho et al., 2016), and naive mutation frequency-based method to evaluate the performance of predicting known cancer-related genes. In particular, the MUFFINN method provided two strategies; DNmax (direct neighbor max), which counts mutations in the most frequently mutated neighbors, and DNsum (direct neighbor sum), which counts mutations in all direct neighbors using the networks HumanNet (Lee et al., 2011) and STRING (Szklarczyk et al., 2015; Cho et al., 2016). For comparison, the best performance between the two networks of each strategy of the top N genes was selected. In addition, other methods were executed in their default settings. Moreover, the precision-recall curves of the top N genes are illustrated in Figure 2. The precision and recall matrices are defined as follows:

precision=(#genes in benchmark)(# genes found in Driver_IRW)#genes found in Driver_IRW    (10)
recall=(#genes in benchmark)(# genes found in Driver_IRW)#genes found in benchmark    (11)

where (# genes in benchmark) represents the number of genes in benchmarks (20/20 rule and IntOGen datasets) dataset, and (# genes found in Drive rIRW) represents the number of top N genes prioritized by Driver-IRW.

FIGURE 2
www.frontiersin.org

Figure 2. Assessment of predictive power of Driver-IRW for known cancer related genes compared with different methods. The precision-recall curves of the top 100 results of Driver-IRW, DawnRank, DriverNet, MUFFINN (contained two results of MUFFINN_DNmax and MUFFINN_DNsum), and naïve frequency-based methods by 20/20 rule gene list for four datasets (A) BRCA, (B) HNSC, (C) KIRC, and (D) THCA, and by IntOGen database for datasets (E) BRCA, (F) HNSC, (G) KIRC, and (H) THCA.

In practice, researchers may only be interested in the top-ranked candidate genes for follow-up experimental validation. Hence, only the top 100 candidate driver genes (detailed lists are in the Supplementary File) were selected to assess their performance. In general, Driver_IRW outperforms most other methods in four datasets regardless of whether assessing with the 20/20 rule or IntOGen metrics, which indicates that Driver_IRW can identify more known cancer drivers than other methods. Overall, the results show that our proposed method performs well in identifying known cancer-related genes.

Analysis of the Improvement of the Transition Matrix

To verify that there are improvements after incorporating the tendency of nodes in random walk, the strategy that assigns all neighbors of node i with uniform transition probabilities (named RW_UniTr) was adopted. Here, the random jumping probability remained the same as that of Driver_IRW. The benchmark data are the 20/20 rule, and the results are shown in Figure 3. The results indicated that the transition tendency in accordance with the neighbors' out-degree and network topological information are important for improving the performance of the method.

FIGURE 3
www.frontiersin.org

Figure 3. The analysis of the improvement of transition matrix. Driver_IRW is our proposed method; RW_UniTr means random walk method with uniform transition probability. The precision-recall curves of top 100 results of Driver_IRW and RW_UniTr by 20/20 rule gene list for four datasets (A) BRCA, (B) HNSC, (C) KIRC, and (D) THCA.

The Analysis of Seed Genes

Based on the successful application of seed genes on disease-related discovery in previous studies (Köhler et al., 2008), known cancer driver genes were used in our method. To evaluate the performance after considering seed genes, the seed genes were removed from our Driver_IRW method (termed as IRW_withoutS). Without the seed genes, the random jumping probabilities of all nodes were set to equal values (1/n, n is the total number of genes), which was equivalent to letting the random walker jump to all nodes with equal probability. Analogous to the analysis of transition matrix, the benchmark here is the 20/20 rule dataset. The results are shown in Figure 4. These results indicate that there is mild improvement after taking the seed genes into consideration.

FIGURE 4
www.frontiersin.org

Figure 4. The analysis of seed genes. Driver_IRW is our proposed method with seed genes; IRW_withoutS means removing the seed genes from Driver_IRW method and setting them with equal probabilities. The precision-recall curves of top 100 results of Driver_IRW and IRW_withoutS by 20/20 rule gene list for four datasets (A) BRCA, (B) HNSC, (C) KIRC, and (D) THCA.

In addition, in order to investigate the impact of the seed nodes and demonstrating whether the results are sensitive to the selection of seed node, we have randomly deleted 10, 30, and 50% nodes for 10 times, respectively, from all seed nodes we used in this article. Next, we applied our method to every subset. Then the accuracy and recall are calculated in accordance with the 20/20 rule dataset for 10 times. Finally, the mean value of accuracy and recall of 10 times are calculated and used to compared with the full seed nodes (Figure 5).

FIGURE 5
www.frontiersin.org

Figure 5. The assessment of different seeds for four types of cancer. Here, the full seeds mean the results of all seed nodes we integrated from CGC and DisGeNet database; 90% seeds, 70% seeds, and 50% seeds mean that we randomly deleted 10, 30, and 50% seed nodes, respectively.

It is obviously that the performances of full seeds and 90% seeds are similar. And with the rise of the percentage we deleted seeds from full seeds, the performance seems becoming a little worse especially in KIRC and THCA dataset. However, in general, the performance on these subsets does not change significantly. It means that the results are not too sensitive to the selection of seed nodes.

Analysis of the Top 10 Candidate Drivers

The overall performance of identifying cancer drivers of the top 100 genes is shown in Figure 2. Here, the top 10 candidate genes were listed to illustrate their importance. First, their statuses were explored regarding whether they are known cancer drivers or candidate drivers in the NCG 6.0 database (Repana et al., 2019), which is a manually curated repository that contained 2,372 genes whose somatic modifications have been reported as known or predicted cancer driver roles (Repana et al., 2019). Then, the mutation frequencies of these genes were calculated to investigate the sensibility to the mutation frequency of our method. The results are shown in Table 1.

TABLE 1
www.frontiersin.org

Table 1. The top 10 candidate drivers of different cancers.

As shown in Table 1, almost all of the top 10 genes are cancer-related according to NCG 6.0. Genes not reported by NCG 6.0 may also have potential effects on cancer. For example, it has been demonstrated that PAK1 is increased in breast cancer and plays a pivotal role in promoting tumor growth and drug resistance (Kumar et al., 2006; Dou et al., 2016). In addition, PAK1 has been reported by some studies to play a key role in the initiation and progression of KIRC (O'Sullivan et al., 2007).

Furthermore, besides the frequently mutated and important candidate drivers, such as TP53 in BRCA and HNSC and NRAS and HRAS in THCA, rare mutated (usually defined as mutated frequency <2% of samples Hou and Ma, 2014) candidate drivers could also be identified through our method (Table 1). Interestingly, TP53 and CTNNB1 rank in the top 10 in four cancer types. It is widely known that the tumor suppressor gene TP53 is frequently mutated in most human cancers and has an important role in the cellular stress response (Petitjean et al., 2007; Hidalgo, 2010). Additionally, CTNNB1 is reported as a potential biomarker using the KIRC corresponding network (Isik and Ercan, 2017). Moreover, mutations in CTNNB1 are related to several human malignancies, such as colorectal cancer (Klaus and Birchmeier, 2008), lung cancer (Schou et al., 2001), HNSC (Jerhammar et al., 2010), and KIRC (Hirata et al., 2012), although it is a rarely mutated gene.

Discussion

The identification of cancer driver genes is a valuable task for cancer genomics analysis (Guo et al., 2018). In this study, we propose Driver_IRW, an improved random walk–based framework, to prioritize cancer genes. Since the transition of a node moving to others is more likely to have tendentiousness in reality, a strategy different from the traditional random walk method was used whereby nodes transit to others based on the out-degree of their neighbors rather than move to their neighbors uniformly according to the degree of themselves. This not only spreads information rapidly but also avoids trapping by dangling nodes. Additionally, the application of the method without seed genes exhibits that the seed genes play a role in driver gene identification. However, the improvement is not prominent. On the one hand, the quality of the seeds may affect the performance; hence, it is necessary to collect high-quality seeds. On the other hand, this might mean that our method is robust to the prior information. In a nutshell, the main differences of our proposed method Driver_IRW with other methods are mainly in two aspects. The first one is the computation of the transition probability compared with some methods such as DawnRank, BRW, and IRWRLDA. The second one is the computation of the prior information compared with DPRank method. DPRank method considers the tendentiousness of the neighbors' degree in the network; however, it neglects the impact of known seeds in the real biology network. The experimental results on four different cancer datasets (Figure 2) indicate that our method is more effective when evaluating known driver discovery than some previous methods. Moreover, the results in Table 1 also show that Driver_IRW can identify not only the frequently mutated genes but also rarely mutated drivers. Comparison of Driver_IRW with the method that uses traditional transition probability indicates that the improvement on transition matrix indeed improves the performance. It should be noted that Driver_IRW was only applied on four TCGA datasets as case studies in this study; it could be extended to other datasets if expression data, mutation data, and corresponding seeds genes are available.

Our method may be a complement of the traditional frequency-based methods and some network-based methods. However, there are also some limitations for this method. In this study, the mutation data are only regarded as the posterior information to filter the candidate driver genes. Generally, cancer evolves to accumulate additional alterations (Nussinov et al., 2019), which might infer more important information to take the mutation and time-dependent alteration data into consideration. Besides genetic aberrations, other events, such as miRNA differential expression and epigenetic changes, can also contribute to the progression of cancer. The expression of mRNAs can be controlled by upregulated miRNAs (Dimitrakopoulos et al., 2018). This may increase the power of our method if more omics information are integrated. In addition, the cancer data used were from all patients of one type. However, there are different subtypes for multiple cancers. For example, BRCA is typically classified as luminal A, luminal B, triple-negative/basal-like, HER2-enriched, and normal-like subtypes, and HNSC contains HPV+ (human papillomavirus positive) and HPV– (human papillomavirus negative) subtypes (Vokes et al., 2015). Moreover, the characteristics, molecular profiles, or specific mutations are usually distinguished among different subtypes. Therefore, future work in dividing the cancer types into different subtypes to research the driver gene separately will be valuable.

In conclusion, Driver_IRW is easy to use for prioritizing cancer genes with the improved random walk–based method. We expect that our method will provide a valuable resource and can be amended in the future.

Data Availability Statement

Publicly available datasets were analyzed in this study. This data can be found here: breast cancer (BRCA) with 1097 samples, head and neck squamous cell carcinoma (HNSC) with 522 samples, kidney renal cell cancer (KIRC) with 534 samples and thyroid cancer (THCA) with 513 samples from TCGA (https://portal.gdc.cancer.gov/).

Author Contributions

P-JW conceived the algorithm, designed the method and drafted the manuscript, analyzed the data, and carried out the experiments. F-XW and C-HZ refined the idea, polished the English expression, and revised the paper. JX and YS participated in the design and revision of the research. JW participated the discussion and the coordination of the study. All authors read and approved the final manuscript.

Funding

This work was supported by the National Natural Science Foundation of China (grant numbers U19A2064, 61873001, 61872220, and 61672037), the Key Project of Anhui Provincial Education De-partment (No. KJ2017ZD01), the Key Project of Academic Funding for Top-notch Talents in Univer-sity of Anhui (No. gxbjZD2016007) and the Natural Science Foundation of Anhui Province (No. 1808085QF209).

Conflict of Interest

The authors declare that the research was conducted in the absence of any commercial or financial relationships that could be construed as a potential conflict of interest.

Acknowledgments

The authors acknowledge the High-performance Computing Platform of Anhui University for providing computing resources.

Supplementary Material

The Supplementary Material for this article can be found online at: https://www.frontiersin.org/articles/10.3389/fgene.2020.00377/full#supplementary-material

Abbreviations

TCGA, The Cancer Genome Atlas; ICGC, International Cancer Genome Consortium; SNVs, Single Nucleotide Variants; CNVs, Copy Number Variations; Indels, Insertions and Deletions; BMR, Background Mutation Rates; BRCA, Breast Cancer; HNSC, Head and Neck Squamous Cell Carcinoma; KIRC, Clear Cell Kidney Carcinoma; THCA, Thyroid Carcinoma; PCC, Pearson Correlation Coefficients; DC, Degree Centrality; BC, Betweenness Centrality; KC, Katz Centrality.

References

Amgalan, B., and Lee, H. (2015). DEOD: uncovering dominant effects of cancer-driver genes based on a partial covariance selection method. Bioinformatics 31, 2452–2460. doi: 10.1093/bioinformatics/btv175

PubMed Abstract | CrossRef Full Text | Google Scholar

Bashashati, A., Haffari, G., Ding, J., Ha, G., Lui, K., Rosner, J., et al. (2012). DriverNet: uncovering the impact of somatic driver mutations on transcriptional networks in cancer. Genome Biol. 13:R124. doi: 10.1186/gb-2012-13-12-r124

PubMed Abstract | CrossRef Full Text | Google Scholar

Bertrand, D., Chng, K. R., Sherbaf, F. G., Kiesel, A., Chia, B. K. H., Sia, Y. Y., et al. (2015). Patient-specific driver gene prediction and risk assessment through integrated network analysis of cancer omics profiles. Nucleic Acids Res. 43:44. doi: 10.1093/nar/gku1393

PubMed Abstract | CrossRef Full Text | Google Scholar

Bobrow, M., and Zhao, S. (2010). International network of cancer genome projects. Nature 464, 993–998. doi: 10.1038/nature08987

PubMed Abstract | CrossRef Full Text | Google Scholar

Chen, X., You, Z., Yan, G., and Gong, D. (2016). IRWRLDA: improved random walk with restart for lncRNA-disease association prediction. Oncotarget 7, 57919–57931. doi: 10.18632/oncotarget.11141

PubMed Abstract | CrossRef Full Text | Google Scholar

Cho, A., Shim, J. E., Kim, E., Supek, F., Lehner, B., and Lee, I. (2016). MUFFINN: cancer gene discovery via network analysis of somatic mutation data. Genome Biol. 17:129. doi: 10.1186/s13059-016-0989-x

PubMed Abstract | CrossRef Full Text | Google Scholar

Ciriello, G., Cerami, E., Sander, C., and Schultz, N. (2012). Mutual exclusivity analysis identifies oncogenic network modules. Genome Res. 22, 398–406. doi: 10.1101/gr.125567.111

PubMed Abstract | CrossRef Full Text | Google Scholar

Croft, D., O'kelly, G., Wu, G., Haw, R., Gillespie, M., Matthews, L., et al. (2010). Reactome: a database of reactions, pathways and biological processes. Nucleic Acids Res. 39(suppl_1), D691–D697. doi: 10.1093/nar/gkq1018

PubMed Abstract | CrossRef Full Text | Google Scholar

Csardi, G. (2006). The igraph software package for complex network research. Interjournal Complex Syst. 1695, 1–9.

Google Scholar

Dees, N. D., Zhang, Q., Kandoth, C., Wendl, M. C., Schierding, W., Koboldt, D. C., et al. (2012). MuSiC: identifying mutational significance in cancer genomes. Genome Res. 22, 1589–1598. doi: 10.1101/gr.134635.111

PubMed Abstract | CrossRef Full Text | Google Scholar

Dimitrakopoulos, C., Kumar, H. S., Häfliger, L., Behr, J., Montazeri, H., Hall, M. N., et al. (2018). Network-based integration of multi-omics data for prioritizing cancer genes. Bioinformatics 34, 2441–2448. doi: 10.1093/bioinformatics/bty148

PubMed Abstract | CrossRef Full Text | Google Scholar

Dou, Q., Chen, H.-N., Wang, K., Yuan, K., Lei, Y., Li, K., et al. (2016). Ivermectin induces cytostatic autophagy by blocking the PAK1/Akt axis in breast cancer. Cancer Res. 76, 4457–4469. doi: 10.1158/0008-5472.CAN-15-2887

PubMed Abstract | CrossRef Full Text | Google Scholar

Gentili, M., Martini, L., Petti, M., Farina, L., and Becchetti, L. (2019). “Biological Random Walks: Integrating heterogeneous data in disease gene prioritization,” in 2019 IEEE Conference on Computational Intelligence in Bioinformatics and Computational Biology (CIBCB) (Siena), 1–8. doi: 10.1109/CIBCB.2019.8791472

CrossRef Full Text | Google Scholar

Gonzalez-Perez, A., Perez-Llamas, C., Deu-Pons, J., Tamborero, D., Schroeder, M. P., Jene-Sanz, A., et al. (2013). IntOGen-mutations identifies cancer drivers across tumor types. Nat Methods 10:1081. doi: 10.1038/nmeth.2642

PubMed Abstract | CrossRef Full Text | Google Scholar

Greenman, C., Stephens, P., Smith, R., Dalgliesh, G. L., Hunter, C., Bignell, G., et al. (2007). Patterns of somatic mutation in human cancer genomes. Eur. J. Cancer Suppl. 6, 153–158. doi: 10.1016/S1359-6349(08)71197-2

CrossRef Full Text | Google Scholar

Gui, Y., Guo, G., Huang, Y., Hu, X., Tang, A., Gao, S., et al. (2011). Frequent mutations of chromatin remodeling genes in transitional cell carcinoma of the bladder. Nat. Genet. 43, 875–878. doi: 10.1038/ng.907

PubMed Abstract | CrossRef Full Text | Google Scholar

Guo, W., Zhang, S. W., Zeng, T., Li, Y., Gao, J., and Chen, L. (2019). A novel network control model for identifying personalized driver genes in cancer. PLOS Comput. Biol. 15:e1007520. doi: 10.1371/journal.pcbi.1007520

PubMed Abstract | CrossRef Full Text | Google Scholar

Guo, W. F., Zhang, S. W., Liu, L. L., Liu, F., Shi, Q. Q., Zhang, L., et al. (2018). Discovering personalized driver mutation profiles of single samples in cancer by network control strategy. Bioinformatics 34, 2441–2448. doi: 10.1093/bioinformatics/bty006

PubMed Abstract | CrossRef Full Text | Google Scholar

Haber, D. A., and Settleman, J. (2007). Cancer: drivers and passengers. Nature 446:145–146. doi: 10.1038/446145a

PubMed Abstract | CrossRef Full Text | Google Scholar

Hidalgo, M. (2010). Pancreatic cancer. N. Engl. J. Med. 362, 1605–1617. doi: 10.1056/NEJMra0901557

PubMed Abstract | CrossRef Full Text | Google Scholar

Hirata, H., Hinoda, Y., Ueno, K., Nakajima, K., Ishii, N., and Dahiya, R. (2012). MicroRNA-1826 directly targets beta-catenin (CTNNB1) and MEK1 (MAP2K1) in VHL-inactivated renal cancer. Carcinogenesis 33, 501–508. doi: 10.1093/carcin/bgr302

PubMed Abstract | CrossRef Full Text | Google Scholar

Hou, J. P., and Ma, J. (2014). DawnRank: discovering personalized driver genes in cancer. Genome Med. 6:56. doi: 10.1186/s13073-014-0056-8

PubMed Abstract | CrossRef Full Text | Google Scholar

Hristov, B. H., and Singh, M. (2017). Network-based coverage of mutational profiles reveals cancer genes. Cell Syst. 5, 221–229.e4. doi: 10.1016/j.cels.2017.09.003

PubMed Abstract | CrossRef Full Text | Google Scholar

Iranzo, J., Martincorena, I., and Koonin, E. V. (2018). Cancer-mutation network and the number and specificity of driver mutations. Proc. Natl. Acad. Sci. U.S.A. 115, E6010–E6019. doi: 10.1073/pnas.1803155115

PubMed Abstract | CrossRef Full Text | Google Scholar

Isik, Z., and Ercan, M. E. (2017). Integration of RNA-Seq and RPPA data for survival time prediction in cancer patients. Computers Biol. Med. 89, 397–404. doi: 10.1016/j.compbiomed.2017.08.028

PubMed Abstract | CrossRef Full Text | Google Scholar

Jerhammar, F., Ceder, R., Garvin, S., Grénman, R., Grafström, R. C., and Roberg, K. (2010). Fibronectin 1 is a potential biomarker for radioresistance in head and neck squamous cell carcinoma. Cancer Biol. Therapy 10, 1244–1251. doi: 10.4161/cbt.10.12.13432

PubMed Abstract | CrossRef Full Text | Google Scholar

Jones, S., Zhang, X., Parsons, D. W., Lin, J. C., Leary, R. J., Angenendt, P., et al. (2016). Core signaling pathways in human pancreatic cancers revealed by global genomic analyses. Science 321, 1801–1806. doi: 10.1126/science.1164368

PubMed Abstract | CrossRef Full Text | Google Scholar

Kanehisa, M., Goto, S., Sato, Y., Furumichi, M., and Tanabe, M. (2011). KEGG for integration and interpretation of large-scale molecular data sets. Nucleic Acids Res. 40, D109–D114. doi: 10.1093/nar/gkr988

PubMed Abstract | CrossRef Full Text | Google Scholar

Klaus, A., and Birchmeier, W. (2008). Wnt signalling and its impact on development and cancer. Nat. Rev. Cancer 8, 387–398. doi: 10.1038/nrc2389

PubMed Abstract | CrossRef Full Text | Google Scholar

Köhler, S., Bauer, S., Horn, D., and Robinson, P. N. (2008). Walking the interactome for prioritization of candidate disease genes. Am. J. Hum. Genet. 82, 949–958. doi: 10.1016/j.ajhg.2008.02.013

PubMed Abstract | CrossRef Full Text | Google Scholar

Kumar, R., Gururaj, A. E., and Barnes, C. J. (2006). p21-activated kinases in cancer. Nat. Rev. Cancer 6, 459–471. doi: 10.1038/nrc1892

PubMed Abstract | CrossRef Full Text | Google Scholar

Lawrence, M. S., Petar, S., Paz, P., Kryukov, G. V., Kristian, C., Andrey, S., et al. (2013). Mutational heterogeneity in cancer and the search for new cancer genes. Nature 499, 214–218. doi: 10.1038/nature12213

PubMed Abstract | CrossRef Full Text | Google Scholar

Lee, I., Blom, U. M., Wang, P. I., Shim, J. E., and Marcotte, E. M. (2011). Prioritizing candidate disease genes by network-based boosting of genome-wide association data. Genome Res. 21, 1109–1121. doi: 10.1101/gr.118992.110

PubMed Abstract | CrossRef Full Text | Google Scholar

Liu, M., Xiong, Z., Ma, Y., Zhang, P., Wu, J., and Qi, X. (2017). DPRank centrality: finding important vertices based on random walks with a new defined transition matrix. Fut. Gen. Comp. Syst. 83, 376–389. doi: 10.1016/j.future.2017.10.036

CrossRef Full Text | Google Scholar

Moreau, Y., and Tranchevent, L.-C. (2012). Computational tools for prioritizing candidate genes: boosting disease gene discovery. Nat. Rev. Gen. 13, 523–536. doi: 10.1038/nrg3253

PubMed Abstract | CrossRef Full Text | Google Scholar

Network, C. G. A. (2008). Comprehensive genomic characterization defines human glioblastoma genes and core pathways. Nature 455, 1061–1068. doi: 10.1038/nature07385

CrossRef Full Text | Google Scholar

Nussinov, R., Jang, H., Tsai, C.-J., and Cheng, F. (2019). Review: precision medicine and driver mutations: computational methods, functional assays and conformational principles for interpreting cancer drivers. PLoS Comput. Biol. 15:e1006658. doi: 10.1371/journal.pcbi.1006658

CrossRef Full Text | Google Scholar

O'Sullivan, G. C., Tangney, M., Casey, G., Ambrose, M., Houston, A., and Barry, O. P. (2007). Modulation of p21-activated kinase 1 alters the behavior of renal cell carcinoma. Int. J. Cancer 121, 1930–1940. doi: 10.1002/ijc.22893

PubMed Abstract | CrossRef Full Text | Google Scholar

Page, L., Brin, S., Motwani, R., and Winograd, T. (1999). The PageRank Citation Ranking: Bringing Order to the Web. Stanford, CA: Stanford InfoLab.

Google Scholar

Petitjean, A., Achatz, M., Borresen-Dale, A.-L., Hainaut, P., and Olivier, M. (2007). TP53 mutations in human cancers: functional selection and impact on cancer prognosis and outcomes. Oncogene 26, 2157–2165. doi: 10.1038/sj.onc.1210302

PubMed Abstract | CrossRef Full Text | Google Scholar

Piñero, J., Bravo, À., Queralt-Rosinach, N., Gutiérrez-Sacristán, A., Deu-Pons, J., Centeno, E., et al. (2016). DisGeNET: a comprehensive platform integrating information on human disease-associated genes and variants. Nucleic Acids Res. 45, D833–D839. doi: 10.1093/nar/gkw943

PubMed Abstract | CrossRef Full Text | Google Scholar

Raphael, B. J., Dobson, J. R., Oesper, L., and Vandin, F. (2014). Identifying driver mutations in sequenced cancer genomes: computational approaches to enable precision medicine. Genome Med. 6, 1–17. doi: 10.1186/gm524

PubMed Abstract | CrossRef Full Text | Google Scholar

Repana, D., Nulsen, J., Dressler, L., Bortolomeazzi, M., Venkata, S. K., Tourna, A., et al. (2019). The Network of Cancer Genes (NCG): a comprehensive catalogue of known and candidate cancer genes from cancer sequencing screens. Genome Biol. 20:1. doi: 10.1186/s13059-018-1612-0

PubMed Abstract | CrossRef Full Text | Google Scholar

Rosenbloom, K. R., Armstrong, J., Barber, G. P., Casper, J., Clawson, H., Diekhans, M., et al. (2015). The UCSC Genome Browser database: 2015 update. Nucleic Acids Res. 43, 670–681. doi: 10.1093/nar/gku1177

PubMed Abstract | CrossRef Full Text | Google Scholar

Schaefer, C. F., Anthony, K., Krupa, S., Buchoff, J., Day, M., Hannay, T., et al. (2008). PID: the pathway interaction database. Nucleic Acids Res. 37(suppl_1), D674–D679. doi: 10.1038/npre.2008.2243.1

PubMed Abstract | CrossRef Full Text | Google Scholar

Schou, I., Ekeberg, Ø., and Ruland, C. M. (2001). Genetic alteration of the beta-catenin gene (CTNNB1) in human lung cancer and malignant mesothelioma and identification of a new 3p21.3 homozygous deletion. Oncogene 20, 4249–4257. doi: 10.1038/sj.onc.1204557

CrossRef Full Text | Google Scholar

Sondka, Z., Bamford, S., Cole, C. G., Ward, S. A., Dunham, I., and Forbes, S. A. (2018). The COSMIC cancer gene census: describing genetic dysfunction across all human cancers. Nat. Rev. Cancer 18, 696–705. doi: 10.1038/s41568-018-0060-1

PubMed Abstract | CrossRef Full Text | Google Scholar

Song, J., Peng, W., and Wang, F. (2019). A random walk-based method to identify driver genes by integrating the subcellular localization and variation frequency into bipartite graph. BMC Bioinformatics 20:238. doi: 10.1186/s12859-019-2847-9

PubMed Abstract | CrossRef Full Text | Google Scholar

Stratton, M. R., Campbell, P. J., and Futreal, P. A. (2009). The cancer genome. Nature 458:719. doi: 10.1038/nature07943

PubMed Abstract | CrossRef Full Text | Google Scholar

Szklarczyk, D., Franceschini, A., Wyder, S., Forslund, K., Heller, D., Huertacepas, J., et al. (2015). STRING v10: protein-protein interaction networks, integrated over the tree of life. Nucleic Acids Res. 43, 447–452. doi: 10.1093/nar/gku1003

PubMed Abstract | CrossRef Full Text | Google Scholar

Vandin, F., Upfal, E., and Raphael, B. J. (2012). De novo discovery of mutated driver pathways in cancer. Genome Res. 22, 375–385. doi: 10.1101/gr.120477.111

PubMed Abstract | CrossRef Full Text | Google Scholar

Vogelstein, B., Papadopoulos, N., Velculescu, V. E., Zhou, S., Diaz, L. A., and Kinzler, K. W. (2013). Cancer genome landscapes. Science 339, 1546–1558. doi: 10.1126/science.1235122

PubMed Abstract | CrossRef Full Text | Google Scholar

Vokes, E. E., Agrawal, N., and Seiwert, T. Y. (2015). HPV-associated head and neck cancer. J. Natl. Cancer Inst. 107:djv344. doi: 10.1093/jnci/djv344

PubMed Abstract | CrossRef Full Text | Google Scholar

Wu, H., Gao, L., Li, F., Song, F., Yang, X., and Kasabov, N. (2015). Identifying overlapping mutated driver pathways by constructing gene networks in cancer. BMC Bioinformatics 16, 1–11. doi: 10.1186/1471-2105-16-S5-S3

PubMed Abstract | CrossRef Full Text | Google Scholar

Zhang, J., Wu, L., Zhang, X., and Zhang, S. (2014). Discovery of co-occurring driver pathways in cancer. BMC Bioinformatics 15, 271–271. doi: 10.1186/1471-2105-15-271

PubMed Abstract | CrossRef Full Text | Google Scholar

Zhang, J., and Zhang, S. (2017). Discovery of cancer common and specific driver gene sets. Nucleic Acids Res. 45:e86. doi: 10.1093/nar/gkx089

PubMed Abstract | CrossRef Full Text | Google Scholar

Zhang, J., and Zhang, S. (2018). The discovery of mutated driver pathways in cancer: models and algorithms. IEEE/ACM Trans. Comput. Biol. Bioinformatics 15, 988–998. doi: 10.1109/TCBB.2016.2640963

PubMed Abstract | CrossRef Full Text | Google Scholar

Zhang, W., Xu, J., Li, Y., and Zou, X. (2018). Detecting essential proteins based on network topology, gene expression data, and gene ontology information. IEEE/ACM Trans. Comput. Biol. Bioinformatics 15, 109–116. doi: 10.1109/TCBB.2016.2615931

CrossRef Full Text | Google Scholar

Keywords: cancer, driver gene, protein–protein network, random walk, centrality

Citation: Wei P-J, Wu F-X, Xia J, Su Y, Wang J and Zheng C-H (2020) Prioritizing Cancer Genes Based on an Improved Random Walk Method. Front. Genet. 11:377. doi: 10.3389/fgene.2020.00377

Received: 13 December 2019; Accepted: 26 March 2020;
Published: 28 April 2020.

Edited by:

Alfredo Pulvirenti, University of Catania, Italy

Reviewed by:

Guosheng Han, Xiangtan University, China
Qinghua Jiang, Harbin Institute of Technology, China

Copyright © 2020 Wei, Wu, Xia, Su, Wang and Zheng. This is an open-access article distributed under the terms of the Creative Commons Attribution License (CC BY). The use, distribution or reproduction in other forums is permitted, provided the original author(s) and the copyright owner(s) are credited and that the original publication in this journal is cited, in accordance with accepted academic practice. No use, distribution or reproduction is permitted which does not comply with these terms.

*Correspondence: Chun-Hou Zheng, zhengch99@126.com

Disclaimer: All claims expressed in this article are solely those of the authors and do not necessarily represent those of their affiliated organizations, or those of the publisher, the editors and the reviewers. Any product that may be evaluated in this article or claim that may be made by its manufacturer is not guaranteed or endorsed by the publisher.